-
Notifications
You must be signed in to change notification settings - Fork 0
/
Copy pathDynamics_IonFlux.tex
3052 lines (2547 loc) · 116 KB
/
Dynamics_IonFlux.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{Dynamics of Ion Flux}
\section{Concentrations}
\label{sec:concentrations}
In previous chapter, we use the concept of {\it 'fraction of open channel'}
(Sect.\ref{sec:concentration-example}) to map from a discontinued quantity (i.e.
the number of channel) to a continuous quantity (i.e. fraction in the range 0 to
1). In a generic context of chemical species, an equivalent concept called {\bf
concentration} is used.
Concentration is a way to describe mixture composition, e.g. the
amount of a solvent in a solution. However, this doesn't necessary
liquid ones.
\subsection{mass (weight/weight) percentage: grams of substance per 100 grams of
sample}
\label{sec:concentration-mass-percentage}
One way is to use mass (weight) percentage (\% w/w)
\begin{equation}
\label{eq:999}
\text{c}_{\% w/w} = \frac{m_\text{substance}}{m_\text{solution}} \times 100\%
\end{equation}
In simple case when solution contains only one solvent and one solute,
then $m_\text{solution} = m_\text{solute}+m_\text{solvent}$.
IMPORTANT: Only weight percentage is the percentage concentration that is always
unambiguous, i.e. does not change with temperature. Unit : (\%).
\begin{verbatim}
1% w/w means 1 gram of substance per every 100gram of sample
\end{verbatim}
\subsection{mass-volume percentage, volume-volume percentage}
Unlike weight-weight percentage (Sect.\ref{sec:concentration-mass-percentage}),
mass-volume percentage or volume-volume percentage are temperature dependent, so
such quantities are unambiguous.
{\bf Mass-volume (or weight-volume) percentage} (\% w/v) is
\begin{equation}
\label{eq:1000}
\text{c}_{\% w/v} = \frac{m_{substance}}{V_{solution}} \times 100\%
\end{equation}
the unit of Volume is [mL], and unit of mass is [g]. So, the overall
unit is in [\% g/mL]. So, it's possible to have solutions of
concentration above 100\% w/v.
{\bf Volume-volume percentage} (\% v/v) is
\begin{equation}
\label{eq:1001}
\text{c}_{\% v/v} = \frac{V_{substance}}{V_{solution}} \times 100\%
\end{equation}
The problem with this usage is that ``volume is only additive for
ideal gas only''. So, the final volume is not a sum of volumes used
when preparing mixture. It means that volume-volume percentage doesn't
sum to 100\%. To convert between concentration types, we need to know
densities of the solution and of both pure solvent and solute.
\subsection{parts per million/billion: ppm, ppb}
For very low concentration, people often use {\bf ``parts per''}
notation.
NOTE: \verb!1% w/w! above can be understood as ``1 mass part per
hundred mass parts'', though very rarely used. Instead, these are
often used
\begin{verbatim}
ppm = parts per million (10^6)
ppb = parts per billion (10^9)
ppt = parts per trillion (10^12)
ppq = parts per quadrillion (10^15)
\end{verbatim}
``ppq'' is rather a theoretical construct than a useful thing. It can
also use volume by volume, rather than mass by mass.
Example: 1mL of \ce{SF6}(g) added to 1000L of hydrogen \ce{H2}(g) is 1ppm
vol/vol, but 73 ppm mass/mass. Another example: 1 atom of lead per $10^6$
beryllium atoms there are 23 ppm of lead mass/mass.
\subsection{molar concentration: M (or mole/(Liter solution)), mM, $\mu$M, nM}
\label{sec:concentration-Molar}
{\bf Molar concentration} (M, mM, $\mu$M, nM) or {\bf molarity} is
\begin{equation}
\label{eq:1002}
c_M = \frac{n_{substance}}{V_{solution}}
\end{equation}
with unit is [mole/L] or [Molar] or [M].
NOTE: Some old books use M/500 unit, it means 1 mol per 500 litres of solution
(or M/500 means 0.002 mol/L).
\textcolor{red}{Molar concentration is the most often used concentration unit as
it makes stoichiometric calculations much easier}, which can be measured with
the highest precision and for most analytical applications this is the preferred
way of expressing concentration.
{\it The only drawback is that it's temperature-dependent}, e.g. the volumes
increase when the temperature increase and molar concentration goes down.
To avoid temperature dependency, we can use {\bf molarity}
(Sect.\ref{sec:concentration-molarity}).
\subsection{-- molarity (mole/kg)}
\label{sec:concentration-molarity}
Unlike {\it molar concentration} concept (Sect.\ref{sec:concentration-Molar});
{\bf molarity} is temperature independent.
\begin{equation}
\label{eq:1003}
c_m = \frac{n_{substance}}{m_{solvent}}
\end{equation}
is expressed in [mole/kg] units.
NOTE: It's rarely used in analytical chemistry, molality is often used in {\bf
physical chemistry}, especially when dealing with substance properties over a
wide range of temperature, or properties change with both temperature and
mixture composition.
\subsection{-- normality}
\label{sec:concentration-normality}
{\bf Normality} (similar to molarity, but use equivalent, not moles) has the
units number of [equivalents per Liter].
So, the same solution can have different normalities for different types of
reactions. So, it's \textcolor{red}{reaction-dependent}. E.g.: 1M sulfuric acid
solution is 2N for acid/base reactions, and 1N in the reaction of barium sulfate
precipitation.
\begin{framed}
Molar mass of the substance
\begin{equation}
\label{eq:1005}
n_\text{substance} = \frac{m_\text{substance}}{m_M}
\end{equation}
with $n_\text{substance}$ is the number of moles.
\end{framed}
\subsection{Molar fraction}
\label{sec:concentration-molar-fraction}
{\bf Molar fraction} is the number of moles of a substance divided by
the total number of moles of all substances. So the range of a value for molar
fraction is from 0 to 1 and {\bf unitless}. The quantity is temperature
independent.
Conversion:
\begin{enumerate}
\item from weight percentage to molality
\begin{equation}
\label{eq:1004}
c_m = \frac{1000\times c_{\% w/v/}}{m_M(100-c_{\% w/w})}
\end{equation}
the factor 1000 is necessary as molality is in [mole/kg] while all
masses are in grams (gr).
\end{enumerate}
\subsection{protein concentration (mg/ml)}
\label{sec:concentration-protein}
In biochemistry, protein concentration is measured in mg/ml, and is measured
using Bradford method (Coomassie Blue reagent; Pierce Chemical
Co., Rockford, IL) with bovine serum albumin (Sigma Chemical, Co.) as standard.
Reference:
\begin{itemize}
\item \url{http://www.chembuddy.com/?left=concentration&right=concentration-follies}.
\end{itemize}
\subsection{Faraday constant}
\label{sec:Faraday-constant}
The particles are often electric charges, whose unit is
Coulomb (C). Definition:
\begin{itemize}
\item {\it 1 Coulomb is the amount of charge that must pass through an
electrolytic cell in order to deposit $1.1180\times 10^3$g of silver
from a solution of silver nitrate}, or
\item 1 Coulomb is the amount of charge on $6.242\times 10^{18}$
electrons.
\end{itemize}
So, the charges on one mole of electrons (which is Avogadro number
$N_A=6.023\times 10^{23}$ of electrons) is
\begin{equation}
\label{eq:1343}
\frac{N_A}{6.242\times 10^{18}} = 9.649\times 10^4 (\C/\text{mol. electrons})
\end{equation}
This quantity is known as {\bf Faraday constant}
\begin{equation}
\label{eq:1344}
1\text{ Faraday } = 9.649\times 10^4 \;\;(\C/\text{mol. electrons})
\end{equation}
Faraday's constant converts quantity of moles to quantity of charge for a
univalent ion, i.e. $z=1$.
\section{Flux $J$}
\label{sec:fluxes}
A flux describes the rate of a reaction, i.e. the speed of reaction
(Sect.\ref{sec:speed-reaction}). In the previous example, we studied the simple
example of the rate of switching between open and close of a cluster of $N$ ion
channels (Sect.\ref{sec:flux-example}).
In chemical reactions, the fraction of 'open' channel in that example is
replaced by the 'concentration' of a given species
(Sect.\ref{sec:concentration}), and the {\bf chemical flux} is used to study the
rate of chemical reactions (Sect.\ref{sec:chemical-flux}).
% \subsection{Current density}
%
% To avoid the variation between cells, all quantities are determined
% based on a unit of membrane area, except the voltage. The information
% is given in Table~\ref{tab:terminology_2}.
%
% \begin{table}[hbt]
% \begin{center}
% \caption{Ionic currents}
% \begin{tabular}{p{2cm}r}
% \hline
% notation & description \\
% \hline\hline
% $I_i$ & ionic current through plasma membrane ($I_{\ce{Na}},
% I_{\ce{K}}$) \\
% & ...$I_i = g_i(V_m-E_i) $ \\
% $I_{app}$ & applied step current (rectangular pulse) (mA/cm$^2$)\\
% $I_m=\frac{I_0}{A}$ & the current density per unit area (mA.cm$^{-2}$) \\
% $i_m$ & the current density per unit length (mA/cm) \\
% $I_\dhpr, I_\LCC$, $I_\Ca$ & the calcium current via DHPR (L-type channels) \\
% $I_\ryr$ & the current of calcium released via RyR from JSR \\
% $I_\Na$ & the current of sodium \\
% $I_\NaCa$ & the current via Na/Ca exchanger \\
% $I_\K$ & the current via potassium channel \\
% \end{tabular}
% \end{center}
% \label{tab:Current}
% \end{table}
%
% With different types of ions, there are respectively different ionic
% currents, Table \ref{tab:Current}. Typically, all currents are indeed
% ``current density'', i.e. in unit [Ampere/unit square], e.g. mA/cm$^2$.
\subsection{Atomic vibrations: thermal energy}
\label{sec:thermal-energy-distribution}
\label{sec:atomic-electronic-energy}
\label{sec:electronic-energy}
Heat causes atoms to vibrate. The vibration amplitude increases with
temperature.
When vibration are sufficient large to break the bonds, a material is
disruptly transformed from one state to another, e.g. from solid to liquid.
\begin{itemize}
\item vibration frequency: $10^{13}$ Hz
\item heat provides the atoms an amount of energy, called
atomic/electronic energy.
Heat causes atomic vibration which generate an amount of energy, called
atomic/electronic energy.
The average atomic/electronic energy due to thermal excitation is of order of
$k_B T$, i.e. $E_\text{average} \sim k_BT$ (with $k_B$ = Boltzmann constant,
and $T$ = temperature in absolute degree Kelvin).
NOTE: $k_BT = 0.026$ (eV) for room temperature.
The distribution of such energy around the mean follow the Boltzmann
distribution function (Sect.\ref{sec:Boltzmann-equation})
\begin{equation}
P(E) = \exp\left( -E/(k_BT)\right)
\end{equation}
$k_B = 1.38 \times 10^{-23} $(J/K) = $8.62\times 10^5 $ (eV/K).
\end{itemize}
\subsection{Atomic diffusion: vacancy diffusion (interdiffusion and
self-diffusion)}
Diffusion is material transport by atomic motion from one lattice site to
another. To move from lattice site to lattice site, atoms need energy to break
bonds with neighbors, and to cause the necessary lattice distortions during
motion from site to another. This energy comes from atomic vibration as
mentioned above.
This type of diffusion, i.e. vacancy diffusion, depends on the presence of
vacancies and therefore increases with the vacancy concentration as the
temperature increases. Motion of vacancies in one direction is equivalent to
motion of atoms in the opposite direction.
Interdiffusion occurs in response to a concentration gradient (more rigorously,
to a gradient in chemical potential).
\begin{figure}[hbtp]
\centerline{\includegraphics[height=5cm,
angle=0]{./images/motion.eps}}
\caption{(A) vacancy diffusion; (B) interstitial diffusion}
\label{fig:motion}
\end{figure}
\subsection{Atomic diffusion: interstitial diffusion (interdiffusion only)}
Interstitial diffusion (depends on temperature). This is generally faster than
vacancy diffusion because there are many more interstitial sites than vacancy
sites to jump to.
\subsection{Probability of energy jump}
\begin{equation}
R_j = R_0 \exp \left( - \frac{E_m}{k_BT} \right)
\end{equation}
\subsection{Diffusion flux (J)}
Diffusion is process which is NOT due to the action of a force , but a result of
the random movements of atoms ( statistical problem).
The flux of diffusing atoms, J, is used to quantify how fast diffusion occurs.
The flux is defined as either in
\begin{enumerate}
\item number of atoms diffusing through unit area and
per unit time, i.e. atoms/(m$^2$.second).
\item mass flux: mass of atoms diffusing through
unit area per unit time, (e.g., kg/(m$^2$.second))
\end{enumerate}
\begin{equation}
J = \frac{M}{A.t} = \frac{\text{mass (or atoms or
moles)}}{\text{(cross-sectional area) x time}}
\end{equation}
unit of J (g/(cm$^2$.sec)) if M is mass (g).
In differential form, the instantaneous flux $J$ is
\begin{equation}
J(t) = \frac{1}{A} \frac{\partial M}{\partial t}
\end{equation}
\subsection{Steady-state diffusion: Fick's first law}
In 1D, at steady state of an isothermal, isobaric system, the diffusion along
the direction $x$ is proportional to the concentration gradient.
\begin{equation}
J = - D \frac{dC}{dx}
\end{equation}
As the slope of the change from high to low concentration
\begin{equation*}
dC/dx = \frac{c_\text{dest}-c_\text{source}}{dx} < 0
\end{equation*}
as $c_\text{dest}=c_\text{low}$ < $c_\text{source}=c_\text{high}$, the minus sign indicate the flux
get positive value and flows in the direction down the concentration gradient.
$D$ is the diffusion coefficient (Sect.\ref{sec:diffusion-coefficient}), and is
typically chosen as a constant as a given temperature.
\subsection{Diffusion coefficient: Diffusion constant}
\label{sec:diffusion-constant}
\label{sec:diffusion-coefficient}
Diffusion coefficient is the measure of mobility of diffusing species.
At steady-state, the diffusion coefficient is considered a constant value at a
given temperature.
\begin{equation}
D = D_0 \exp \left( - \frac{Q_d}{RT} \right)
\end{equation}
with
\begin{itemize}
\item $D_0 =$ temperature-independent pre-exponential (m$^2$/sec);
\begin{equation}
D_0 = \frac{k_BT}{h}
\end{equation}
with $h = $ Planck's constant ($6.6 \times 10^{-34}$ J/sec), and $k_B = $
\item $Q_d$ = activation energy for diffusion (J/mole, or eV/atom)
\item $R = $ universal gas constant ()
\item $T = $ temperature (K)
\end{itemize}
If we take the log on both sides
\def\ln{{\text{ln}}}
\begin{equation}
\ln D = \ln D_0 - \frac{Q_d}{R} \left( \frac{1}{T} \right)
\end{equation}
or
\def\log{{\text{log}}}
\begin{equation}
\log D = \log D_0 -\frac{Q_d}{2.3 R} \left( \frac{1}{T} \right)
\end{equation}
\subsection{Nonsteady-state diffusion: Fick's second law}
In most real situations the concentration profile and the the diffusion flux and
the concentrations change with time, i.e. the concentration gradient are
changing with time, Fig.\ref{fig:diffusion-in-time-space}.
\begin{equation}
\frac{\partial C}{\partial t} = \frac{\partial}{\partial t}\left( D
\frac{\partial C}{\partial x} \right) = D \frac{\partial^2 C}{\partial x^2}
\end{equation}
So the time-varying diffusion follows the so-called Fick's second law.
\begin{figure}[hbtp]
\centerline{\includegraphics[height=5cm,
angle=0]{./images/diffusion-in-time-space.eps}}
\caption{Diffusion in time and space}
\label{fig:diffusion-in-time-space}
\end{figure}
Diffusion constant $D$ (distance$^2$/time) is a quantity that tell how fast a
species is translocated from one place to another based on Fickian Law
(deterministic). For a given distance $x$ (distance), the time for movement
is defined as
\begin{equation}
t = \frac{x^2}{2*D}
\end{equation}
The time-constant for this transfer is known as the reciprocal of the time,
$\tau$ (sec$^{-1}$)
\begin{equation}
\tau = \frac{1}{t}
\end{equation}
\begin{framed}
A connection between ionic mobility $u$ and diffusion constant $D$
is
\begin{equation}
\label{eq:1227}
D = \frac{u.R.T}{|z|F}
\end{equation}
with T = absolute temperature [K], R = Faraday constant [J/(mol.K)]
\end{framed}
\section{Chemical flux: Fluxes in aqueous media}
\label{sec:chemical-flux}
For (transmembrane) flux of currents via (selective pores) ion channels, let's
check Sect.\ref{sec:equat-ionic-curr}. In this section, we discuss flux across
two compartments
\ref{sec:flux-example}
\ref{sec:fluxes}
In transport phenomena, a flux is defined as the amount of a substance
that flows through a unit area (or unit volume) per a unit of
time. Generally, the amount is in term of density (e.g. mass per unit
volume or mass per unit surface area), so a flux is a vector. In
chemical diffusion, the unit of flux
\begin{itemize}
\item can be (mol.cm$^{-2}$.s$^{-1}$)
\item or (mol.cm$^{-3}$.s$^{-1}$)
\item or (mol.L.s$^{-1}$) or ($\mu$M/sec).
\end{itemize}
We typically use \verb!cm! rather than metre in chemical diffusion.
There are 4 physical laws that dictates the movement of ions
\begin{enumerate}
\item drift movement under electric field (Sect.~\ref{sec:drift-ions})
\item diffusion of ions under concentration
gradient (Sect.~\ref{sec:ionic-diffusion})
\item the relationship between the two above processes, i.e. the
diffusion coefficient $D_c$ and the drift mobility $\mu$
(Sect.~\ref{sec:einstein-relation})
\item the basic principle of separation of charges in biological systems, e.g.
if there is any barrier such as the plasma membrane that separate the two spaces
\end{enumerate}
These laws lead to fundamental equations in computational cell
biology: Nernst-Planck, Nernst, Goldman-Hodgkin-Katz equations, and
Donnan equilibrium equation.
\subsection{Molar flux: Nernst-Planck equation}
\label{sec:flux-molar-flux}
\label{sec:Nernst-Planck-equation}
\textcolor{red}{Effect 1}: The molar flux for diffusive properties in terms of
the diffusion coefficient ($D_s$) and the local concentration of the property
($c_s$). This is Fick's law (Sect.\ref{sec:fickian-diffusion}).
\begin{equation}
M'_s = -D_s \frac{dc_s}{dx}
\end{equation}
\textcolor{red}{Effect 2}: The molar flux of solute due to electrophoretic
effects is defined in terms of the concentration of s (i.e. $c_s$), the
electrical mobility ($u_s$ - the velocity of S in the fluid), and the local
potential ($\Psi$).
\begin{equation}
M'_s = - u_s c_s \frac{d\Psi}{dx}
\end{equation}
The two affects above are fundamentally additive, i.e.
\begin{equation}
M_s = -D_s \frac{dc_s}{dx} - u_s c_s \frac{d\Psi}{dx}
\end{equation}
This is known as {\bf Nernst-Planck equation} (after Nernst 1888,1889; Planck,
1890). It extends Fick's law for the case where the diffusing particles are also
moved with respect to the fluid by electrostatic forces.
\begin{equation}
\begin{split}
\frac{\partial c_s}{\partial t} &= -\nabla(J) \\
J &= - \left[ D \nabla c_s - u c_s + \frac{Dz_s e}{k_BT} c_s (\nabla \Psi +
\frac{\partial A}{\partial t}) \right]
\end{split}
\end{equation}
Its dimension depend on those used to express the ionic concentration an
velocity. NORMALLY: The unit is mol/(cm$^2$.c).
SUMMARY: The total ionic flux for the \verb!k!-th ion type, $\bar{j}_k$, is
given by the sum of ionic fluxed due to (1) diffusion (D), and (2) electric
field (e).
Using {\bf Einstein relationship}
\begin{equation}
\bar{j}_k = \bar{j}_{k,D} + \bar{k}_{k,e} = -D_k \left( \nabla c_k +
\frac{c_k z_k F}{RT} \nabla \Phi \right)
\end{equation}
\begin{mdframed}
Ion flows from one place to another in an aqueous media can be the result from
two driving forces: from an electric field force
(Sect.\ref{sec:flow-by-electric-field-force}) and those resulting from a
diffusional force (Sect.\ref{sec:flow-by-diffusional-force}). For the flow of
ions across a barrier, such as plama membrane, read
Sect.\ref{sec:equat-ionic-curr}.
The {\bf Nernst-Planck equation} describes the flux of an ion (charged
chemical species) in a fluid environment under the affect of diffusion and
electric field.
\begin{equation}
\label{eq:1228}
\overrightarrow{j} = \overrightarrow{j_D} + \overrightarrow{j_e} =
-D (\nabla c + \frac{c.z.F}{R.T}\Delta \Phi)
\end{equation}
with $\overrightarrow{j}$ = ionic flux [mol/(cm$^2$.sec)].
\end{mdframed}
However, what we care about is the current across the membrane, not the molar
flux. It is discussed in Sect.\ref{sec:ionic-flow}.
\subsection{Nerns-Einstein relationship}
\label{sec:Einstein-relationship}
Nernst-Einstein relationship defines
\begin{equation}
D_s = \frac{u_s RT}{z_s F}
\end{equation}
with $z_s$ is the valence of S.
\subsection{Ionic flows}
\label{sec:ionic-flow}
In order to get from the molar flux (Sect.\ref{sec:flux-molar-flux}) to the
current we need only multiply by the valence and Faraday's constant
\begin{equation}
I_s = z_s.F.M_s
\end{equation}
\ref{sec:nernst-equation}
We can easily convert the ionic flux to the electric current density $I$ by
multiplying the ionic flux by $zF$ (i.e. the number of charges carried
by each mole).
The ionic flux $\bar{j}_k$ can be converted to an electric current density
$I$ by multiplying the former by \verb!z_k.F!
\begin{equation}
I_k = \bar{j}_k \times z.F = -D_k z_k F \left( \nabla c_k +
\frac{c_k z_k F}{RT} \nabla \Phi \right)
\end{equation}
The unit of electric current density $J_{k}$ is [C/(cm$^2$.s)] = [A/cm$^2$].
The equivalent form is
\begin{equation}
I_k = - \left( \mu_k R T \frac{z_k}{|z_k|} \nabla c_k + u_k c_k
|z_k| F. \nabla \Phi \right)
\end{equation}
\url{http://www.bem.fi/book/03/03.htm}
\begin{mdframed}
\begin{equation}
\label{eq:1242}
\begin{split}
I &= -DzF (\nabla c + \frac{c.z.F}{R.T}\Delta \Phi) \\
&= - \left(u.R.T.\frac{z}{|z|}\nabla c + u.c.|z|F\nabla\Phi\right)
\end{split}
\end{equation}
with $I$ = electric current density [C/(sec.cm$^2$)] = [A/cm$^2$].
\end{mdframed}
By integrating the Nernst-Planck equation across the cell membrane -
the center of the membrane is taken as the origin of coordinates for a membrane
of thickness $h$ - for a steady-state and arbitrary potential profile, the
diffusive flux across the cell membrane is (Jacquez, Schultz, 1974)
\begin{equation}
I_k^d = \frac{-h P_k}{\int_{-h/2}^{h/2} \exp\left( z_k F \psi(x) /(RT)
\right)dx}
\left[ c_k^i Z^{z_k}_i \exp\left(\frac{z_k F V}{2 RT} \right)
- c_k^o Z^{-z_k}_o \exp\left( -\frac{z_k F V}{2 RT} \right)
\right]
\end{equation}
NOTE:
\begin{enumerate}
\item $P_k$ = permeability of ion $k$ - Sect.\ref{sec:permeability}
\item $z_k$ = valence of ion $k$
\item $\psi(x)$ = the potential at point $x$ along the membrane
NOTE: $\psi(-h/2)$ = from the outside at $-h/2$;
$\psi(h/2)$ = from the inside at $h/2$
\item $c_k^i, c_k^o$ = the concentration (activity) of ion $k$ at
intracellular and outer-side of membrane, respectively.
\item $Z$ factor is defined as
\begin{equation}
Z_i = \exp (F V_i /(RT)) \qquad;
Z_o = \exp (F V_o / (RT))
\end{equation}
with $V_i = \psi_i - \psi(h/2)$; and $V_o = \psi(-h/2) - \psi_o$ - which are
potential differences between inner and outer bulk phases with the adjacent
membrane surfaces.
NOTE: The transmembrane potential $\Vm = \psi_i - \psi_o = V_i + V + V_o$
\end{enumerate}
\subsection{- caused by electric field force}
\label{sec:flow-by-electric-field-force}
\label{sec:electric-field}
We consider two points: O (with reference potential $\Phi_O$) and P (with
potential $\Phi_P$). So, the work required to move one unit charge
is
\begin{equation}
\label{eq:work-move-1-molecule-ion}
W_e = (\Phi_P - \Phi_O)
\end{equation}
Whenever a force exerts on a unit charge, an electric field is defined, in the
opposite direction of the translocation. The {\bf electric field} is defined as
{\it the force that exerts on a unit charge}.
The work $W_e$ required to move Q quantity of charge from
point O to point P is
\begin{equation}
W_e = Q (\Phi_P - \Phi_O)
\end{equation}
with $W_e$ [Joules, J], $\Phi$ [Volt], Q [Coulomb, C]).
\begin{mdframed}
Instead of using Coulomb, the quantity of ions in electrophysiology is often
used as {\it moles}. Suppose the valence of an ion is $z$ and the change on one
mole of electrons is $F$ (Faraday constant), then the work required to move one
mole of an ion (i.e. an $N_A$ number of molecules of
that ion) of valence $z$ [unitless] from point O to point P is
\begin{equation}
W_e = zF (\Phi_P - \Phi_O)
\end{equation}
NOTE: for $\Ca$ ion, $z_\Ca = 2$.
\end{mdframed}
The work required to move a unit positive charge (i.e. $Q=1$) from O to P (of
distance $ds$) against the electric field $\vec{E}$ is defined as
\begin{equation}
dW = - \vec{E} \cdot \vec{ds}
\end{equation}
with $\vec{ds}$ is the vector of displacement.
Combined with eq.\ref{eq:work-move-1-molecule-ion}, we have
\begin{equation}
\Phi_P - \Phi_O = - \vec{E} \cdot \vec{ds}
\end{equation}
Applying Taylor series expansionof the scalar field $\Phi$ around the point O
along the path $s$
\begin{equation}
\Phi_P = \Phi_O + \frac{d\Phi}{ds} ds + \ldots
\end{equation}
The second term is known as {\it directional derivative of $\Phi$ in the
direction $s$} and by vector-analytic properties of the gradient is given by
$\nabla \Phi \cdot \vec{ds}$.
If P is very close to O, the remaining higher terms may be neglected. So, we
have $\Phi_P - \Phi_O = \nabla \Phi \cdot \vec{ds}$.
\subsection{- caused by diffusional force}
\label{sec:flow-by-diffusional-force}
If the ionic concentration is not uniform between adjacent compartments, this
leads to a passive redistribution of ionic concentration, which can be described
using {\bf Fickian diffusion} formula (Sect.\ref{sec:ficks-law-diffusion})
\begin{equation}
\label{eq:1226}
\overrightarrow{j_D} = -D\nabla c
\end{equation}
with D = diffusion constant [cm$^2$/sec] (Sect.\ref{sec:diffusion-constant}), c
= ionic concentration [mol/cm$^3$] (or Molar=mole/litre), $\overrightarrow{j_D}$
[mol/(cm$^2$.sec)].
\subsection{-- ((((()))))}
In a metabolism process, the product C of one reaction can be the input (i.e.
the reactant) of another one. Hence, the concentration of C, i.e. [C], can
change due to either the consumption of C or the production of C.
Sect.\ref{sec:concentration} describes how a concentration is defined.
\begin{equation}
\label{eq:135}
\begin{split}
\ce{A + B <=> C} \\
\ce{C + D -> E}
\end{split}
\end{equation}
The amount of C (in moles) that create/consumed in a unit volume per
unit of time is called the
{\bf chemical flux}\footnote{\url{http://en.wikipedia.org/wiki/Flux}}
of species C, denoted by $J$.
The equation is given
\begin{equation}
\label{eq:85}
J = [C]k
\end{equation}
with $[C]$ is the concentration of the product C
[mol.m$^{-3}$.s$^{-1}$], $k$ is the reaction rate.
{\bf NOTE}: Flux (from Latin: fluxxus - flow)
{\it indicates the amount of flow go through a unit of surface (area)
per unit time}.
There are different context for using flux, here, we talks about
{\bf chemical flux}, denoted by $J_{chem}$ [mol.m$^{-3}$.s$^{-1}$].
If we can call the amount of C produced in a unit time is $J_{in}$ and
the amount of C consumed in a unit time [s] is $J_{out}$, then the
change in concentration (in the direction of adding) of C in a unit
time [s] is
\begin{equation}
\frac{d[C]}{dt}=J_{in} - J_{out}
\end{equation}
Let's consider an isobaric, isothermic reaction, i.e. $p=const,
T=const$. Then, based on eq. \eqref{Gibbs} introduced in the previous
chapter, the change in Gibb's free energy of a particular substance is
\begin{equation}
\label{eq:64}
dG= -SdT + \mu dN + Vdp = \mu dN
\end{equation}
with $\mu$ is the
{\bf chemical
potential}\footnote{\url{http://en.wikipedia.org/wiki/Chemical_potential}},
i.e.the potential per mole.
Chemical potential of a component in a solution can be thought of in
many
ways\footnote{\url{http://www.geol.umd.edu/pages/facilities/lmdr/chmpot.htm}}.
In our case, it is defined as the rate at which the extensive internal
energy increase as the number of moles of the component in question
increase.
\begin{equation}
\label{eq:136}
\mu = \frac{\partial U}{\partial N_i}|_{S,V,N_{j\neq i}}
\end{equation}
Its unit is {\it the ``energy'' per mole} [J/mol], $dN$ is the change in
the number of moles for such substance (negative for loss,
positive for gain). Under such condition, the change in the Gibb's
free energy is caused by the change in the number of moles.
% or its variant {\it ``energy'' per mole}
{\bf NOTE}: In general thermodynamics, studied particles don't have to
be only atoms or molecules (i.e. the objects of chemistry). They can
be electrons, holes, or anything else that can be identified and
numbered. The ``chemical potential'' of the electrons, however, is
still a major parameter of the system (to the annoyance of the solid
state physicists - they therefore usually call it ``Fermi
energy'')\footnote{\url{http://www.tf.uni-kiel.de/matwis/amat/def_en/kap_2/advanced/t2_4_1.html}}.
\section{Flux within a single compartment}
\label{sec:flux-within-compartments}
Flux within a single compartment refers to the rate of formation or losing of a
species, i.e. the amount of change over a unit of time, e.g. $J_E = d[E]/dt$.
\section{Flux across compartmental border}
\label{sec:flux-across-compartments}
When spatial dimension is considered, then we need to also estimate the
the flux, i.e. concentration change, due to diffusion from one compartment to
another.
\section{Duffision}
\label{sec:diffusion}
\subsection{Diffusion formula}
Depending on the diffusion, the
formula can be simplified in two ways
\begin{enumerate}
\item isotropic:
\begin{equation}
\label{eq:1430}
\nabla.(D_C\nabla[\Ca]) = D_C(\partial^2[Ca]/\partial r^2 +
(2/r)\partial[\Ca]/\partial r)
\end{equation}
\item anisotropic:
\begin{equation}
\label{eq:1431}
\nabla.(D_C\nabla[\Ca]) = D_{Cx}\partial^2[\ca]/\partial x^2 +
D_{Cy}\partial^2[\ca]/\partial y^2 + D_{Cz}\partial^2[\ca]/\partial z^2
\end{equation}
\end{enumerate}
\subsection{How fast of a diffusion: effective diffusion constant}
\label{sec:effective-diffusion-constant}
Under a standard environment ({\it in vitro}), the diffusion of a species is
known as true diffusion. However, under an {\it in vivo} environment, the diffusion can be
slower, due to the viscosity in the media, etc. This diffusion is known as {\bf
effective diffusion constant}. To model kinetics of species inside the cell, we
need to consider using effective diffusion constants, and similarly with
effective volume (to be discussed in the next section).
\section{Currents (movements of ions)}
\label{sec:currents}
Current is the total amount of positive charges (in unit of Coulombs) flowing
across a fixed point in the wire per unit of time
\begin{eqnarray}
\label{eq:409}
I = \frac{dQ}{dt}
\end{eqnarray}
with the unit of current is {\bf Ampere} = A = Coulomb/sec = C/sec
(Sect.\ref{sec:current-ampere}).
The current is often calculated using the formula
\begin{equation}
I_s = \bar{g}_S \times \left( \text{driving forces} \right)
\end{equation}
with the driving force can be
\begin{enumerate}
\item linear - Sect.\ref{sec:ohms-law}
\item non-linear - Sect.\ref{sec:GHK_current}
\end{enumerate}
\subsection{current in physics}
\label{sec:current-in-physics}
By convention, in physics, current is the direction of the movement of positive
charge (or the opposite sign of the electron movement).
\subsection{current in electrophysiology}
\label{sec:current-in-electropysiology}
\label{sec:ionic-currents}
By convention, in physics, current is the direction of the movement
of positive charge (or the opposite sign of the electron movement).
However, in electrophysiology, the sign of the (density) current is positive if
the positive charged ion moves from inside to outside of the cell, and vice
versa. This is opposite for negative charged ions, e.g. $\Cl$. For the stimulus
current, if we use
\begin{equation}
\label{eq:1365}
\Csc \frac{dV_m}{dt}=-I_\ion + I_\stim
\end{equation}
then $I_\stim$ (from outside to inside) is positive. Otherwise, if we
use
\begin{equation}
\label{eq:1366}
\Csc \frac{dV_m}{dt}=-(I_\ion + I_\stim)
\end{equation}
then $I_\stim$ (from outside to inside) is negative. Again, the unit
$\Csc$ is $\mu$F/cm$^2$, and $V_m$ is [mV], while $I_\ion, I_\stim$ is
in $\mu$A/cm$^2$.
\begin{figure}[hbt]
\centerline{\includegraphics[height=4cm,
angle=0]{./images/electric_field.eps}}
\caption{Electric field}
\label{fig:elec_field}
\end{figure}
At a point in space, using Kirchoff's law (i.e. the law of conservation), the
applied current $I_\app$ is equal to the sum of the membrane current via
capacitor $I_m$ and the ionic current $I_\ion$ (positive if outward),
Fig.~\ref{fig:electric_current}.
\begin{equation}
\label{eq:689}
I_\app = I_\ion + I_m + I_{\text{axial}}
\end{equation}
with unit [ampere/unit area]. NOTE: $I_m = \frac{dV_m}{dt}$
%So, the transmembrane current $I_m$ is given by
%the sum of the ionic currents and the current flowing into the membrane
% capacity
If we convert to density current, i.e. current per unit surface area, then
\begin{equation}
\Csc \frac{dV_m}{dt} = - \sum j_{ion} + \frac{1}{A} I_\app - j_{\text{axial}}
\end{equation}
with the equation of individual inonic currents $I_{ion}$ can be selected from
Sect.\ref{sec:equat-ionic-curr}.
\begin{enumerate}
\item current density: nA/$\mum^2$
\item specific membrane capacitance: $\mu F/\mum^2$
\item surface area (A) : $\mum^2$
\item current injection ($I_\app$): nA
\end{enumerate}
If an AP is initiated at all points along the fibre simultaneously, the membrane
potential at all points along the fibre would be the same (or uniform). Then the
axial current $I_{\text{axial}}$ will therefore be zero.
% This type of response is called a 'membrane' action potential by
% Hodgkin-Huxley, and is rewritten with $I_m=0$ \citep{noble1962mhh}.
Finally, we obtain the widely use formula to model the electrical
property of a cell membrane
\begin{equation}
\label{eq:33}
\Csc \frac{dV_m}{dt} = -\sum_i g_i (V_m-E_i) + \frac{1}{A} I_\app %+
% I_{\text{axial}}
\end{equation}
with $\Csc$ in ($\mu$F/cm$^2$), $V_m$ in [mV], $g_i$ in (mS/cm$^2$),
and $I_\app$ in ($\mu$A) and $A$ is surface area.
\subsection{Current: Ampere}
\label{sec:ampere-current}
\label{sec:current-ampere}
The movement of 1 Coulomb per second is known as the unit of
electrical current, denoted as Ampere (A). It means that 1 (A) = 1
(C/sec), or
\begin{equation}
\label{eq:1345}
1 \text{ Faraday } = 9.649\times 10^4 (\text{A.sec})
\end{equation}
\begin{table}[hbt]
\begin{center}
\caption{Ionic currents}
\begin{tabular}{p{2cm}r}
\hline
notation & description \\
\hline\hline
$I_i$ & ionic current through plasma membrane ($I_{\ce{Na}},
I_{\ce{K}}$) \\
& ...$I_i = g_i(V_m-E_i) $ \\
$I_{app}$ & applied step current (rectangular pulse) (mA/cm$^2$)\\
$I_m=\frac{I_0}{A}$ & the current density per unit area (mA.cm$^{-2}$) \\
$i_m$ & the current density per unit length (mA/cm) \\
$I_\dhpr, I_\LCC$, $I_\Ca$ & the calcium current via DHPR (L-type channels) \\
$I_\ryr$ & the current of calcium released via RyR from JSR \\
$I_\Na$ & the current of sodium \\
$I_\NaCa$ & the current via Na/Ca exchanger \\
$I_\K$ & the current via potassium channel \\
\end{tabular}
\end{center}
\label{tab:Current}
\end{table}