physical principles and formalisms of electrical excitability

Upload: alkemir

Post on 25-Feb-2018

223 views

Category:

Documents


0 download

TRANSCRIPT

  • 7/25/2019 Physical Principles and Formalisms of Electrical Excitability.

    1/53

    C H A P T E R 6

    A A * I I

    A

    A

    A

    Physical principles and formalisms

    Departments of Physiology and Neurology, Albert Einstein

    College of Medicine, New Yo rk , Ne w YoFk

    Departmen t of Biophysics, T he Rockefeller Unive rsity,

    New York , New York

    of electrical excitability

    C H A P T E R C O N T E N T S

    Membrane Potentials and Ionic Fluxes

    Fundamental concepts

    Equilibrium state

    Nonequilibrium state

    Summary

    The Donnan equilibrium

    Phase-boundary potentials

    Two important, examples of equilibr ium situ ations

    Electrodes

    -

    he measurement of potential differences

    Quasi-equilibrium systems

    A membrane with a very large fixed-charge density

    An oil membrane

    Homogeneous uncharged membrane

    Homogeneous membranes with special properties

    Mosaic membranes

    Formal Consequences of Voltage-dependent Conductances

    Ion transport (the Nernst-Planck flux equations)

    The nature of electrical excitability

    Reasons for believing th at electrical excitability does not

    Hodgkin-Huxley equivalent circuit

    Current-voltage (I-V) characteristics

    Negative-slope conductance

    Changing the I - V characteristic without change of the g-V

    result from the shifting of ionic profiles

    characteristic

    Voltage-dependent Conductance in Thin Lipid Membranes

    The unmodified t hin lipid membrane

    Formation

    Permeability and electrical properties

    Carriers

    Channel formers

    A mosaic membrane formed with two modifiers

    Summary

    Monazomycin

    Alamethicin

    Excitability-inducing material

    Nonvoltage-dependent modifiers

    Voltage-dependent modifiers

    system, the modified thin lipid (or bilayer) mem-

    brane, which illustrates most of the relevant phe-

    nomena associated with nerve excitation. Before dis-

    cussing this model, however, we develop more or less

    from first principles the concepts of membrane poten-

    tials and ionic fluxes. This forms a ra ther large pa rt

    of the article and may be superfluous for the more

    sophisticated reader. Nevertheless we have included

    this material because, despite its importance for un-

    derstanding nerve excitation, it is rather inaccessible

    to students and investigators attempting to make

    initial contact with the neurophysiological literature.

    In the second section we discuss some of the formal

    aspects of the behavior of systems containing ele-

    ments whose conductances are profoundly affected by

    the voltage across them. With thi s as background, we

    then discuss the fascinating voltage-dependent phe-

    nomena that can arise in suitably doped thin lipid

    membranes.

    MEMBRANE

    POTENTIALS

    A N D IONIC FLUXES

    Fundamental Concepts

    Here we consider two general situations: equilib-

    rium and nonequilibrium states. We analyze the

    equilibrium sta te from both th e thermodynamic and

    the statistical mechanical viewpoint; the nonequili-

    brium state is handled by the Nernst-Planck flux

    equations.

    EQUILIBRIUM STATE.

    hermodynamic approach.

    The

    fundamental relation that we need from thermody-

    Single channels

    Summary and conclusion

    _ _ _ _ ~

    namics is that at thermal equilibrium the electro-

    chemical potential, pi, f any species i is the same in

    all phases to which the species has access. Thus, for

    phases 1 and 2 we can write

    pi (1) =

    pi

    (2) (1)

    provided that species i can move between the two

    phases. Equation

    1

    is the starting point of all our

    IN

    THIS CHAPTER

    we intend to elucidate the presently

    understood physicochemical principles and formal-

    isms underlying the electrical excitability of biologi-

    cal membranes. The primary analysis is of a model

    161

  • 7/25/2019 Physical Principles and Formalisms of Electrical Excitability.

    2/53

    162

    HA N D BOOK O F PHYSIOLOGY HE N ER V OU S SY STEM I

    thermodynamic treatments. For a n ideal solution, p,

    is given by

    p , = p , + RT In

    X ,

    + PV, + z , F

    $

    + any other

    relevant work terms (2)

    where p , is the standard chemical potential (in the

    particular phase being considered) of the ith species,

    X ,

    is mole fraction of ith species,

    V ,

    s partial molar

    volume of ith species, z, is valence of ith species (0,

    k l , +2,

    .

    . .), P

    is

    hydrostatic pressure of the phase, 4

    is

    electrostatic potential of the phase,

    R is

    the gas

    constant, T is temperature in degrees Kelvin, and F

    is the Faraday. (If the solution is nonideal, an activ-

    ity coefficient, y , , is included with the mole fraction

    term. In this article we deal only with ideal solu-

    tions.) If species i is dilute, then Equation 2 can be

    rewritten as

    p, = p ,

    + RT

    I n c, +

    PV, + z,

    F 4 +

    . . . (2a)

    where

    c,

    is the concentration of i.

    ( p l ( 0 )

    n Eq. 2a is

    different from p , n Eq. 2 because the standard state

    must be newly defined in going from mole fraction

    units to concentration units.)

    A s

    indicated in Equation 2 , other work terms must

    be included if they contribute

    to

    the potential energy

    of the species i. For example, if the system is in a

    gravitational field (or ultracentrifuge), there will be a

    term for the gravitational potential energy. For the

    systems we consider, these terms do not arise. In fact,

    even the

    PV

    term will generally be trivial and there-

    fore not ente r into our treatments. Thus, for our

    purposes,

    p,

    can be written as

    p, =

    p, (

    + RT

    In c,

    +

    z , F

    $

    (2b)

    Stat i s t ical mechanical approach. The basic rela-

    tion tha t we need from statistical mechanics is tha t at

    thermal equilibrium the particles satisfy the Boltz-

    mann distribution a t all points in the system to which

    they have access. That is, a t any point x (considering

    only a one-dimensional situation)

    c , x ) =

    C,,e-%(x)/h7

    (3)

    where,

    w ,

    XI is the potential energy per particle of

    the ith species a t

    x ,

    c,, is its concentration a t th e point

    defined as zero potential energy, and

    k

    is th e Boltz-

    mann constant. Multiplying the numerator and de-

    nominator of the exponent by

    N A ,

    Avogadros num-

    ber, we can write Equation 3 in the form

    c , x ) = c , , e - , s ) lRT (3a)

    where

    W,

    is the potential energy per mole of the ith

    species. If W,

    is

    purely electrostatic energy, then

    Equation 3a becomes

    c , ( x ) =

    ~ ~ e - 2 ~W X ) l R T

    (3a)

    For example, consider particles in a gravitational

    field. Then Equation 3 becomes

    =

    c e -l l l v s : k ? (4)

    where m is the mass of the particle and g is the

    acceleration due to gravity. If there were no thermal

    energy, all the particles would sit a t x = 0, he point

    of minimum potential energy. On the other hand if

    there were no gravitational field and only thermal

    energy, the particles would be homogeneously dis-

    tributed throughout space. Equation 4 is the compro-

    mise when both terms operate.

    It is important to realize th at Equation 3a is equiv-

    alen t to Equation

    1.

    (Take the logarithm of both sides

    of Equation 3a and identify W, as all the terms on the

    right in Equation

    2a

    except

    RT

    In

    ci.)

    Thus the ther-

    modynamic statement that the electrochemical po-

    tential of a species is the same in all phases is equiva-

    lent to the statistical mechanical statemen t that the

    molecules of the species satisfy the Boltzmann distri-

    bution.

    NONEQUILIBRIUM STATE.

    The bases of our discussion of

    ion transport a re the Nernst-Planck flux equations

    where $,+and

    $k-

    represent the flux (in mol/s)

    of

    the

    jth and kth ion, respectively, across unit area at any

    point x in the system,

    u ,

    s th e molar mobility of the

    jth cation,

    uk

    is the molar mobility of the kth

    anion,

    c,+

    and

    ck-

    represent the concentration of the

    jth and kth ions, respectively, a t any point I, nd

    is the electrical potential

    at

    any point x . (Again we

    are considering a one-dimensional situation with gra-

    dients of concentration and electrical potential occur-

    ring only in the x direction.)

    Whereas our basic equilibrium equations (Eqs. 1

    and 3) have

    a

    solid foundation in thermodynamics

    and statistical mechanics, the flux equations are less

    firmly grounded in theory, and in some instances

    they are even grossly incorrect. We therefore wish to

    make a few points concerning them tha t will give the

    reader some feel for their meaning.

    Equations 5 can be written i n the form

    where

    D = uRT

    (7)

    The Boltzmann distribution iS

    a

    quantitative state-

    merit of the balanced but incessant competition be-

    tween potential energy

    ( w )

    nd thermal energy ( k T ) .

    I some

    heoretical problems associated with justifying the use

    of the Boltzmann distribution in electrolyte theory do not concern

    us

    67).

  • 7/25/2019 Physical Principles and Formalisms of Electrical Excitability.

    3/53

    CHAPTER 6: PHYSICAL PRINCIPLES

    A N D

    FORMALISMS

    OF ELECTRICAL EXCITABILITY 163

    For

    a

    nonelectrolyte (z

    =

    O), Equation 6 simply be-

    comes

    dc

    dx

    4 = D-

    which is Ficks

    first

    law of diffusion. [Equation

    7

    is

    the famous Nernst-Einstein relation between the dif-

    fusion constant an d th e frictional coefficient (13,

    14);

    the mobility (u)

    s

    simply the reciprocal of the fric-

    tional coefficient.] Thus th e flux of a species is propor-

    tional

    to

    the negative gradient of its concentration.

    On the other hand for an ion

    at

    uniform concentra-

    tion throughout the system dc/dx = 01, Equation

    6

    becomes

    which is simply the equation for electrophoresis.

    That is, the ionic flux

    is

    proportional to the electric

    field (-d+/dx). Thus Equation 6 says that if there is

    both a concentration gradient and an electric field,

    the ionic flux is a l inear sum of the fluxes that would

    arise from each effect alone.

    Another way of looking at Equations

    5

    is to say

    that

    flux = concentration x velocity

    and

    velocity = mobility x driving force,2

    so that combining we have

    flux = mobility

    x

    concentration

    x

    driving

    Z F9)

    T dc

    4 = u x c x

    dx

    forc

    Substituting Equation 2b, the thermodynamic

    expression for the electrochemical potential, into

    Equation 8 we obtain Equation 5, the Nernst-Planck

    flux equations (if p,( is constant). Note that Equa-

    tion

    8

    reduces to the thermodynamic equilibrium

    condition, Equation

    1,

    when

    4i = 0

    at all points; or

    Equation

    5

    reduces to the BoItzmann distribution,

    Equation 3ar, when Cbi

    =

    0.

    [The relationship of the

    Boltzmann distribution (equilibrium state ) to the

    flux equations (nonequilibrium state) is clearly seen

    upon differentiating Equation 3a

    d%

    T

    dci

    CI dx du

    (which is the same relation a s obtained by se tting di

    = 0 in Equation 5). Thus a t equilibrium the sum of

    the electrical force

    ( -

    z,F d$ldx) and the diffusional

    force

    ( -

    RT/ci dc,/dx) cting on an ion is zero at

    every point in the system, and hence there is no flux

    of matter. When these two forces do not balance,

    there is a net driving force on the ion and hence a

    flux.

    For our purposes the Nernst-Planck flux equations

    (Eqs. 5) are perfectly adequate. Nevertheless the

    reader should be aware th at there exist situations for

    ziF 0

    The flux equation states that the force on an ion is the

    sum of two terms: -RTlc

    dcidx

    and -zF d$ldx. The

    second term is th e electrical force familiar from ele-

    mentary electrostatic theory, but the first term is

    more subtle.

    It is

    a phenomenological force resulting

    from the random (Brownian) motion of each individ-

    ual ion. Though the random movements of each ion

    are equally likely to be in the positive or negative

    direction, statistically it appears as if there is a force

    operating in the direction of the concentration gra-

    dient (4 1).

    Equation

    5

    is

    a

    special case of the more general

    expression

    which st ates tha t the driving force on the ith species

    is the negative gradient of its chemical potential.

    In a condensed phase such as water, an ion subjected to a

    force very quickly accelerates to its terminal velocity. In practice,

    therefore (unless we are dealing with very high frequencies),

    force gives rise to velocity rather than to acceleration.

    which these equations are insufficient. In particular,

    Equations

    5

    (or more explicitly Eq.

    8)

    say that the

    only driving force on a species i is the negative gra-

    dient of chemical potential for that species alone; it

    neglects the coupling of the gradients of chemical

    potentials for other species j to th e flux of species i. (A

    familiar example of such coupling is solvent drag,

    where the gradient of the chemical potential of water

    gives rise not only to a flow of water, but also to a flow

    of solute dissolved in the water.) Though these cou-

    plings, expressed as cross coefficients to forces acting

    :e

    (5)

    on other species, are extensively used by those who

    deal with the formalism of irreversible thermody-

    namics (36), they do not concern us here. For the

    systems we consider (which are

    of

    neurophysiological

    interest), they introduce corrections tha t are a t best

    second order.

    SUMMARY. We consider dilute, ideal solutions of ions

    in which gradients of concentration, potential, or

    solvent composition occur in only the x di re~ t ion .~

    The sta rting point for the t reatment of these systems

    is one of the following equations

    4.

    -

    dc, d%

    I uiRT ziuiFci 5) Transport

    dx dx

    For the general case, d/dz is replaced by

    C

    n all equations.

  • 7/25/2019 Physical Principles and Formalisms of Electrical Excitability.

    4/53

    164 HA N D BOOK OF PHYSIOLOGY THE NERVOUS SYSTEM I

    m e m b r a n e

    C I -

    F I G .

    1.

    The Donnan system.

    Two Impor tan t Examples of Equi l ibr ium S i tua tions

    THE DONNAN

    E Q U I L I B R I U M .

    Aside from its intrinsic

    physiological intere st, the Donnan equilibrium illus-

    trates almost all the important concepts and difficul-

    ties associated with membrane potentials; we there-

    fore analyze it in some detail. The Donnan equilib-

    rium arises when a membrane separates two solu-

    tions containing both permeant and impermeant

    ions. For simplicity we restrict our trea tment to the

    case of an infinitesimally thin membrane separating

    two infinite solutions; both solutions contain a single

    permeant univalent positive (Na+) nd negative (Cl-)

    ion species, and, in addition, one solution contains a

    univalent impermeant ion, zN, where

    z = + 1

    (Fig.

    1). (The mechanism of membrane semipermeability

    is irrelevant to our treatm ent. In practice, the Don-

    nan situation commonly arises when a porous mem-

    brane, such as dialysis tubing, holds back a macro-

    ion (e.g., protein) which cannot

    fit

    through the pores.

    Usually the solvent, water, is also permeant.)

    Thermodynamic ana lys i s . The conditions that

    must be satisfied are, from Equation

    1

    /- .I

    ( l ) ~ N . I 2 ) (9a)

    k l

    (1 )

    = P(I

    ( 2 ) (9b)

    which become, upon substituting from Equation 2b

    &,+ +

    RT In a+], -t F ,

    (10a)

    = &;;+ + RT In [Na+In+

    F I / J ~

    Note that there is no such equation for N, since we

    have specified that it cannot move between the two

    aqueous phases. (We have excluded a third equation

    equat ing the chemical potential

    of

    water on the two

    sides of the membrane. Although the osmotic effect

    accompanying the Dorinan equilibrium is of consider-

    able general physiological interest, it is not signifi-

    can t for electrophysiology.) Adding Equation 10a and

    10b we obtain

    or

    This is known a s the Gibbs-Do nnan condi t ion, and

    r

    is called the Don nan ratio .

    From Equations 10a or lob we can directly obtain

    the potential difference across the membrane

    RT a+],

    v =

    ($, - J J ~ )=

    ~

    In

    ~

    F

    a+],

    RT

    [Cl-ll RT

    F

    [Cl-]? F

    -

    - -

    In ~=

    -

    ~ In r (13)

    To

    calculate r, and hence the membrane potential,

    we invoke the familiar electroneutrality condition

    on both sides of the membrane

    [Na+I1= [Cl-1, =

    c I

    a+],

    +

    z[N]

    =

    LCl-1, (14b)

    (I t is to be understood tha t the electroneutrality con-

    dition holds only for remote regions on eith er side

    of

    the membrane, that is, for macroscopic regions. As

    will be seen presently, the al terna tive treatment by

    means of the Poisson-Boltzmann equation establishes

    the concentrations and potential as continuous func-

    tions ofx, and indeed regions very close to the mem-

    brane are not electrically neutral

    .)

    Combining Equa-

    tions 14 and 12 we obtain

    (14a)

    (electroneutrality condition)

    Figure 2 is a sketch of r as a function of the imper-

    meant ion (N) concentration for

    z =

    +1. As [N+]

    increases,

    r

    decreases from

    1;

    that is, permeant posi-

    tive ion concentration decreases on side 2 and per-

    meant anion concentration increases. In the limit of

    large +I, there is virtually no Na+ on side 2 and

    [Cl-1,

    =

    +I. The converse occurs for large -1.

    Note also from Equation 13 th at t he membrane po-

    tential, \Ir, will be positive or negative depending on

    whether

    z

    is

    +

    1 or -1 and th at the absolute value of

    \I

    increases as [N] increases. Figure 3 is a diagram of

    r

    -

    ~

    = r

    Na+;l,

    -

    [Cl-1,

    a+:[, [Cl-1,

    FI G . 2 .

    Qualitative plots of the Donnan ratio, r, as a function

    of the impermeant ion concentration,

    IN],

    for an impermeant ion

    of

    valence

    +1

    or - 1 . (See Eq.

    1 5 . )

  • 7/25/2019 Physical Principles and Formalisms of Electrical Excitability.

    5/53

  • 7/25/2019 Physical Principles and Formalisms of Electrical Excitability.

    6/53

    166

    HANDBOOK OF PHYSIOLOGY THE NERVOUS SYSTEM I

    and

    [Cl-]+,

    =

    [Cl~]~,e+~YH

    FIG.

    5.

    The two-phase system.

    generated from a given initial condition. Thus, al-

    though the impermeability of N is the ultimate cause

    for the membrane potential and ion asymmetries,4

    the $ and concentration functions are inexorably

    coupled. That is, the

    $

    profile affects the concentra-

    tion profiles which in turn affect the

    J,

    profile, and

    which is the Donnan potential as given in Equation

    13. Furthermore by equa ting the two expressions for

    q

    we see that

    lCI-I+,lNal,

    =

    [Cl-lL,[Na+l-, so on.

    We have dealt in some detail with the Donnan

    equilibrium because the general nature of the

    results

    which we shal l discuss shortly. Basically the thermo-

    which is the Gibbs-Donnan condition a s in Equation

    we have the more general statement

    2. (Note tha t bY

    Equations 16a and 16b

    applies to many other equilibrium systems, one of

    lC1-l[Natl = [C-J-,[Na+]_,

    =

    constant

    dynamic approach gives the-concentrations and po-

    at every point x . )

    Second, we now finally see the specific charge sepa-

    ration that gives rise to the electrostatic potential.

    There exist narrow space-charge regions on the two

    sides of th e membrane; in solution 1 he space charge

    is negative and in solution 2 it is positive, with of

    course

    .r:& = -

    J1:pbr

    (24)

    The potential, $, and the ion concentrations, a+]

    and [Cl-I, vary continuously in these regions to their

    final values in the remote regions.

    A

    comparison of

    Figure 3 with Figure 4 clearly illustrates the differ-

    ences between the thermodynamic and th e statistical

    mechanical point of view. In the former case, the

    $

    and concentration functions are constant a t thei r re-

    mote values on the two sides of the membrane, with

    a discontinuity in these values occurring at the plane

    of the membrane; in the latter case these functions

    are seen to vary continuously from one remote region

    to the other.

    The ex ten t of the space-charge regions is deter-

    mined by the Debye length, L,) Eq. 23). Roughly

    speaking, p falls about e-fold for every Debye length.

    For 0.1

    M

    salt solution in water

    ( E =

    801,

    L,, =

    10

    A;

    thus the region where electroneutrality is signifi-

    cantly violated extends about 40-50

    A,

    too small for

    direct sampling with a pipette. Note from Equation

    23 th at L,, is directly proportional to 4 2 nd inversely

    proportional to f l a a Thus in low dielectric con-

    stant media, L,, will tend to be smaller, whereas in

    media of low ionic strength,

    L,,

    will be larger. We

    shall refer back to this point later when discussing

    ion distributions in a lipid, or hydrocarbon, phase.

    Finally it should be realized that the system is a

    typically nonlinear one, which makes it difficult to

    describe how the J, and concentration functions are

    tentials far from the membrane, or interface,

    whereas the Poisson-Boltzmann treatment explicitly

    describes how the membrane potential arises from

    space-charge regions near the membrane; electroneu-

    trality holds only a t remote (many Debye lengths)

    regions.

    PHASE-BOUNDARY POTENTIALS. In th e Donnan equilib-

    rium, the asymmetry of permeable-ion distributions

    (the Donnan condition) and the membrane poten-

    tia l ar ise because a macro-ion in one of the aqueous

    solutions

    is

    impermeant. Here we consider the case in

    which no impermeant ion is present, but the two

    phases are different (for simplicity we consider them

    immiscible); for example, one phase

    is

    water and the

    other is oil (Fig. 5).

    Thermodynam i c anal y s i s . From Equations

    1

    and 2b

    we have

    pi;;,+,,

    + RT In a+], + F$,

    = p l&t )2

    RT In [Nail2

    +

    F$2

    (25a)

    + RT In [Cl-1,

    -

    F $ ,

    = p;: )l-)s

    RT In [ClV], - F$2

    (25b)

    These are the same equations employed for the Don-

    nan equilibrium (Eqs. 10a and lob); this time, how-

    ever, the p s are not the same on sides l and 2,

    because the solvents are different. The conditions of

    electroneutrality for remote regions also give

    a+], = [Cl-I, = [NaCIl, (264

    [Na+12

    =

    [Cl-I,

    =

    [NaCll, (26b)

    Substituting Equation 26 into Equation 25 and add-

    ing we obtain

    For the t rue Donnan case, where N

    is

    not fixed

    at

    a constant

    value to th e right of the membrane, we would h ave

    an

    Equation

    16c, stating th at N satisfies the Boltzmann dis tributio n for x

    >

    0.

    The final result is not much different from th at shown in Fig. 4 ,

    except that

    N

    is perturbed upward near the membrane.

  • 7/25/2019 Physical Principles and Formalisms of Electrical Excitability.

    7/53

    CHAPTER

    6:

    PHYSICAL

    PRINCIPLES

    AND

    FORMALISMS

    OF

    ELECTRICAL

    EXCITABILITY 167

    Upon substituting Equation 26 into Equation 25 and

    subtracting we also obtain

    *O/\V =_ ( 2 - 1 )

    (28)

    (dk+,] P$L)J - (PL:%+)y PII.l-),)

    -

    2F

    Equation 27 is an expression for the oil-water (o/w)

    partition coefficient

    of

    NaCl in terms of the standard

    chemical potentials of Na+and C1- in the two phases.

    Equation 28 gives the phase-boundary potential in

    term s of these same quantities.

    It is instructive to transform these expressions into

    ones containing the intrinsic partit ion coefficient of

    each ion. At the boundary between the two phases

    there is a discontinuity in standard chemical poten-

    tial and hence a discontinuity in ion concentrations.

    Of course, in (classical) reality there are never actual

    discontinuities; all functions are continuous. Never-

    theless, because th e boundary between two phases is

    established through short-range van der Waals inter-

    actions, the phase transition occurs over molecular

    distances (-2 A) and hence can be practically treated

    as a discontinuity, compared to the space-charge

    re-

    gions that can extend over tens or even hundreds of

    angstroms. Thus the p*.))s,nd hence t he ion concen-

    trations, can change precipitously over a distance

    where the electrostatic potential,

    $,

    remains un-

    changed.s With this in mind, and assuming that

    there is not a layer

    of

    dipoles at th e interface, we can

    write for th e conditions at the boundary

    Equations 30a and 30b are expressions for the intrin-

    sic partition coefficients

    G.1

    of Na+ and C1-, respec-

    a Image forces, which make a major contribution to the intrin-

    sic partition coefficient, extend over distances longer than the

    phase-transition region, but these are still in general much

    shorter than the space-charge regions.

    tively. That is, they a re th e partition coefficients tha t

    would be observed in remote regions

    i f

    somehow elec-

    trostatic interaction among ions did not occur. Sub-

    stituting Equation 30 into Equation 27, we have

    p o / w = d(PNa+)(PCI-) (31)

    and substituting Equation

    30

    into Equation 28 gives

    ,

    (32)

    Equation 31 states that the macroscopically ob-

    served partition coefficient for NaCl is the geometric

    mean of the individual partition coefficients

    @ )

    for

    Na+ and C1-. Note th at if one of these approaches

    zero, the parti tion coefficient for the sal t approaches

    zero. Equation 32 states that the phase-boundary

    potential is determined by the rat io of the individual

    partition coefficients. Note that if these are equal,

    the phase-boundary potential is zero. Also note that

    in contrast to the Donnan potential, the phase-bound-

    ary potential is not a function of the NaCl concentra-

    tion (provided, of course, the partition coefficients

    are

    not concentration dependent).

    Stat is t ical mechanical analys is .

    Had we stopped

    our thermodynamic analysis with Equations 27 and

    28, we would have been in much the same position we

    were in with our thermodynamic treatment of the

    Donnan equilibrium; the concentrations of NaCl are

    different in the two phases, and there is a potential

    difference between th e two phases (Fig. 6). Again th e

    formal expressions are perfectly correct, and again

    the origin of the boundary potential

    is

    obscure. By

    extending our thermodynamic analysis to the parti-

    tioning occurring a t the boundary, we have given a

    fairly strong hint as to the source

    of

    the boundary

    potential. Clearly, applying the Poisson-Boltzmann

    analysis with these boundary conditions will show

    that space-charge regions exist in both the aqueous

    and oil phases. Figure 6 will then be transformed into

    the more complete Figure 7. We shall not go through

    the formal treatment since, aside from the mathe-

    matical details, there are no new physical principles

    that have not already been considered for the Donnan

    case. The extent of the space-charge regions in the

    two phases will depend on their respective Debye

    I

    ql=O

    I

    FIG. 6 . Concentrations and potentials in the water and oil

    phases. The potential, 2, in the oil phase i s positive, because we

    have assumed that the partition coefficient for sodium, pNa+,

    between oil and water

    is

    greater than that for chloride,

    p c , - .

  • 7/25/2019 Physical Principles and Formalisms of Electrical Excitability.

    8/53

    168

    HANDB OOK OF PHYSIOLOGY -

    THE

    NERVOUS

    SYSTEM

    I

    lengths. Since

    L D =

    Jmhere tends to be a

    compensating effect between dielectric constant and

    salt concentration. Thus a low dielectric constant by

    itself would lead to a smaller Debye length, but in

    general a low dielectric constant

    is

    accompanied by a

    small partition coefficient of the salt between water

    and the low dielectric constant phase, which leads

    to

    a larger Debye length. In the cases we consider later

    of membranes with an essentially hydrocarbon inte-

    rior, the concentration term strongly predominates,

    so that the space-charge region in the membrane

    interior is much more extended than the one in the

    aqueous phase. Of course, regardless of the extent of

    the space-charge regions, overall charge conservation

    must always obtain

    pdx = - 1: p d x

    (24)

    We might also note that a truly continuous treat-

    ment would not show the discontinuities in ion con-

    centrations at = 0, as is depicted in Figure

    7,

    but

    rather would show

    a

    continuous transition over a

    distance of a few angstroms. Such a t reatment would

    require a "van der Waals, image force,

    . . .

    -Boltz-

    mann" analysis of this region, in analogy to the Pois-

    son-Boltzmann treatment. Needless to say, the much

    more complex nature

    of

    these forces makes such an

    analysis extraordinarily difficult (if not impossible).

    il ,

    Electrodes - he Measurement

    of Potential Difference

    Up to this point in our discussion of potentials

    associated with ionic systems, we have not described

    how one goes about measuring these potentials. Be-

    fore considering other examples

    of

    membrane poten-

    tials, we must discuss this problem. We are con-

    I

    *O/

    w

    N a t , q

    *--- - -

    N o t l $/'

    I /

    A

    r

    I

    I S o l u t i o n

    1

    I S o l u t i o n 2 I

    FIG.

    8. Method for measuring the potential difference across a

    membrane.

    cerned not with the technical questions of which am-

    plifiers to use

    o r

    what brand of oscilloscope is best,

    but rather with an important theoretical question

    (which also happens to have important practical im-

    plications). The basic problem is the following: in

    order to measure the potential difference across a

    membrane, we must insert a pair of electrodes into

    the system- ne electrode on each side of the mem-

    brane (Fig. 8). By necessity there will exist

    at

    each

    solution-electrode (soln/elec) interface a potential,

    generally called an

    electrode pot entia l .

    Thus the po-

    tential th at we measure,

    Y,,, , ,~i5u,. , ,c, ,

    s in principle the

    algebraic sum of three potentials: the membrane po-

    tential (the quantity

    of

    interest) plus two electrode

    potentials

    -

    q'measured

    -

    Vrrnernbrane

    4-

    qeleclsoln I +

    q s o i n

    2leier

    (33)

    How then do we make contact with the solutions so

    that the sum of the last two terms in Equation 33 is

    negligible? This

    is

    crucial, for when we go to measure

    a membrane potential, we want to measure a quan-

    tity tha t is a unique property of the ionic system and

    not a quantity that is dependent on the particular

    pair

    of

    electrodes we happen to choose.

    To illustrate more concretely the problem

    of

    meas-

    urement, consider again the Donnan system of Fig-

    ure 1. With two theoretical formulations we have

    shown that

    RT [Cl-1,

    In~

    [Cl-1,

    - --

    * ,,.ml,r;,nC.

    -

    ( 1 3 )

    and the question now is, can we measure it? Suppose

    we use reversible Ag-AgC1 electrodes. [The electrode

    reaction i s AgCl + e + Ag")'+ C1- (soln).] What then

    is

    ~ , , , ( . ; l S U l . ( . t , ?

    The answer turns out t o be zero. Let us

    see why. The "electrode potential" of a Ag-AgC1 elec-

    trode in contact with a solution containing C1- is

    where qo s the so-called standard potential. (In thi s

    case it is the potential

    of

    the half cell when the

    solution is one molar in chloride.9 If we keep our sign

    FIG.

    7. Sketch of concentration and potential profiles for a

    phase-boundary equi librium a s determined from the Poisson-

    Boltzmann analysis (cf. Fig. 6, which is the result

    of

    the thermo-

    dynamic analysis; as in

    Fig. 6, P , ~ . >

    p c i - ) . The space-charge

    density is shown in th e lower par t

    of

    the figure.

    It is interesting to note that Eq. 34 can be derived by exactly

    the Same methods we have employed in treating our pure ionic

    systems. Thus, since the system is in equilibrium and

    C 1 ~

    an

    move between the two phases (solid and solution), we have from

    our thermodynamic relation

  • 7/25/2019 Physical Principles and Formalisms of Electrical Excitability.

    9/53

    CHAPTER 6: PHYSICAL PRINCIPLES AND FORMALISMS O F ELECTRICAL EXCITABILITY

    169

    convention straight we then have

    RT

    + (*(,

    - p

    n [Cl-I,) = 0

    r

    * i * Iv r / w1 n

    2

    The electrode potentials exactly cancel the mem-

    brane potential, and operationally we measure no

    potential difference

    at

    all. We certainly made a poor

    choice of electrodes

    We could have predicted th is result purely from the

    second law of thermodynamics without going

    through the above algebra. Since each electrode

    is

    in

    equilibrium with its solution and the solutions are in

    equilibrium (Donnan) with each other, the entire

    system must be in equilibrium. If there indeed were a

    potential difference between the electrodes, we could

    construct a perpetual motion machine of the second

    kind. That is, we could connect a load between the

    two electrodes and do work. At one electrode, C1-

    would go into th e solution, and at the other electrode,

    C1- would come out; the overall chemical composition

    of the solutions would not change. If the Donnan

    condition became perturbed by the transport of NaCl

    from one solution to the other, we could merely pause

    for a while and let the Donnan condition reestablish

    itself (utilizing, of course, only thermal energy).

    When one electrode becomes almost depleted of AgC1,

    we merely switch the electrodes from one solution to

    the other, a process tha t, in principle, requires negli-

    gible work. Thus we could indefinitely convert ther -

    mal energy into work without any other change in

    the universe -a clear violation of the second law of

    thermodynamics.

    Since a reversible pair of electrodes will always

    give =

    0,

    we must try something different.

    Why not use a pair of stainless-steel wires? We might

    indeed measure the correct Donnan potential, but

    then aga in we might not. The problem is that there is

    not a well-defined process dominating the potential of

    k I (solid) = k 1 soh)

    &? solid) - FJlplrc

    &','

    (sold

    +

    R T In

    ICI-I

    - F&,

    or

    and rearranging, we obtain

    Eq.

    34 where

    and

    ~ ' , 4 , . , h < , l ( elPC - l W d

    1

    Po = j

    (d'i) solid) -

    i' (soh))

    Similarly a Poisson-Boltzmann analysis would show an extended

    space-charge region in the solution near the electrode and a very

    narrow one in the electrode itself.

    the steel wires. What is the dependence of thi s poten-

    tial on Na+ and C1- (or in th e more general case of the

    Donnan equilibrium, on any other permeant ions

    present in the system) concentration? What effect

    does the macro-ion N have on this potential? It is

    possible that none of these have significant effect on

    the electrode potential, and therefore the two elec-

    trode potentials will be equal and cancel each other

    out, leaving the Donnan potential a s the only quan-

    tity measured. But we cannot be sure , since we have

    no theory to work from.

    It turns out that we

    can

    measure the membrane

    potential by introducing appropriate sa lt bridges. In-

    stead of putt ing t he Ag-AgC1 electrodes directly into

    the solutions, we place them into 3

    M

    (or saturated)

    KC1 and make contact to the solutions through t he 3

    M

    KCl (Fig.

    9) .

    Now at

    first

    glance it might appear

    that things are even worse, because the measured

    potential is now the sum of five potentials instead of

    three:

    In addition to the two electrode potentials, there are

    now two liquid junction potentials -one between 3 M

    KCl and solution 1 and one between 3 M KCl and

    solution 2. On reflection, however, it is clear that

    things are not worse than before, for since the Ag-

    AgCl electrodes are in identical solutions (i.e., 3 M

    KCl), the electrode potentials must be equal and

    hence cancel. This then leaves the two liquid junction

    potentials with 3

    M

    KC1, and it t urn s out that these

    are small, both because K+ and C1- have the same

    mobility and because their concentration is large

    (40).

    Given a sal t bridge with a large concentration of

    a salt whose ions have equal mobilities, it can be

    shown tha t the liquid junction potential between this

    bridge and any "reasonable" solution is small

    (40).

    (KCl happens to be the most convenient salt to use.)

    Thus the only term left in Equation 35 hat is either

    not negligible or does not cancel out is 9m ml ri nhe

    quantity we wish to measure.

    Note t hat , although solutions 1 and 2 in our Don-

    nan example contain C1-, the presence of thi s partic-

    ular ion is not relevant to our reason for using 3 M

    KCl; these bridges would be equally effective with

    chloride-free solutions. Also note th at since the metal

    Ag/AgCI, ,Ag/AgCI

    rne m i

    one

    FIG. 9. Method of measuring membrane potential by making

    contact with the solutions through

    3

    M

    KCI

    junctions.

  • 7/25/2019 Physical Principles and Formalisms of Electrical Excitability.

    10/53

    170

    HANDBOOK

    O F

    PHYSIOLOGY HE NERVOUS SYSTEM I

    electrodes are in identical solutions, any identical

    pair of electrodes in theory could be used (even stain-

    less-steel wires). For practical reasons of stability and

    convenience, the most commonly used electrodes ar e

    either Ag-AgC1 or calomel (Hg-Hg,Cl,).

    It should also be clear that the measurement, by

    use of the arrangement shown in Figure

    9,

    of a finite

    potential difference for a Donnan equilibrium is not a

    violation of the second law, since with the salt

    bridges present, the complete system of membrane

    and

    electrodes is not in equilibrium.

    Finally, we must warn the reader who wishes to

    pursue th is subject

    of

    the measurement of membrane

    potentials fur ther tha t he will come across the view of

    certain purists who claim that, since one can only

    measure ~ l , l , , l , , , ~ , . , l l , , . through some such artifice as the

    introduction of liquid junctions, it is not sensible to

    even talk about membrane potentials. That is, it

    is

    impossible to assign values to the individual te rms in

    Equation

    35,

    and hence one can only talk about

    *,)

    ),,,,,

    In fact, there are some ingenious argu-

    ments that prove that most of the measured Don-

    nan potential occurs not a t the membrane but a t the

    contact between salt bridges and solution. It is not

    our purpose to contribute to these polemics. We hope

    that the Poisson-Boltzmann analysis of the Donnan

    equilibrium will convince anyone th at indeed there is

    a membrane potential intrinsic to that system and

    that our treatment of the phase-boundary potential

    (and other systems we shal l discuss shortly) will also

    convince the skeptic of the reality of membrane po-

    tentials. It might also be germane to point out that if

    electrophysiologists had taken the purists critique of

    membrane potentials seriously, the subjects of elec-

    trical excitability, receptor potentials, and postsyn-

    aptic potentials would never have gotten

    off

    the

    ground, and most neurophysiologists today would be

    out of business.

    Quasi-equilibrium S,ystems

    Before we take up the problem of diffusion poten-

    tials involving the flux equations, we consider two

    nonequilibrium systems t ha t a re sufficiently close to

    equilibrium that they can be treated with good accu-

    racy by the methods already employed. The consider-

    ations developed here a re particularly relevant to our

    future discussions of bilayer membranes.

    SITY. uppose tha t we modify the Donnan system in

    Figure

    1 so

    that the solution with the macro-ion

    separates

    two

    solutions of NaCl (Fig. 10). Further-

    more, let the macro-ion, N, be uniformly distributed

    and immobile (as in the ion-exchange resin we dis-

    cussed previously (see subsection

    Statistical mechan-

    ical analysis (Poisson-Boltzmann equation).

    Then

    the middle compartment (m) of Figure

    10

    is a mem-

    brane (in point

    of

    fact, an ion-exchange membrane)

    separating solutions 1and 2 . Let us also assume that

    A

    MEMBRANE

    WITH A

    V E R Y LARGE FIXED-CHARGE DEN-

    the concentration of fixed charge is large compared to

    the concentration of NaCl in the two compartments;

    that

    is,

    [Nl

    +

    [NaCI],, [NaCll,. If the membrane

    thickness is large compared to the Debye length in m,

    then with solutions of equal concentration on the two

    sides, there exist two Donnan potentials of the same

    magnitude,

    as

    shown in Figure

    1 l A

    (where for the

    sake of concreteness we take z = + l ) . igure 1 l A is

    simply a symmetrical duplication of Figure 3 . At each

    interface, there is a large Donnan potential jump

    between solution and membrane, but there is no

    potential difference across the membrane, because

    these two jumps are equal.

    Now consider the case in which th e concentrations

    of NaCl in solutions

    1

    and 2 are not equal (e.g.,

    [NaCll, < [NaCl],). This system is no longer in equi-

    librium, and NaCl will diffuse slowly from solution

    1

    to 2; within the membrane there is a concentration

    1 ) m ( 2 )

    FIG. 10. A n ion-exchange membrane separating two solu-

    tions.

    Concentra t ion

    I /

    /i

    P o t e n t i a l

    p r o f i l e

    63

    \ C o n c e n t r a t i o n

    I I p r o f i l e s

    w

    Pot en i a I

    o r o f i l e

    = O i

    1)

    ( m )

    2

    FIG. 11.

    A:

    concentration and potential profiles for an ion-

    exchange membrane of large positive fixed-charge density sepa-

    rati ng two solutions of equal concentration of NaCI. B : same as

    A , except tha t NaCl concentrations in compartments 1 and

    2

    a r e

    unequal. (T he Na and C1- profiles within the m embrane have a

    small negative slope tha t is not clearly seen in the figure.)

  • 7/25/2019 Physical Principles and Formalisms of Electrical Excitability.

    11/53

    CHAPTER

    6: PHYSICAL PRINCIPLES AND

    FORMALISMS

    OF ELECTRICAL

    EXCITABILITY

    171

    gradient of Na+ an d C1- -the concentration profiles

    within the membrane have a finite slope. However,

    the high concentration of fixed charge, N +, permits

    only a very small concentration of Na+ in th e mem-

    brane; thus the salt concentration gradient will be

    quite shallow, and diffusion of NaCl across the mem-

    brane will be very slow indeed. To a

    first

    approxima-

    tion we can therefore neglect th e finite slopes within

    the membrane and the slow flux of NaCl across the

    membrane, and tr eat t he system by our equilibrium

    methods. Thus, as in the case when the concentra-

    tions of NaCl in solutions

    1

    and

    2

    were equal, we

    again have two Donnan equilibria. This time, how-

    ever, they are not identical, and the re exists across

    the membrane a potential difference (Fig. 11B).

    For highly charged membranes, the counter ion is

    virtually the only mobile ion present within the

    membrane. Thus the membrane is "permselective"

    for C1- (if z = -1, it is permselective for Na+). Note

    th at this permselectivity arises purely from the Don-

    nan effect and is not dependent on an y steric factors.

    On the basis of th e permselectivity for C1-, we can

    immediately calculate the membrane potential from

    our thermodynamic equilibrium criterion

    k T ( 1 )

    = k T ( 2 )

    pi ;' +

    RT

    In [Cl-I, - FICII

    = & ;)-

    +

    R T In [Cl-I, - FI, I~

    If instead the membrane contained a negative fixed

    charge,

    it

    would

    be

    permselective for Na+,and by the

    same argument as above we would obtain

    R T "af],

    Y = + - l n -

    F "a+],

    In general, for a membrane that is permselective for

    an ion, i, of valence z (regardless

    of

    the mechanism

    for the permselectivity), we obtain, by equating the

    electrochemical potential of the ion on the two sides

    of the membrane

    (37)

    Equation

    37

    is often called the

    Nerns t

    equation, and

    v'

    is called the Nerns t potent ia l .

    It is instructive to see how the membrane potential

    can also be derived from the algebraic sum of the two

    Donnan potentials. Since N+

    s

    very large, within the

    membrane [Cl-1,

    --

    "+I. The Donnan potential be-

    tween solution

    1

    and th e membrane

    (m)

    is

    and between solution 2 and the membrane

    Combining we have

    which is Equation 36. Thus the transmembrane po-

    tential is made up of the difference between two large

    Donnan potentials.

    We have been assuming that the membrane thick-

    ness is large compared to the Debye length within it ,

    and we have therefore been able to draw the concen-

    tration and potential profiles as in Figure

    11,

    without

    worrying about the very th in space-charge regions. If

    the membrane thickness and the Debye length were

    comparable, the continuous profiles in the space-

    charge regions would have to be explicitly calculated.

    Also it would now be meaningless to speak of the

    transmembrane potential as the sum of two Donnan

    potentials

    at

    each interface, since the distinction be-

    tween interface and electroneutral membrane inte-

    rior no longer exists; the space-charge region extends

    throughout the enti re membrane. To explicitly calcu-

    late th e membrane potential would require t he solu-

    tion of the Poisson-Boltzmann equation. It turns out,

    however, that even in such a thin membrane, vir-

    tually the only mobile ion present is C1-, for which

    the thin membrane is still permselective. Thus our

    equilibrium results are still applicable, and the mem-

    brane potential will still be the Nernst potential for

    c1-.

    AN

    OIL MEMBRANE.

    Let us now, in the same way th at

    we extended the Donnan system of Figure

    1

    o make

    an ion-exchange membrane in Figure 10, extend the

    water-oil system of Figure 5 to make an oil mem-

    brane bounded by water phases

    as

    in Figure 12.

    Suppose, to start with, that the membrane

    is

    thick

    compared to

    the

    Debye length within it and that

    the

    NaCl concentrations on the two sides differ. If we

    assume tha t the partition coefficients of one or both of

    the ions is very small, then again we can neglect the

    small gradients of NaCl within the membrane and

    the slow flux of NaCl across the membrane and

    treat

    the system

    as

    being

    at

    equilibrium. The concentra-

    tion and potential profiles are shown in Figure 13,

    where for the sake of concreteness we have made the

    Na+ partition coefficient considerably larger tha n

    th at of C1-. We see that there are two large positive

    (H,O)

    ; (o i l ) ;

    H,O)

    1) m ) ( 2 )

    FIG. 12.

    An oil membrane separating two NaCl aqueous solu-

    tions,

  • 7/25/2019 Physical Principles and Formalisms of Electrical Excitability.

    12/53

    172 H ANDBO O K O F

    PHYSIOLOGY

    - THE NERVOUS SYSTEM I

    Not

    1 - 1 -

    I

    2

    I

    I

    Concentration

    :

    ( o i l )

    I

    Na+.CI- prof I

    les

    I +m I

    I I p r o f i l e

    -

    Potential

    I

    1

    I I

    I I

    I

    F IG . 13 .

    Concentration and potential profiles for a thick oil

    membrane separating two NaCl aqueous solutions. The potential

    within the membrane

    ($,)

    is positive, because we have assumed

    that N a+ partitions better into the membrane than C1-.

    phase-boundary potentials, but they are equal and

    given by Equation

    32

    As we pointed out earli er, the phase-boundary poten-

    tial is not

    a

    function of the NaCl concentration in the

    aqueous phase. Even though N a+ is much more fa-

    vored in the oil phase than C1- and consequently

    there are large phase-boundary potentials, the total

    membrane potential is zero. This result contrasts

    sharply with that for the high-density fixed-charge

    membrane we discussed earlier,

    across which the

    Nernst potential appears.

    The above analysis was predicated on the mem-

    brane being thick compared to the Debye length

    within it. Consider now a membrane of thickness

    comparable with t he Debye length. Instead of Figure

    13,

    we must now draw the complete ionic profiles

    including the space-charge regions, since there is no

    electroneutral region within the membrane. The pro-

    files in Figure 14 ar e an extension of those in Figure

    7 . For comparison we have redrawn Figure 13 in

    Figure

    14A,

    exaggerating the space-charge regions

    for

    a

    thick membrane; in Figure 14B we have ex-

    panded the scale, since the space-charge regions ex-

    tend through the enti re membrane.

    To calculate the membrane potential we cannot use

    Equation

    32,

    but must solve the Poisson-Boltzmann

    equation. By inspection of the profiles, however, we

    see

    that Na+

    is

    virtually the only ion in the mem-

    brane; that is, the membrane is permselective for

    Na+. Consequently the membrane potential mus t be

    given by the Nernst potential for Na+

    This is true provided the mobilities of Na+ and C1- in the

    membrane are equal. If they are not, then,

    as

    we shall see in a

    later section, there will be a diffusion potential due to the asym-

    metry in mobility. Nevertheless our main conclusion continues to

    hold; namely, the phase-boundary potentials do not contribute to

    the membrane potential.

    Thus for the thick membrane (membrane thickness

    * LJ

    the membrane potential is zero, whereas for

    the th in membrane (membrane thickness c , ;

    if

    c ,

    > c p ,V,, the potential a t

    I =

    0 )

    would change sign. T hat is, th e plots for

    u

    >

    u

    and

    u < u

    would

    be interchanged.

    current-voltage characteristic of thi s circuit for u

    =

    u,

    u

    > u, and u oes

    change during

    the

    tran-

    FIG.

    21. Concentration profiles within the membrane for the

    bi-ionic case of a membrane separating 0.1

    M

    NaCl from 0.1

    M

    KCI. A , I = 0; B , I is large and positive; C, I is large and negative.

    10.1M NaCl

    ? I

    FIG. 22. Sketch of the steady-state current-voltage character-

    istic for the bionic case of Fig. 21. Note the characteristics(dashed

    l ines) for a membrane separating either symmetrical 0.1 M NaCl

    solutions or symmetrical 0.1 M KCl solutions, which the bi-ionic

    characteristic approaches asymptotically for large positive and

    negative potentials, respectively. Also shown are the chord and

    slope resistance lines at point P .

  • 7/25/2019 Physical Principles and Formalisms of Electrical Excitability.

    18/53

    178 HANDBOOK OF

    PHYSIOLOGY

    - THE

    NERVOUS

    SYSTEM I

    T

    t I

    F I G . 2 3 .

    Nonlinear steady-state current-voltage characteristic

    with chords drawn at

    I , ,

    9,)nd

    I x ,

    qJ. f current is stepped

    fro ml, toZ,, the voltage originally attains the value a t P and then

    increases with ti me along the vertical line from

    P

    until it reaches

    the value q 2 . imilarly, if curre nt is stepped from I , to I,, the

    voltage originally attains the value at Q and then moves along

    the vertical line from Q until it reaches the value 9,.

    sient from one steady state to another (81.1

    The current-voltage characteristic in Figure 22 also

    illustrates several concepts tha t will be important for

    our later discussion of excitable systems: the system

    showsrectification; tha t is, for voltages (i.e., values of

    q

    -

    of the same magnitude but opposite sign, the

    (absolute values of the) currents a re

    not

    equal. Also,

    as for any circuit element with a nonlinear current-

    voltage characteristic, the term

    resistance

    is ambigu-

    ous and has a t least two reasonable meanings. One is

    the chord resistance given by

    (58)

    In this case, since

    \If,,

    s not voltage dependent, the

    chord resistance is identical to the integral resist-

    ance. Another equally valid meaning for resistance is

    the

    slope resistance

    given by

    (59)

    Both of these terms are illustrated in Figure

    22

    at a

    given value

    of

    q.

    n discussing membrane resistance

    for a nonlinear system, one must always specify the

    resistance referred to.

    Of

    course,

    for

    linear current-

    voltage characteristics, such as in Figure 19, the two

    resistances are identical, and no distinction need be

    made.

    It is very instructive to consider how the system

    passes i n t im e from one steady-state point to another

    (Fig. 23, which is a redrawing of the characteristic in

    Fig. 22). Suppose we are sitting at the point ( I I ,PI)

    and suddenly step the current up to I,. The ins tanta-

    neous voltage will be given by I , times the chord

    resistance at

    I , ,

    since the ionic profiles have not had

    time

    to

    change. As they rearrange themselves, the

    resistance rises (Na+

    is

    replacing K + in the mem-

    brane) and the voltage increases until it attains the

    steady-state value for I,. Similarly, if the current is

    suddenly stepped from

    I ,

    to

    14,

    he instantaneous

    voltage will be given by

    I ,

    times the chord resistance

    at

    I,,

    and then the (absolute value of the ) voltage will

    decrease

    as the ionic profiles rearrange. The re-

    sponses are schematized in Figure 24.

    We see that the membrane resistance is not only

    nonlinear, it is also

    t ime uariant .

    The time depend-

    ence enters because of the time required for the ionic

    profiles

    to

    shift from one steady-state configuration

    to

    another.

    As

    the profiles are shifting, the integral

    resistance is continuously changing. (Y,, is also con-

    tinuously changing, but we are neglecting that in

    this discussion.) Note also th at the response in Figure

    24A is phenomenologically similar to the response of

    the RC network in Figure 25A, whereas the re-

    sponse in Figure

    24B

    is phenomenologically similar

    to

    the response of the RL network in Figure 2%. It

    can be shown formally tha t the AC-impedance char-

    acteristics of time-variant resistances can display

    phenomenological capacitances and inductances (42).

    To

    summarize the time-variant aspect of the system:

    the instantaneous response to a change in current (or

    voltage) is along the chord resistance; the voltage (or

    1

    A

    I

    F I G . 24. The change of voltage with time in response to steps

    of cur ren t. These responses follow from the a nalysis of Fig. 23 as

    given in the legend and text.

    P ?

    FIG. 25. RC

    ( A )

    and RL ( B ) networks that would give re-

    sponses similar to those shown in Fig. 24.

  • 7/25/2019 Physical Principles and Formalisms of Electrical Excitability.

    19/53

    CHAPTER 6: PHYSICAL PRINCIPLES AND FORMALISMS O F

    ELECTRICAL EXCITABILITY

    179

    current) then relaxes to the point on the steady-state

    current-voltage characteristic appropriate for the

    new current (or voltage).

    We have been discussing th e behavior of a membrane

    whose only property is tha t

    it

    allows stir ring on both

    sides for maintenance of boundary conditions; other-

    wise the membrane imposes no new restrictions on

    ion movement. Our real example has been coarse

    filter paper. We now wish to consider two examples

    in which the membrane phase is not merely a con-

    fined salt solution of composition grading between

    that of the two aqueous phases bathing the mem-

    brane. After discussing the two examples, we shall

    comment on the general situation of membranes with

    special properties.

    Fixed-charge membrane. We have already consid-

    ered the quasi-equilibrium situation th at arises when

    the fixed-charge density of

    a

    membrane is very large

    and the membrane separates a single salt at two

    different concentrations. We now extend the treat-

    ment of fixed-charge membranes to the case of lower

    charge densities. Since the co-ion is no longer ex-

    cluded, transport (when

    Z

    = 0) of ions and sa lt cannot

    be neglected. (Even in membranes of large fixed-

    charge density, significant fluxes of counter ions oc-

    cur; e.g., for the bi-ionic case of NaCl vs. KCl sepa-

    rated by

    a

    high-density negatively charged mem-

    brane, there is, a t

    Z

    = 0, a flux of Na+ from compart-

    ment

    1

    o compartment 2, and an equal and opposite

    flux of K+ from

    2

    to

    1.

    Our treatment will of course

    also apply to these cases.) We do not go through the

    details of the analysis, but merely point out the high-

    lights of the principles involved.

    The general analysis of fixed-charge membranes is

    a straightforward combination of the Donnan equilib-

    rium and the flux equations

    (71). It

    is assumed that

    the interfacial processes are rapid compared to trans-

    port through the membrane interior; hence one as-

    sumes that equilibrium is maintained a t each inter-

    face a t al l times (even in the face of curren t flow). For

    the fixed-charge membrane, this means that the

    Donnan equilibrium is satisfied at each interface.

    Thus th e concentrations of ions just within the mem-

    brane satisfy the Donnan condition with respect to

    the ions just outside the membrane. Subject to these

    new boundary concentrations, the ions then move

    according to the flux equations. Essentially then the

    fixed charge has established new boundary condi-

    tions for the Nernst-Planck regime of ions. Current

    flow will shift the concentration profiles within the

    interior of the membrane, but

    it

    is assumed th at th e

    concentrations Ijust inside the membrane are not

    perturbed by the current. Figure 26A illustrates the

    concentration profiles for the single-salt case in a

    positive fixed-charge membrane. If +I = 0, the

    system reduces to the single-salt case for an un-

    charged membrane, and if +I % [NaCll,,,, the sys-

    HOMOGENEOUS MEMBRANES WITH SPECIAL PROPERTIES.

    I

    I

    ( 1 ) ( m )

    ( 2 )

    FIG.

    26.

    Concentration profiles (A ) and potential profile ( B ) or

    a positive fixed-charge membrane separating NaCl solutions at

    different concentrations. Note the Donnan

    jumps

    in concentra-

    tions and potential at the interfaces.

    tem approaches the quasi-equilibrium case of a mem-

    brane with a very large fixed-charge density, de-

    scribed earlier. Note tha t in th is lat ter case the mem-

    brane is almost exclusively permeable to C1- not

    because the mobility of chloride in the membrane is

    so much larger tha n th at of sodium, but because the

    concentration of chloride is so much greater than that

    of sodium. This illustrates a n important general prin-

    ciple applicable to both electrolytes and nonelectro-

    lytes: the permeability of a membrane for

    a

    species is

    generally dependent on two factors. One is the mobil-

    ity of the species in the membrane phase, and the

    other is the ability of the species to enter the mem-

    brane in the first place. Unless one has other infor-

    mation, i t is impossible to tell which is responsible for

    a

    given molecule being poorly permeable.

    For a fixed-charge membrane, the total membrane

    potential

    is

    the algebraic sum of three terms: two

    Donnan potentials at the interfaces, generally called

    rr, and

    rr2,

    and a Nernst-Planck diffusion potential

    within the membrane, called (GI, - 2m). Thus

    (

    60)

    (The potential profile is illustrated for the single-salt

    case in Fig. 26B.) To find - 2m) one must solve

    the flux equations subject to the new boundary condi-

    tions established by the Donnan equilibria at the

    interfaces

    (71).

    The electroneutrality condition

    within the membrane now takes the form

    C

    cj+ +

    ZN -

    X C ~ -

    0 (61)

    Note tha t for the bi-ionic case, the Donnan potentials

    ml nd

    rr2

    are equal an d hence cancel, so that the

    membrane potential in Equation 60 is given just by

    the internal diffusion potential -

    Oil membrane. The analysis of th is system is com-

    pletely analogous to that of the fixed-charge mem-

    brane, but the equilibria a t the boundaries are now

    partition equilibria, not Donnan equilibria. Subject

    to the new concentrations established just within th e

    membrane, the ions once again diffuse according to

    the Nernst-Planck flux equations. (The concentration

    profile for the single-salt case is illustrated in Fig.

    27.) Once again the membrane potential is the alge-

    braic sum of three potentials: two phase-boundary

    potentials and an inte rnal diffusion potential. Thus

    9 = T(o/u)a +

    T( u /o) z

    + ( I m - 2m)

    (62)

    = T +

    ~2

    + ( 1,

    -

    2m)

  • 7/25/2019 Physical Principles and Formalisms of Electrical Excitability.

    20/53

    180

    HANDBOOK

    O F PHYSIOLOGY

    THE

    NERVOUS SYSTEM I

    As we pointed out earlier (see subsection

    Quasi-egui-

    Zibrium Systems.

    AN OIL

    MEMBRANE),

    the phase-

    boundary potentials are equal for the singe-salt case

    and hence cancel. Thus, despite the relative solubili-

    ties of the anion and cation in the oil phase, the

    membrane potential will only be a function of their

    relative mobilities in this phase. Therefore, for the

    single-salt case, Equation 62 reduces to

    u - u R T cZrn

    In

    J

    =--

    u + u F c,,,

    On the other hand, the phase-boundary potentials

    will in general be different for the bi-ionic case, and

    all three terms in Equation 62 will contribute to

    II/.

    General considerations.

    The two examples we have

    jus t given illustrate the general approach taken with

    membrane transport. One assumes that the mem-

    brane can conceptually be divided into three regions:

    two interfacial regions and an interior region. It is

    assumed that interfacial processes are

    so

    fast that

    equilibrium conditions prevail there. These processes

    establish new boundary conditions Ijust within the

    membrane, and transport then takes place through

    the membrane interior according to the flux equa-

    tions. (This analysis incidentally is equally applica-

    ble to nonelectrolyte transport.) The membrane po-

    tential is the algebraic sum of two equilibrium poten-

    tia ls and one diffusion potential. Since the membrane

    substance is discontinuous at the interface, there

    exist discontinuities in concentrations and electrical

    potential at each interface, but the assumption of

    equilibrium is equivalent to assuming th at the elec-

    trochemical potential of each species is nevertheless

    continuous across the interface. [For example, in a

    fixed-charge membrane there are jumps in concen-

    tration and electrical potential at the interface (see

    Fig. 261, but the electrochemical potential of each ion

    on the two sides of the interface is the same. In fact,

    these characteristic discontinuities of the Donnan

    equilibrium were derived on the basis of equality of

    electrochemical potential.] Of course, one could be

    more sophisticated and expand these discontinuities

    into the actual space-charge regions present at each

    interface.

    There are two points worth noting about this anal-

    ysis of membrane transport and potentials. Fi rst , it is

    possible that situations might be found where the

    assumption of very fast interfacial processes relative

    I C 1 n I

    1 )

    m )

    ( 2 )

    FIG.

    27. Concentration profile

    for

    an

    oil

    membrane separat-

    ing a single salt at t w o different concentrations.

    to diffusion within the membrane begins to break

    down. It would th en no longer be possible to assume

    equilibrium a t the boundaries, and a n explicit analy -

    sis of the boundary kinetics would have to be included

    in the overall transport. The previous type of analy-

    sis, however, still gives a good qualitative picture

    even in this instance. For example, suppose that

    interfacial events were not fast enough to maintain

    the Donnan conditions. Nevertheless the

    sign

    of the

    change of ion concentrations across the interface can

    be obtained from the Donnan analysis. That is, for a

    positively fixed-charge membrane, t he anion concen-

    tration will be elevated across the interface and the

    cation concentration depressed. The interfacial po-

    tentials will then be somewhat reduced from their

    equilibrium values.

    The second point is that the analysis presupposes

    the possibility

    of

    dividing the membrane up into in-

    terfacial regions and interior. Clearly this becomes a

    very tenuous assumption when the membrane

    is

    of

    the order of

    50-p\

    thickness, as with th e lipid bilayers.

    In fact, the analysis breaks down.

    It

    is no longer

    possible to break up the membrane potential as in

    Equation 62, since the space-charge regions extend

    throughout the membrane. Here, a n explicit analysis

    combining the flux equations and Poissons equation

    with the appropriate boundary conditions must be

    used. It is interesting that for nonelectrolyte trans-

    port across the bilayers (e.g., isotopic water flux), the

    procedure of dividing the membrane into three re-

    gions leads to predictions in good agreement with

    experimental results

    (18).

    It appears that partition

    equilibrium is attained rapidly with respect to diffu-

    sion even through a very thin

    (-50

    A)

    region.

    MOSAIC MEMBRANES.e have been considering mem-

    branes whose properties do not vary in a plane paral-

    lel to the membrane surfaces; tha t

    is,

    the membranes

    are homogeneous. We now wish to comment on mo-

    saic membranes consisting of regions with different

    permeability characteristics. For example, the mem-

    brane may contain some regions permselective for

    Na +, other regions permselective for

    K + ,

    and still

    others relatively nonselective among univalent ions.

    As we shall see later, such membranes are of direct

    relevance to biological membranes, particularly ex-

    citable membranes, and can be experimentally real-

    ized with modified thin lipid membranes. At this

    point we are not concerned with t he mechanisms for

    the selectivities, but accept them as given. (If the

    membrane consisted of cation and anion selective

    regions, we could imagine that each region was a

    patch of high fixed-charge density ion-exchange

    resin; such membranes can be fabricated with ion-

    exchange beads imbedded in an inert matrix .)

    For a membrane with regions in parallel,

    the

    equivalent circuit has the form shown in Figure 28.

    (For completeness we have included the membrane

    capacitance, which we discuss in a later section.)

  • 7/25/2019 Physical Principles and Formalisms of Electrical Excitability.

    21/53

    CHAPT E R 6:

    PHYSICAL PRINCIPLES

    AND FORMALIS MS OF ELECTRICAL EXCITABILITY 181

    I

    - a

    '.i

    7

    t

    t 1

    1

    FIG

    28.

    Equivalent circuit for

    a

    mosaic membrane. The indi-

    vidual elements are shown to

    be

    ideally selective

    for

    a given ion,

    but this need not necessarily be the case.

    Here we have indicated that in principle the conduct-

    ances (g,)can be voltage dependent. (In treating par-

    allel circuits, it is more convenient to use conduct-

    ances than resistances.) The emfs , however, are in-

    variant, being determined by the concentrations of

    ions on the two sides of the membrane and the perme-

    ability characteristics of the regions. If the regions

    are permselective, as in Figure 28, each emf is the

    Nernst potential for the permeant ion of tha t region.

    This circuit should be contrasted to the one for the

    homogeneous membrane given in Figure 16. In that

    case there is a single conductance and one emf, both

    of which can be voltage dependent. For the mosaic

    membrane, we have from elementary circuit theory

    z,

    = gj 9

    Ej)

    (63)

    for each element of the circuit. Summing over all

    elements, we have from Kirchhoff

    s

    law

    where

    z = T Z I (65)

    Equation 64 bears the same relation to the mosaic

    membrane as Equation 38 bears to the homogeneous

    membrane. Comparing the two, we see that each

    equation has an Ohm's law term and also another

    term that may be called the diffusion potential. In a

    formal sense the two expressions are quite similar.

    Although the flux equations describe the move-

    ment of ions across the membrane, the equivalent

    circuit of Figure 16 does not indicate that there is ion

    flux when Z

    =

    0. In contrast, the equivalent circuit of

    Figure 28 depicts the local currents th at flow through

    the elements of the membrane even when no current,

    I ,

    is being passed across the membrane. Under these

    circumstances, the membrane potential is given ex-

    clusively by the second term in Equation 64, whereas

    Equation 65 becomes

    zzj = 0

    j

    It

    should be understood that the dissipative pro-

    cesses for the homogeneous and mosaic membranes

    are quite different. In the former the ions flow

    through a common region, and hence there are no

    local current flows. In the mosaic membrane ion

    movement is through local currents. Despite these

    physical differences,

    it

    happens tha t the

    steady-state

    properties of a homogeneous membrane can also be

    formally represented by the circuit in Figure

    28 (20).

    FORMAL CONSEQUENCES OF VOLTAGE-

    DEPENDENT CONDUCTANCES

    The N a ture

    of

    Electrical Excitability

    In

    the

    subsection

    Ion Transport (the Nern st-

    Planck Flux Equations) we developed the properties

    of a homogeneous membrane and showed that even

    with such a simple membrane as coarse filter paper,

    it

    is

    possible to observe rectification, nonlinearities,

    and time transien ts in th e membrane potential. Since

    these properties are also characteristic of electrically

    excitable biological membranes, the question arises

    whether the mechanisms operating in the simple

    systems we discussed are also responsible for gener-

    ating action potentials and bioelectric phenomena.

    We remind the reader th at the nonohmic behavior we

    have described is due to the shifting of the ionic

    profiles of mobile ions within the membrane in the

    face of an applied voltage or current; this change in

    the ionic profiles leads to changes in the membrane

    conductance and the membrane emf. The question

    then is whether such a mechanism could account for

    action potentials of nerve and muscle.

    It

    is very difficult to give a

    definitive

    answer to this

    question. It can probably be proved that

    it

    is impossi-

    ble to reproduce electrical excitability with a filter

    paper membrane, even with the most bizarre mixture

    of ions on the two sides of the membrane. This does

    not preclude the possibility that a membrane with a

    fixed-charge density varying as arc coth x3", or one

    with some continuously varying function of the die-

    lectric constant of

    its

    ''oil" interior, or a combination

    of these st ruc tures could not generate action poten-

    tials. Although no one has demonstrated, either theo-

    retically or experimentally, that one can develop an

    excitable membrane simply from electrophoretically

    moving ionic profiles within the membrane, neither

    has anyone presented proof that it is impossible to do

    so. [In fact, if one allows the ionic profiles to be

    shifted not only by the electric field (i.e., electropho-

    resis) but also by solvent drag produced by electro-

    osmosis and hydrostatic pressure gradients, then one

    can indeed develop, with a membrane made merely of

    sintered glass, oscillatory behavior and complex all-

    or-none responses th at ar e in some ways phenomeno-

    logically similar to excitable tissue (72, 73).] There

    are nevertheless strong indications for believing th at

    such

    a

    mechanism cannot be the basis for biological

  • 7/25/2019 Physical Principles and Formalisms of Electrical Excitability.

    22/53

    182

    HANDBOOK

    O F PHYSIOLOGY

    -

    THE NERVOUS

    SYSTEM I

    excitability, and indeed there is compelling evidence

    that excitable membranes are not homogeneous, but

    rather mosaic structures. We shall discuss the bases

    for the above assertions and a t the same time present

    the current view of the na ture of biological excitable

    systems.

    REASONS F O R

    BELIEVING

    THAT ELECTRICAL EXCITAB IG

    ITY

    DOES NOT RESULT

    FROM THE SHIFTING

    OF IONIC

    PROFILES. Th e magn i tude of rect if ication. In excitable

    cells such as the squid giant axon, the ratio of the

    conductance a t large depolarizations to the conduct-

    ance at large hyperpolarizations can be several

    hundred to one. Since the major conducting ions are

    Na+ (on the outside) and

    K +

    (in the axoplasm), with

    approximately equal total sal t concentration on each

    side of the membrane, the situat ion is approximately

    bi-ionic. In such a case, rectification is achieved by

    exchanging in the membrane a more mobile ion with

    a less mobile ion, depending on the sign of the poten-

    tial. Assuming tha t the contribution of all other ions

    to the membrane conductance can be neglected (an

    extreme case), the limiting rectification ratio is given

    by the ratio of the sodium and potassium mobilities in

    the membrane (i.e.,

    u K / u N a ) ,

    hich is approximately

    1.5 in free aqueous solution.

    If

    this phenomenon ac-

    counted for the observed rectification ratio of several

    hundred, either the potassium mobility within the

    membrane must be abnormally large, or the sodium

    mobility must be abnormally small, or both.

    Th e t ime scale of the a ct ion potent ia l . In our discus-

    sion of trans ien t behavior during ionic profile shifts,

    we omitted the time scale. Since the relaxation of the

    profiles is basically a diffusion process, the times

    involved will be roughly those required for the root

    mean square displacement

    (?)I/*

    of an ion t o be equal

    to the membrane thickness. This

    is

    given by the well-

    known result from Brownian motion theory

    (13,

    14)

    x2 = 2Dt (66)

    In aqueous solution, the diffusion coefficient, D,

    for

    small ions such as Na+ and K+ is of the order

    cm2/s. Taking the membrane thickness a s

    100

    A

    (lo- ';

    cm), we find from Equation

    66

    that the time involved

    for the displacement of concentration profiles is of the

    order of s, that

    is,

    0.1 p s . The events associated

    with a n action potential, occurring

    as

    they do on the

    millisecond time scale, are four orders of magnitude

    slower. We would therefore have to postulate that t he

    diffusion constants (i .e. , mobilities) of ions within the

    membrane are at least a factor of

    lo4

    times smaller

    th an t he mobilities of ions in free solution.

    Dissociation

    of

    sod ium and po tass ium conduc t -

    ances. One of the distinguishing features of the

    permeability changes occurring during a n action po-

    tential (or under voltrage-clamp conditions) is th e dif-

    ferent time course of these changes for different ions.

    In th e squid giant axon, for example, sodium permea-

    bility rises and begins to fall before potassium perme-

    -

    ability changes significantly. In fact, Hodgkin and

    Huxley's analysis of excitability in the giant axon

    was possible only because they could clearly resolve

    and separate the time course of the sodium and potas-

    sium conductance changes (29). Since then, t he belief

    that these components are separate has been rein-

    forced with the discovery of pharmacological agents

    that inhibit one, but not both, conductances. For

    example, tetrodotoxin (TTX), the puffer fish poison,

    can completely and reversibly block the sodium con-

    ductance changes without affecting the magnitude

    and the time course of the potassium conductance

    transients (49). Conversely, tetraethylammonium

    (TEA) reduces the potassium conductance without

    significantly affecting the sodium transients (2) .

    It is difficult t o imagine a homogeneous regime

    of ions where th e permeability changes for one ion are

    not intimately coupled with those of the other ions. In

    the bi-ionic case tha t we considered, for example, th e

    movement of Na+ into or ou t of th e membrane was

    accompanied by the obligatory and opposite move-

    ment of K + . It

    is

    even more difficult t o conceive of an

    agent, acting in a homogeneous membrane, that

    could inhibit transients associated with one ion and

    yet not affect those associated with the other ions in

    the membrane. For these reasons, virtually all elec-

    trophysiologists believe that excitable membranes

    are mosaic structures of the form described in the

    subsection

    MOSAIC

    MEMBRANES, and although Hodg-

    kin and Huxley did not explicitly stat e this in their

    basic papers (29-331, they clearly had this picture in

    mind. Certainly the simplest way in which the so-

    dium and potassium conductances could be function-

    ally independent is for the regions of sodium and

    potassium permeability to be physically separated.

    HODCKIN-HUXLEY EQUIVALENT CIRCUIT. nalysis of

    electrical excitability of the squid giant axon mem-

    brane, and any other excitable membrane, starts

    with the equivalent circuit for a mosaic membrane of

    the type shown in Figure 28. The ions involved in the

    circuit can vary depending on the particular biologi-

    cal membrane

    o r

    the ions in the solution that bathes

    the membrane.

    Because much is known about the

    squid axon, we use this system as representative of

    the others, with the understanding t hat the pr inc i -

    ples operative there are generally believed to apply to

    all excitable membranes. For the squid giant axon in

    normal seawater , the equivalent circuit of Figure 28

    takes the particular form shown in Figure 29 (32) .

    The sodium and potassium conductances are voltage

    dependent, whereas the leakage conductance (g,)-

    probably a lumping together of several ion permea-

    bilities- s constant. The sodium and potassium

    emf's are,

    of

    course, given by the Nernst potentials

    for these ions, and are assumed to be constant, be-

    cause the concentrations of these ions in seawater

    and axoplasm are constant (unde r the experimental

    conditions). Since [Na+loutside [Na+linside nd

  • 7/25/2019 Physical Principles and Formalisms of Electrical Excitability.

    23/53

    CHAPTER 6: PHYSICAL PRINCIPLES

    AND

    FORMALISMS OF ELECTRICAL EXCITABILITY

    183

    O u t s i d e ( s e a w a t e r )

    c

    I n s id e ( o x o p l a s m )

    FIG

    29. The Hodgkin-Huxley equivalent circuit for the squid

    giant axon membrane in normal seawater.

    [K+loutside < [K+linside>

    ENa and EK are of opposite

    polarity. Ultimately this system would run down

    (Na+coming in and K + going out), unless some coun-

    terbalancing process exists. This energy-requiring

    process

    is

    the role assigned to metabolism. The leak-

    age emf

    (E,)s,

    like the leakage conductance, proba-

    bly a lumped parameter. Although the leakage ele-

    ment modifies the behavior of the system somewhat,

    the dominating factors are the voltage-dependent so-

    dium and potassium conductance elements; for the

    most part, therefore, we confine o u r comments to

    these and ignore the leakage element.

    The Hodgkin-Huxley description of the action po-

    tent ial is briefly summarized as follows: in the rest-

    ing state,

    g,

    >>gNa o tha t the resting potential sits

    near EK.At threshold depolarization,

    g N a

    has in-

    creased enough tha t the membrane furthe r depolar-

    izes. This leads to a further increase in g N a , which

    leads to further depolarization, and so on. The net

    result of thi s process is the rising phase of the action

    potential, which tends toward EN, a t

    its

    peak. The

    falling phase results from a turning off of g,, (called

    sodium inactivation) and a turning on of

    g K ;

    both

    processes act to move the membrane potential back to

    EK.

    Regenerative behavior develops because the

    membrane potential is determined by the relative

    conductances (permeabilities) of the membrane to

    potassium and sodium [if we neglect the capacitance

    current, th e membrane potential is given at all times

    by the last ter m in Equation

    64,

    which in this case is

    (gNaEXa

    +

    g,EK)/(gNa

    +

    gK)],

    while a t the same time

    these conductances are functions of the membrane

    potential. The functional dependence ofg,, and

    g,

    on

    membrane potential is determined in voltage-clamp

    experiments. Tha t is, the membrane potential is held

    fixed (via external electrodes) at various levels and

    the transient changes in the sodium and potassium

    conductances recorded (29-31, 33).

    The physics underlying t he voltage-dependent con-

    ductances is still not understood; it

    is

    a major un-

    solved problem in electrophysiolo