2
Dr. Murtadha Al-Shareifi e-Library
Plasma Membrane,
Membrane Transport,
and Resting Membrane
Potential
AC TIVE
L EARNING
Cellular Physiology
Upon mastering the material in this chapter you should be
able to:
• Understand how proteins and lipids are assembled
to form a selectively permeable barrier known as the
plasma membrane.
• Explain how the plasma membrane maintains an internal
environment that differs significantly from the extracellular fluid.
• Understand how voltage-gated channels and ligandgated channels are opened.
• Explain how carrier-mediated transport systems differ
from channels.
• Understand the importance of adenosine
triphosphate–binding cassette transporters to
T
he intracellular fluid of living cells, the cytosol, has a
composition very different from that of the extracellular fluid (ECF). For example, the concentrations of
potassium and phosphate ions are higher inside cells than
outside, whereas sodium, calcium, and chloride ion concentrations are much lower inside cells than outside. These
differences are necessary for the proper function of many
intracellular enzymes; for instance, the synthesis of proteins by the ribosomes requires a relatively high potassium
concentration. The plasma membrane of the cell creates
and maintains these differences by establishing a permeability barrier around the cytosol. The ions and cell proteins
needed for normal cell function are prevented from leaking
out; those not needed by the cell are unable to enter the cell
freely. The plasma membrane also keeps metabolic intermediates near where they will be needed for further synthesis
or processing and retains metabolically expensive proteins
inside the cell.
The plasma membrane is necessarily selectively permeable. Cells must receive nutrients to function, and they
must dispose of metabolic waste products. To function in
coordination with the rest of the organism, cells receive and
send information in the form of chemical signals, such as
24
•
•
•
•
•
O BJ ECTIVES
lipid transport and development of multidrug
resistance.
Explain the difference between primary and secondary
active transport.
Explain how epithelial cells are organized to
produce directional movement of solutes
and water.
Explain how many cells can regulate their volume when
exposed to osmotic stress.
Understand why the Goldman equation gives the value
of the membrane potential.
Understand why the resting membrane potential
in most cells is close to the Nernst potential
for K+.
hormones and neurotransmitters. The plasma membrane has
mechanisms that allow specific molecules to cross the barrier
around the cell. A selective barrier surrounds not only the cell
but also every intracellular organelle that requires an internal
milieu different from that of the cytosol. The cell nucleus,
mitochondria, endoplasmic reticulum, Golgi apparatus, and
lysosomes are delimited by membranes similar in composition to the plasma membrane. This chapter describes the
specific types of membrane transport mechanisms for ions
and other solutes, their relative contributions to the resting
membrane electrical potential, and how their activities are
coordinated to achieve directional transport from one side
of a cell layer to the other.
●
PLASMA MEMBRANE
STRUCTURE
The first theory of membrane structure proposed that cells
are surrounded by a double layer of lipid molecules, a lipid
bilayer. This theory was based on the known tendency of
lipid molecules to form lipid bilayers with low permeability to water-soluble molecules. However, the lipid bilayer
theory did not explain the selective movement of certain
Dr. Murtadha Al-Shareifi e-Library
●
25
Chapter 2 / Plasma Membrane, Membrane Transport, and Resting Membrane Potential
Extracellular fluid
Glycoprotein
Integral proteins
Sphingolipid
Hormone
receptor
Glycolipid
TK TK
Phospholipid
Cholesterol
Channel
Peripheral protein
Lipid raft
Cytoplasm
water-soluble compounds, such as glucose and amino acids,
across the plasma membrane. In 1972, Singer and Nicolson
proposed the fluid mosaic model of the plasma membrane,
which described the organization and interaction of proteins
with the lipid bilayer (Fig. 2.1). With minor modifications,
this model is still accepted as the correct picture of the structure of the plasma membrane.
Plasma membrane consists of different
types of membrane lipids with different
functions.
Lipids found in cell membranes can be classified into two
broad groups: those that contain fatty acids as part of the
lipid molecule and those that do not. Phospholipids are an
example of the first group, and cholesterol is the most important example of the second group.
Phospholipids
The fatty acids present in phospholipids are molecules with
a long hydrocarbon chain and a carboxyl terminal group.
The hydrocarbon chain can be saturated (no double bonds
between the carbon atoms) or unsaturated (one or more
double bonds present). The long hydrocarbon chain tends to
avoid contact with water and is described as hydrophobic.
The carboxyl group at the other end is compatible with water
and is termed hydrophilic. Fatty acids are said to be amphipathic because both hydrophobic and hydrophilic regions
are present in the same molecule.
Phospholipids are the most abundant complex lipids
found in cell membranes. They are amphipathic molecules
formed by two fatty acids (normally, one saturated and one
unsaturated) and one phosphoric acid group substituted on
the backbone of a glycerol or sphingosine molecule. This
arrangement produces a hydrophobic area formed by the two
fatty acids and a polar hydrophilic head. When phospholipids
are arranged in a bilayer, the polar heads are on the outside
and the hydrophobic fatty acids on the inside (see Fig. 2.1).
It is difficult for water-soluble molecules and ions to pass
directly through the hydrophobic interior of the lipid bilayer.
● Figure 2.1 The fluid mosaic model of
the plasma membrane. Lipids are arranged
in a bilayer. Cholesterol provides rigidity to the
bilayer. Integral proteins are embedded in the
bilayer and often span it. Some membranespanning proteins form pores and channels;
others are receptors. Peripheral proteins do
not penetrate the bilayer. Lipid rafts form
stable microdomains composed of sphingolipids and cholesterol.
The phospholipids, with a backbone of sphingosine
(a long amino alcohol), are usually called sphingolipids and
are present in all plasma membranes in small amounts.
They are especially abundant in brain and nerve cells. Ceramide is a lipid second messenger that is generated from the
sphingolipid sphingomyelin (see Chapter 1, “Homeostasis
and Cellular Signaling,” for a discussion of lipid secondmessenger signaling).
Glycolipids are lipid molecules that contain sugars and
sugar derivatives (instead of phosphoric acid) in the polar
head. They are located mainly in the outer half of the lipid
bilayer, with the sugar molecules facing the ECF. Proteins
can associate with the plasma membrane by linkage to the
extracellular sugar moiety of glycolipids.
Cholesterol
Cholesterol is an important component of mammalian
plasma membranes. The proportion of cholesterol in plasma
membranes varies from 10% to 50% of total lipids. Cholesterol has a rigid structure that stabilizes the cell membrane
and reduces the natural mobility of the complex lipids in
the plane of the membrane. Increasing amounts of cholesterol make it more difficult for lipids and proteins to move in
the membrane. Some cell functions, such as the response of
immune system cells to the presence of an antigen, depend
on the ability of membrane proteins to move in the plane of
the membrane to bind the antigen. A decrease in membrane
fluidity resulting from an increase in cholesterol will impair
these functions.
Lipid microdomains
Aggregates of sphingolipids and cholesterol can form stable microdomains termed lipid rafts that diffuse laterally in
the phospholipid bilayer. The protein caveolin is present in
a subset of lipid rafts (termed caveolae), causing the raft to
form a cavelike structure. It is believed that one function of
both noncaveolar and caveolar lipid rafts is to facilitate interactions between specific proteins by selectively including
(or excluding) these proteins from the raft microdomain.
●
Dr. Murtadha Al-Shareifi e-Library
26
Part I / Cellular Physiology
Clinical Focus / 2.1
Hereditary Spherocytosis
Red blood cells (erythrocytes) contain hemoglobin, which
binds oxygen for transport from lungs to other organs and tissues. A normal erythrocyte has a marked biconcave structure,
which permits the cell to deform for easier passage through
narrow capillaries, especially in the spleen. The unusual shape
is maintained by the cytoskeleton, formed by both integral
and peripheral plasma membrane proteins. A key player is the
long filamentous protein known as spectrin, a peripheral protein that forms a meshwork on the cytoplasmic surface. Linkages between spectrin and several integral proteins serve to
anchor the plasma membrane to the cytoskeleton. Hereditary
spherocytosis (HS) is a genetic disease that affects proteins
in the erythrocyte membrane, and the result is a defective
cytoskeleton. The incidence of HS is 1 in 5,000. The most
common defect is deficiency of spectrin, and the result is that
regions of the membrane break off because they are no longer anchored to the cytoskeleton. The remaining membrane
reseals spontaneously, but after several “sheddings,” the cell
eventually becomes small and spherical. The presence of
these microspherocytes is the hallmark of HS disease, which
can range from asymptomatic to severe hemolytic anemia.
Hemolysis (cell bursting) is present because the spherocytes
are fragile to osmotic stress. Any entry of water will increase
intracellular volume, and the cell membrane will break and
release the hemoglobin. In contrast, normal biconcave eryth-
For example, lipid rafts can mediate the assembly of membrane
receptors and intracellular signaling proteins as well as the
sorting of plasma membrane proteins for internalization.
Proteins are integrally and peripherally
associated with the plasma membrane.
Proteins are the second major component of the plasma
membrane, present in about equal proportion by weight
with the lipids. Two different types of proteins are associated
with the plasma membrane. Integral proteins (or intrinsic proteins) are embedded in the lipid bilayer; many span
it completely, being accessible from the inside and outside
of the membrane. The polypeptide chain of these proteins
may cross the lipid bilayer once or may make multiple passes
across it. The membrane-spanning segments usually contain
amino acids with nonpolar side chains and are arranged in
an ordered a-helical conformation. Peripheral proteins (or
extrinsic proteins) do not penetrate the lipid bilayer. They are
in contact with the outer side of only one of the lipid layers—
either the layer facing the cytoplasm or the layer facing the
ECF (see Fig. 2.1). Many membrane proteins have carbohydrate molecules, in the form of specific sugars, attached to
the parts of the proteins that are exposed to the ECF. These
molecules are known as glycoproteins. Some of the integral membrane proteins can move in the plane of the membrane, like small boats floating in the “sea” formed by the
lipid bilayer. Other membrane proteins are anchored to the
cytoskeleton inside the cell or to proteins of the extracellular
matrix.
rocytes can swell without bursting to accommodate osmotic
water entry. Hence the lifespan of a spherocyte is considerably shorter than the 90- to 120-day lifespan of an erythrocyte.
The shorter life is often compensated by accelerated production of new red blood cells (erythropoiesis) in an otherwise
healthy individual. However, the anemia may be severe during
an illness involving fever because erythropoiesis is slowed by
fever. The anemia is usually the reason that HS patients complain of tiredness or loss of stamina and shortness of breath
during exercise. Splenomegaly (enlarged spleen) is present
in 75% of HS patients. The spherocytes cannot deform their
shape, and they are trapped and destroyed (by hemolysis)
in the spleen, which may be the reason it increases in size.
Removal of spherocytes by the spleen is a major cause of
the anemia and the shortened lifespan. Surgery to remove the
spleen is often recommended if the anemia is severe and will
improve spherocyte survival significantly. This enables most
patients with HS to maintain a normal hemoglobin level.
Treatment of patients with HS has produced an appreciation of the role of the spleen in maintaining the integrity of
the red blood cell population. Laboratory investigation of HS
has helped the understanding of the structure and function
of the membrane cytoskeleton, which is present in almost all
types of cells, and the point mutations that produce defective
spectrin have been identified.
The proteins in the plasma membrane play a variety
of roles. Many peripheral membrane proteins are enzymes,
and many membrane-spanning integral proteins are carriers or channels for the movement of water-soluble molecules
and ions into and out of the cell. Another important role of
membrane proteins is structural; for example, certain membrane proteins in the erythrocyte help maintain the biconcave shape of the cell. Finally, some membrane proteins serve
as highly specific receptors on the outside of the cell membrane to which extracellular molecules, such as hormones,
can bind. If the receptor is a membrane-spanning protein, it
provides a mechanism for converting an extracellular signal
into an intracellular response.
●
SOLUTE TRANSPORT
MECHANISMS
All cells must import oxygen, sugars, amino acids, and small
ions and export carbon dioxide, metabolic wastes, and secretions. At the same time, specialized cells require mechanisms
to transport molecules such as enzymes, hormones, and
neurotransmitters. The movement of large molecules is carried out by endocytosis and exocytosis, that is, the transfer
of substances into or out of the cell, respectively, by vesicle
formation and vesicle fusion with the plasma membrane.
Cells also have mechanisms for the rapid movement of ions
and solute molecules across the plasma membrane. These
mechanisms are of two general types: passive transport,
which requires no direct expenditure of metabolic energy,
Dr. Murtadha Al-Shareifi e-Library
●
27
Chapter 2 / Plasma Membrane, Membrane Transport, and Resting Membrane Potential
and active transport, which uses metabolic energy to move
solutes across the plasma membrane.
Import of extracellular materials occurs
through phagocytosis and endocytosis.
Phagocytosis is the ingestion of large particles or microorganisms, usually occurring only in specialized cells such
as macrophages (Fig. 2.2). An important function of macrophages is to remove invading bacteria from the body.
The phagocytic vesicle (1–2 mm in diameter) is almost as
large as the phagocytic cell itself. Phagocytosis requires
a specific stimulus. It occurs only after the extracellular
particle has bound to the extracellular surface. The particle is then enveloped by expansion of the cell membrane
around it.
Endocytosis is a general term for the process in which
a region of the plasma membrane is pinched off to form an
endocytic vesicle inside the cell. During vesicle formation,
some fluid, dissolved solutes, and particulate material from
the extracellular medium are trapped inside the vesicle and
internalized by the cell. Endocytosis produces much smaller
endocytic vesicles (0.1–0.2 mm in diameter) than phagocytosis. It occurs in almost all cells and is termed a constitutive
process, because it occurs continually and specific stimuli are
not required. In further contrast to phagocytosis, endocytosis originates with the formation of depressions in the cell
membrane. The depressions pinch off within a few minutes
after forming and give rise to endocytic vesicles inside the
cell.
Two types of endocytosis can be distinguished (see
Fig. 2.2). Fluid-phase endocytosis is the nonspecific uptake
of the ECF and all its dissolved solutes. The material is
trapped inside the endocytic vesicle as it is pinched off inside
the cell. The amount of extracellular material internalized by
this process is directly proportional to its concentration in
the extracellular solution. Receptor-mediated endocytosis
is a more efficient process, which uses receptors on the cell
surface to bind specific molecules. These receptors accumulate at specific depressions known as coated pits, so named
because the cytosolic surface of the membrane at this site is
covered with a coat of several proteins. The coated pits pinch
off continually to form endocytic vesicles, providing the cell
with a mechanism for rapid internalization of a large amount
of a specific molecule without the need to endocytose large
volumes of ECF. The receptors also aid the cellular uptake
of molecules present at low concentrations outside the cell.
Receptor-mediated endocytosis is the mechanism by which
cells take up a variety of important molecules, including hormones, growth factors, and serum transport proteins such
as transferrin (an iron carrier). Foreign substances, such as
diphtheria toxin and certain viruses, also enter cells by this
pathway.
Export of macromolecules occurs through
exocytosis.
Many cells synthesize important macromolecules that are
destined for exocytosis or export from the cell. These molecules are synthesized in the endoplasmic reticulum, modified in the Golgi apparatus, and packed inside transport
vesicles. The vesicles move to the cell surface, fuse with the
cell membrane, and release their contents outside the cell
(see Fig. 2.2).
There are two exocytic pathways—constitutive and regulated. Some proteins are secreted continuously by the cells
Endocytosis
Phagocytosis
Extracellular
fluid
Exocytosis
Fluid-phase
endocytosis
Receptor-mediated
endocytosis
Ligand
Receptor
Plasma
membrane
Coated pit
Cytoplasm
● Figure 2.2 The transport of macromolecules across the plasma membrane by the formation
of vesicles. Particulate matter in the extracellular fluid (ECF) is engulfed and internalized by phagocytosis.
During fluid-phase endocytosis, ECF and dissolved macromolecules enter the cell in endocytic vesicles
that pinch off at depressions in the plasma membrane. Receptor-mediated endocytosis uses membrane
receptors at coated pits to bind and internalize specific solutes (ligands). Exocytosis is the release of macromolecules destined for export from the cell. These are packed inside secretory vesicles that fuse with the
plasma membrane and release their contents outside the cell.
●
Dr. Murtadha Al-Shareifi e-Library
28
Part I / Cellular Physiology
that make them. Secretion of mucus by goblet cells in the
small intestine is a specific example. In this case, exocytosis
follows the constitutive pathway, which is present in all cells.
In other cells, macromolecules are stored inside the cell in
secretory vesicles. These vesicles fuse with the cell membrane
and release their contents only when a specific extracellular
stimulus arrives at the cell membrane. This pathway, known
as the regulated pathway, is responsible for the rapid “ondemand” secretion of many specific hormones, neurotransmitters, and digestive enzymes.
Diffusion
High
Membrane
Low
Uncharged solutes cross the plasma
membrane by passive diffusion.
Any solute will tend to uniformly occupy the entire space
available to it. This movement, known as diffusion, is a result
of the spontaneous Brownian (random) movement that all
molecules experience. A drop of ink placed in a glass of water
will diffuse and slowly color all the water. The net result of
diffusion is the movement of substances from regions of high
concentration to regions of low concentration. Diffusion is
an effective way for substances to move short distances.
The speed with which the diffusion of a solute in water
occurs depends on the difference of concentration, the size
of the molecules, and the possible interactions of the diffusible substance with water. These different factors appear in
the Fick law, which describes the diffusion of any solute
in water. In its simplest formulation, the Fick law (also known
as Fick's law) can be written as
J = DA(C1 - C2)/DX
(1)
where J is the flow of solute from region 1 to region 2 in the
solution; D is the diffusion coefficient of the solute, which is
determined by factors such as solute molecular size and interactions of the solute with water; A is the cross-sectional area
through which the flow of solute is measured; C is the concentration of the solute at regions 1 and 2; and DX is the distance between regions 1 and 2. Sometimes, J is expressed in
units of amount of substance per unit area per unit time, for
example, mol/cm2/h, and is also referred to as the solute flux.
The principal force driving the passive diffusion of an
uncharged solute across the plasma membrane is the difference of concentration between the inside and the outside of the
cell (Fig. 2.3). In the case of an electrically charged solute, such
as an ion, diffusion is also driven by the membrane potential,
which is the electrical gradient across the membrane. Movement of charged solutes and the membrane potential will be
discussed in greater detail later in this chapter.
Diffusion across a membrane has no preferential direction; it can occur from the outside of the cell toward the
inside or from the inside of the cell toward the outside. For
any substance, it is possible to measure the permeability
coefficient (P), which gives the speed of the diffusion across
a unit area of plasma membrane for a defined driving force.
The Fick law for the diffusion of an uncharged solute across a
membrane can be written as
J = PA(C1 - C2)
(2)
which is similar to equation 1. P includes the membrane
thickness, the diffusion coefficient of the solute within the
● Figure 2.3 The diffusion of gases and lipid-soluble
molecules through the lipid bilayer. In this example, the
diffusion of a solute across a plasma membrane is driven
by the difference in concentration on the two sides of the
membrane. The solute molecules move randomly by
Brownian movement. Initially, random movement from left to
right across the membrane is more frequent than movement
in the opposite direction because there are more molecules
on the left side. This results in a net movement of solute from
left to right across the membrane until the concentration of
solute is the same on both sides. At this point, equilibrium
(no net movement) is reached because solute movement
from left to right is balanced by equal movement from right
to left.
membrane, and the solubility of the solute in the membrane.
Dissolved gases, such as oxygen and carbon dioxide, have
high permeability coefficients and diffuse across the plasma
membrane rapidly. As a result, gas exchange in the lungs
is very effective. Because diffusion across the plasma membrane usually implies that the diffusing solute enters the
lipid bilayer to cross it, the solute’s solubility in a lipid solvent (e.g., olive oil or chloroform) compared with its solubility in water is important in determining its permeability
coefficient.
A substance’s solubility in oil compared with its solubility in water is its partition coefficient. Lipophilic (lipidsoluble) substances, such as gases, steroid hormones, and
anesthetic drugs, which mix well with the lipids in the
plasma membrane, have high partition coefficients and, as a
result, high permeability coefficients; they tend to cross the
plasma membrane easily. Hydrophilic (water-soluble) substances, such as ions and sugars, do not interact well with the
lipid component of the membrane, have low partition coefficients and low permeability coefficients, and diffuse across
the membrane more slowly.
Solutes such as oxygen readily diffuse across the lipid
part of the plasma membrane by simple diffusion. Thus, the
relationship between the rate of movement and the difference
in concentration between the two sides of the membrane is
linear (Fig. 2.4). The larger the difference in concentration
(C1 − C2), the greater the amount of substance crossing the
membrane per unit time.
Dr. Murtadha Al-Shareifi e-Library
●
29
Chapter 2 / Plasma Membrane, Membrane Transport, and Resting Membrane Potential
Carrier-mediated transport
10
Rate of solute entry (mmol/min)
Rate of solute entry (mmol/min)
Simple diffusion
5
1
2
3
Solute concentration (mmol/L)
outside cell
● Figure 2.4 A graph of solute transport across a
plasma membrane by simple diffusion. The rate of solute
entry increases linearly with extracellular concentration of
the solute. Assuming no change in intracellular concentration, increasing the extracellular concentration increases the
gradient that drives solute entry.
Integral membrane proteins facilitate
diffusion of solutes across the plasma
membrane.
For many solutes of physiologic importance, such as ions,
sugars, and amino acids, the rate of transport across the
plasma membrane is much faster than expected for simple
diffusion through a lipid bilayer. Furthermore, the relationship between transport rate and concentration difference of
these hydrophilic substances follows a curve that reaches
a plateau (Fig. 2.5). Membrane transport with these characteristics is often called facilitated diffusion or carriermediated diffusion, because an integral membrane protein
facilitates (or assists) the movement of a solute through the
membrane. Integral membrane proteins can form pores,
channels, or carriers, each of which facilitates the transport
of specific molecules across the membrane.
There are a limited number of pores, channels, and carriers in any cell membrane; thus, increasing the concentration of the solute initially uses the existing “spare” pores,
channels, or carriers to transport the solute at a higher rate
than by simple diffusion. As the concentration of the solute
increases further and more solute molecules associate with
the pore, channel, or carrier, the transport system eventually
reaches saturation, when all the pores, channels, and carriers are involved in translocating molecules of solute. At this
point, additional increases in solute concentration do not
increase the rate of solute transport (see Fig. 2.5).
The types of integral membrane protein transport
mechanisms considered here can transport a solute along
its concentration gradient only, as in simple diffusion. Net
movement stops when the concentration of the solute has the
same value on both sides of the membrane. At this point, with
reference to equation 2, C1 = C2 and the value of J is 0. The
transport systems function until the solute concentrations
10
Vmax
5
1
2
Solute concentration (mmol/L)
outside cell
3
● Figure 2.5 A graph of solute transport across a
plasma membrane by facilitated diffusion. The rate of transport is much faster than that of simple diffusion (see Fig. 2.4)
and increases linearly as the extracellular solute concentration
increases. The increase in transport is limited, however, by the
availability of channels and carriers. Once all are occupied by
solute, further increases in extracellular concentration have no
effect on the rate of transport. A maximum rate of transport
(Vmax) is achieved that cannot be exceeded.
have equilibrated. However, equilibrium is attained much
faster than with simple diffusion.
Membrane pores
A pore provides a conduit through the lipid bilayer that is
always open to both sides of the membrane. Aquaporins in
the plasma membranes of specific kidney and gastrointestinal tract cells permit the rapid movement of water. Within
the nuclear pore complex, which regulates movement of
molecules into and out of the nucleus, is an aqueous pore
that only allows the passive movement of molecules smaller
than 45 kDa and excludes molecules larger than 62 kDa. The
mitochondrial permeability transition pore and mitochondrial voltage-dependent anion channel (VDAC),
which cross the inner and outer mitochondrial membranes,
promote mitochondrial failure when formed, resulting in the
generation of reactive oxygen species and cell death.
Gated channels
Small ions, such as Na+, K+, Cl−, and Ca2+, cross the plasma
membrane faster than would be expected based on their partition coefficients in the lipid bilayer. The electrical charge of
an ion makes it difficult for the ion to move across the lipid
bilayer. The rapid movement of ions across the membrane,
however, is an aspect of many cell functions. The excitation
of nerves, the contraction of muscle, the beating of the heart,
and many other physiologic events are possible because of
the ability of small ions to enter or leave the cell rapidly. This
movement occurs through selective ion channels.
Ion channels are composed of several polypeptide
subunits that span the plasma membrane and contain a gate
●
Dr. Murtadha Al-Shareifi e-Library
30
Part I / Cellular Physiology
Number of channels open
Lipid
bilayer
0
Selectivity
filter
1
2
3 pA
3
100 msec
Figure 2.7 A patch clamp recording from a frog
muscle fiber. Ions flow through the channel when it opens,
generating a current. The current in this experiment is
about 3 pA and is detected as a downward deflection in the
recording. When more than one channel open, the current and
the downward deflection increase in direct proportion to the
number of open channels. This record shows that up to three
channels are open at any instant.
●
Integral protein
Gate
Aqueous pore
● Figure 2.6 An ion channel. Ion channels are formed
between the polypeptide subunits of integral proteins that span
the plasma membrane, providing an aqueous pore through
which ions can cross the membrane. Different types of gating
mechanisms are used to open and close channels. Ion channels are often selective for a specific ion.
that determines if the channel is open or closed. Specific
stimuli cause a conformational change in the protein subunits to open the gate, creating an aqueous channel through
which the ions can move (Fig. 2.6). In this way, ions do not
have to enter the lipid bilayer to cross the membrane; they are
always in an aqueous medium. When the channels are open,
the ions diffuse rapidly from one side of the membrane to the
other down the concentration gradient. Specific interactions
between the ions and the sides of the channel produce an
extremely rapid rate of ion movement; in fact, ion channels
permit a much faster rate of solute transport (about 108 ions/s)
than the carrier-mediated systems discussed below.
Ion channels have a selectivity filter, which regulates the
transport of certain classes of ions such as anions or cations
or specific ions such as Na+, K+, Ca2+, and Cl− (see Fig 2.6).
The amino acid composition of the channel protein does not
appear to confer the ion selectivity of the channel.
The patch clamp technique has revealed a great deal of
information about the characteristic behavior of channels
for different ions. The small electrical current caused by ion
movement when a channel is open can be detected with this
technique, which is so sensitive that the opening and closing
of a single ion channel can be observed (Fig. 2.7). In general,
ion channels exist either fully open or completely closed, and
they open and close very rapidly. The frequency with which
a channel opens is variable, and the time the channel remains
open (usually a few milliseconds) is also variable. The overall
rate of ion transport across a membrane can be controlled by
changing the frequency of a channel opening or by changing
the time a channel remains open.
Most ion channels usually open in response to a specific stimulus. Ion channels can be classified according to
their gating mechanisms, the signals that make them open
or close. There are voltage-gated channels and ligand-gated
channels. Some ion channels are more like membrane pores
in that they are always open; these ion transport proteins
are referred to as nongated channels (see Chapter 3, “Action
Potential, Synaptic Transmission, and Maintenance of Nerve
Function”).
Voltage-gated ion channels open when the membrane
potential changes beyond a certain threshold value. Channels
of this type are involved in conducting the excitation signal
along nerve axons and include sodium and potassium channels (see Chapter 3, “Action Potential, Synaptic Transmission, and Maintenance of Nerve Function”). Voltage-gated
ion channels are found in many cell types. It is thought that
some charged amino acids located in a membrane-spanning
a-helical segment of the channel protein are sensitive to the
transmembrane potential. Changes in the membrane potential cause these amino acids to move and induce a conformational change of the protein that opens the way for the ions.
Ligand-gated ion channels cannot open unless they
first bind to a specific agonist. The opening of the gate is produced by a conformational change in the protein induced
by the ligand binding. The ligand can be a neurotransmitter arriving from the extracellular medium. It can also be
an intracellular second messenger, produced in response to
some cell activity or hormone that reaches the ion channel
from the inside of the cell. The nicotinic acetylcholine receptor channel found in the postsynaptic neuromuscular junction (see Chapter 3, “Action Potential, Synaptic Transmission,
and Maintenance of Nerve Function” and Chapter 9, “Blood
Components”) is a ligand-gated ion channel that is opened by
an extracellular ligand (acetylcholine). Examples of ion channels gated by intracellular messengers also abound in nature.
This type of gating mechanism allows the channel to open
or close in response to events that occur at other locations
in the cell. For example, a sodium channel gated by intracellular cyclic guanosine monophosphate (cGMP) is involved in
the process of vision (see Chapter 4, “Sensory Physiology”).
This channel is located in the rod cells of the retina and
opens in the presence of cGMP. The generalized structure
of one subunit of an ion channel gated by cyclic nucleotides
is shown in Figure 2.8. There are six membrane-spanning
Dr. Murtadha Al-Shareifi e-Library
●
31
Chapter 2 / Plasma Membrane, Membrane Transport, and Resting Membrane Potential
(Fig. 2.9). During carrier-mediated transport, binding of the
solute to one side of the carrier induces a conformational
change in the protein, which closes one gate and opens the
second gate, allowing the solute to pass through the membrane. As with pores and channels, carriers function until the
solute concentrations have equilibrated.
Carrier-mediated transport systems have several
characteristics:
Out
1
2
3
6
5
4
In
Binding
site
H2N
COOH
A
• They eventually reach saturation at high substrate concentration (see Fig 2.5).
• They have structural specificity, meaning each carrier system recognizes and binds specific chemical structures (a
carrier for d-glucose will not bind or transport l-glucose).
IV
I
III
II
B
● Figure 2.8 Structure of a cyclic nucleotide-gated ion
channel. (A) The secondary structure of a single subunit has six
membrane-spanning regions and a binding site for cyclic nucleotides on the cytosolic side of the membrane. (B) Four identical
subunits (I–IV) assemble together to form a functional channel that
provides a hydrophilic pathway across the plasma membrane.
regions, and a cyclic nucleotide-binding site is exposed to the
cytosol. The functional protein is a tetramer of four identical subunits. Other cell membranes have potassium channels
that open when the intracellular concentration of calcium
ions increases. Several known channels respond to inositol
1,4,5-trisphosphate, the activated part of G proteins, or adenosine triphosphate (ATP). The epithelial chloride channel
that is mutated in cystic fibrosis is normally gated by ATP.
Carrier-mediated transport moves a range
of ions and organic solutes passively
across membranes.
In contrast to pores and ion channels, integral membrane
proteins that form carriers provide a conduit through the
membrane that is never open to both sides of the membrane
at the same time. This is due to the presence of two gates
A
• They allow the transport of polar (hydrophilic) molecules
at rates much higher than that expected from the partition coefficient of these molecules.
B
• They show competitive inhibition by molecules with similar chemical structure. For example, carrier-mediated
transport of d-glucose occurs at a slower rate when molecules of d-galactose are also present. This is because
galactose, structurally similar to glucose, competes with
glucose for the available glucose carrier proteins.
A specific example of carrier-mediated transport is
the movement of glucose from the blood to the interior of
cells. Most mammalian cells use blood glucose as a major
source of cellular energy, and glucose is transported into
cells down its concentration gradient. The transport process in many cells, such as erythrocytes and the cells of fat,
liver, and muscle tissues, involves a plasma membrane protein called GLUT1 (glucose transporter-1). The erythrocyte
GLUT1 has an affinity for d-glucose that is about 2,000fold greater than the affinity for l-glucose. It is an integral
membrane protein that contains 12 membrane-spanning
a-helical segments.
Carrier-mediated transport, like simple diffusion, does
not have a directional preference. It functions equally well
bringing its specific solutes into or out of the cell, depending on the concentration gradient. Net movement by carriermediated transport ceases once the concentrations inside
and outside the cell become equal.
● Figure 2.9 The role of a carrier
protein in facilitated diffusion of solute
molecules across a plasma membrane.
In this example, solute transport into the
cell is driven by the high solute concentration outside compared with inside.
(A) Binding of extracellular solute to the
carrier, a membrane-spanning integral
protein, may trigger a change in protein
conformation that exposes the bound
solute to the interior of the cell. (B) Bound
solute readily dissociates from the carrier
because of the low intracellular concentration of solute. The release of solute may
allow the carrier to revert to its original
conformation (A) to begin the cycle again.
●
Dr. Murtadha Al-Shareifi e-Library
32
Part I / Cellular Physiology
The anion exchange protein (AE1), the predominant
integral protein in the mammalian erythrocyte membrane,
provides a good example of the reversibility of transporter
action. AE1 is folded into at least 12 transmembrane a helices and normally permits the one-for-one exchange of Cl−
and HCO3− ions across the plasma membrane. The direction
of ion movement is dependent only on the concentration
gradients of the transported ions. AE1 has an important
role in transporting CO2 from the tissues to the lungs. The
erythrocytes in systemic capillaries pick up CO2 from tissues
and convert it to HCO3−, which exits the cells via AE1. When
the erythrocytes enter pulmonary capillaries, the AE1 allows
plasma HCO3− to enter erythrocytes, where it is converted
back to CO2 for expiration by the lungs (see Chapter 21,
“Control of Ventilation”).
Active transport systems move solutes
against gradients.
All the passive transport mechanisms tend to bring the cell
into equilibrium with the ECF. Cells must oppose these equilibrating systems and preserve intracellular concentrations of
solutes, in particular ions that are compatible with life.
Primary active transport
Integral membrane proteins that directly use metabolic
energy to transport ions against a gradient of concentration
or electrical potential are known as ion pumps. The direct
use of metabolic energy to carry out transport defines a primary active transport mechanism. The source of metabolic
energy is ATP synthesized by mitochondria, and the different ion pumps hydrolyze ATP to adenosine diphosphate
(ADP) and use the energy stored in the third phosphate
bond to carry out transport. Because of this ability to hydrolyze ATP, ion pumps also are called ATPases.
The most abundant ion pump in higher organisms is the
sodium–potassium pump or Na+/K+-ATPase. It is found in
the plasma membrane of practically every eukaryotic cell and
is responsible for maintaining the low sodium and high potassium concentrations in the cytoplasm by transporting sodium
out of the cell and potassium ions in. The sodium–potassium
pump is an integral membrane protein consisting of two
subunits. The a subunit has 10 transmembrane segments and
is the catalytic subunit that mediates active transport. The
smaller b subunit has one transmembrane segment and is
essential for the proper assembly and membrane targeting of
the pump. The Na+/K+-ATPase is known as a P-type ATPase
because the protein is phosphorylated during the transport
cycle (Fig. 2.10). The pump counterbalances the tendency of
sodium ions to enter the cell passively and the tendency of
potassium ions to leave passively. It maintains a high intracellular potassium concentration, which is necessary for protein
synthesis. It also plays a role in the resting membrane potential by maintaining ion gradients. The sodium–potassium
pump can be inhibited either by metabolic poisons that stop
the synthesis and supply of ATP or by specific pump blockers,
α subunit
β subunit
K+
K+
Out
Lipid bilayer
ATP
In
1
● Figure 2.10 Function of the sodium–potassium
pump. The pump is composed of two large a subunits
that hydrolyze adenosine triphosphate (ATP) and transport
the ions. The two smaller b subunits are molecular chaperones that facilitate the correct integration of the a subunits
into the membrane. In step 1, three intracellular Na+ bind
to the a subunit, and ATP is hydrolyzed. Phosphorylation
(Pi) of the a subunit results in a conformational change,
exposing the Na+ to the extracellular space (step 2). In
step 3, the Na+ diffuses away and two K+ bind, resulting
in dephosphorylation of the a subunit. Dephosphorylation
returns the a subunit to an intracellular conformation. The
K+ diffuses away, and ATP is rebound to start the cycle
over again (step 6). ADP, adenosine diphosphate.
ADP
Na+
Na+
Na+
5
Na +
Na +
Na +
K+
Pi
K+
2
4
Na+
Na+
Na+
K+
K+
3
K+
Pi
Pi
Dr. Murtadha Al-Shareifi e-Library
●
33
Chapter 2 / Plasma Membrane, Membrane Transport, and Resting Membrane Potential
From Bench to Bedside / 2.1
Multidrug Resistance and Cancer
Some studies indicate that up to 40% of human cancers
develop resistance to multiple anticancer agents, and it is a
major problem when treating patients with malignant disease.
A number of cellular mechanisms are known to lead to drug
resistance, but it is now recognized that the most common
mechanism is the efflux of chemotherapeutic drugs from the
tumor cell by adenosine triphosphate–binding cassette (ABC)
family drug transporters. The reduced accumulation of drugs
in tumor cells was originally thought to result from overexpression of a 170-kDa protein termed P-glycoprotein (now
known as MDR1 or ABCB1) in the plasma membrane, resulting in expulsion of cytotoxic drugs, thereby reducing the intracellular concentration to below the threshold for cell killing.
Since the discovery of P-glycoprotein, a number of additional
drug resistance transporters have been identified. These
include the 190-kDa multidrug resistance (MDR)-associated
proteins (MRP1), or ABCC1, and MRP2, or ABCC2, and
the breast cancer–resistance protein (BCRP), or ABCG2.
Unlike the classical mammalian transport systems, which can
be very selective in their substrate preferences, the multidrug
transporters are highly promiscuous and can recognize and
transport a wide range of substrates. P-glycoprotein and
BCRP preferentially exclude large positively charged molecules, whereas the MRP family can exclude both uncharged
molecules and water-soluble anionic compounds. Thus,
the most logical approach to reversing MDR is to find compounds that can block ABC transporter activity.
First-generation modulators of MDR were the calciumchannel blocker, verapamil, and the immunosuppressant,
cyclosporine A. Although these compounds were effective in
such as digoxin, a cardiac glycoside used to treat a variety of
cardiac conditions.
As the Na+/K+-ATPase specifically moves sodium and
potassium ions against their concentration or electrical
potential, a number of other pumps move specific substrates
across membranes utilizing the energy released by ATP
hydrolysis.
• Calcium pumps are P-type ATPases located in the plasma
membrane and the membrane of intracellular organelles.
Plasma membrane Ca2+ ATPases pump calcium out of the
cell. Calcium pumps in the membrane of the endoplasmic
reticulum and in the sarcoplasmic reticulum membrane
within muscle cells (termed SERCAs for sarcoplasmic and
endoplasmic reticulum calcium ATPases) pump calcium
into the lumen of these organelles. The organelles store
calcium and, as a result, help maintain a low cytosolic
concentration of this ion (see Chapter 1, “Homeostasis
and Cellular Signaling”).
• The H+/K+-ATPase is another example of a P-type ATPase
present in the luminal membrane of the parietal cells
in the oxyntic (acid-secreting) glands of the stomach.
By pumping protons into the lumen of the stomach in
exchange for potassium ions, this pump maintains the low
pH in the stomach that is necessary for proper digestion.
cell culture and animal models, side effects due to the high
dose of drug required to reverse chemotherapeutic drug
resistance halted clinical trials. A number of derivatives (structurally similar molecules) for verapamil and cyclosporine A
were subsequently tested. Unfortunately, these compounds
interfered with drug metabolism and elimination, resulting in
overexposure to cytotoxic chemotherapeutic agents. Using
structure–activity relationships and combinatorial chemistry,
a third generation of MDR modulators are being studied that
are effective at nanomolar concentrations and do not affect
the pharmacokinetics of chemotherapeutic agents.
The search for modulators of MDR is also currently
focused on naturally occurring compounds, with the rationale
that such products would be less toxic. Naturally occurring
modulators of MDR include curcumin, coumarin, flavonoids,
chokeberry, and mulberry leaves. Curcumin, in particular,
has been shown to inhibit all three major ABC transporters,
MDR1, MRP, and BCRP. Low oral bioavailability and rapid
metabolism have prompted investigation into improved
methods of delivery such as encapsulation in liposomes.
Finally, the use of transcriptional and translational inhibitors to specifically block synthesis of MDR1, MRP, and
BCRP is under investigation. However, the recent discovery
of single-nucleotide polymorphisms among the ABC drug
transporters may complicate this approach. Polymorphisms
in MDR1 have been shown to alter expression and function
of the protein. Such changes in transporter function could
contribute to variation between different individuals and ethnic groups in both chemotherapeutic drug response and the
effect of inhibitors on transporter activity.
It is also found in the colon and in the collecting ducts of
the kidney. Its role in the kidney is to secrete H+ ions into
the urine, when blood pH falls, and to reabsorb K+ ions
(see Chapter 25, “Neurogastroenterology and Motility”).
• Proton pumps or H+-ATPases are found in the membranes of the lysosomes and the Golgi apparatus. They
pump protons from the cytosol into these organelles,
keeping the inside of the organelles more acidic (at a
lower pH) than the rest of the cell. These pumps, classified as V-type ATPases because they were first discovered in intracellular vacuolar structures, have now been
detected in plasma membranes. For example, the proton
pump in the plasma membrane of specialized bone and
kidney cells is characterized as a V-type ATPase. The
secretion of protons by osteoclasts helps to solubilize the
bone mineral and creates an acidic environment for bone
breakdown by enzymes. The proton pump in the kidney
is present in the same cells as the H+/K+-ATPase and helps
to secrete H+ ions into the urine when blood pH falls.
• ATP-binding cassette (ABC) transporters are a super
family of transporters composed of two transmembrane
domains and two cytosolic nucleotide-binding domains.
The transmembrane domains recognize specific solutes
and transport them across the membrane using a number
●
Dr. Murtadha Al-Shareifi e-Library
34
Part I / Cellular Physiology
of different mechanisms, including conformational
change. The nucleotide-binding domain, or ABC domain,
has a highly conserved sequence. ABC transporters are
involved in a number of cellular processes, including
nutrient uptake, cholesterol and lipid trafficking, resistance to cytotoxic drugs and antibiotics, cellular immune
response, and stem cell biology.
• ABCA1, a member of the ABC subfamily A, has an
important role in effluxing cholesterol, phospholipids,
and other metabolites out of cells. ABCA1 transfers lipids
and cholesterol to lipid-poor high-density lipoproteins
(HDLs). ABCA1 is a unique ABC transporter because it
is also a receptor, binding the lipid-poor HDL to facilitate
the loading of the cholesterol that the transporter is moving out of the cell.
• ABC subfamily C transporters play a crucial role in the
development of multidrug resistance (MDR). There
are a number of different transporters encoded by
multiple MDR genes. The MDR1 transporter is widely
distributed in the liver, brain, lung, kidney, pancreas,
and small intestine and transports a wide range of antibiotics, antivirals, and chemotherapeutic drugs out of
the cell. MDR-associated protein transporters are a
related class of ABCC transporters that also interfere
with antibiotic and chemotherapy. The cystic fibrosis
transmembrane conductance regulator (ABCC7) is
another member of this family.
• Organic anion transporting polypeptides (OATPs)
are members of the solute carrier family and are highly
expressed in the liver, kidney, and brain. OATPs transport anionic and cationic chemicals, steroid, and peptide
backbones generally into cells. Thyroxine, bile acids, and
bilirubin are important solutes transported by OATPs.
These transporters also import agents such 3-hydroxy3-methylglutaryl-CoA reductase inhibitors (statins),
angiotensin-converting enzyme inhibitors, angiotensin
receptor II antagonists, and cardiac glycosides into cells.
• F-type ATPases are located in the inner mitochondrial
membrane. This type of proton pump normally functions in reverse. Instead of using the energy stored in
ATP molecules to pump protons, its principal function
is to synthesize ATP by using the energy stored in a
gradient of protons that is crossing the inner mitochondrial membrane down its concentration gradient. The
proton gradient is generated by the respiratory chain.
Secondary active transport
The net effect of ion pumps is maintenance of the various
environments needed for the proper functioning of organelles, cells, and organs. Metabolic energy is expended by
the pumps to create and maintain the differences in ion
concentrations. Besides the importance of local ion concentrations for cell function, differences in concentrations
represent stored energy. An ion releases potential energy
when it moves down an electrochemical gradient, just as
a body releases energy when falling to a lower level. This
energy can be used to perform work. Cells have developed
Out
In
Out
Na+
Na+
Na+
Na+
Na+
Na+
Na+
In
Na+
Na+
Na+
Na+
Na+
Na+
Na+
Na+
Na+
Na+
Na+
Na+
Na+
Na+
Na+
Na+
Na+
Na+
Na+
1. Na+ binds
2. Solute ( ) binds
3. Translocation
4. Na+ released
5. Solute released
● Figure 2.11 Mechanism of secondary active transport. A solute is moved against its concentration gradient by
coupling it to Na+ moving down a favorable gradient. Binding
of extracellular Na+ to the carrier protein (step 1) may increase
the affinity of binding sites for solute, so that solute also can
bind to the carrier (step 2), even though its extracellular concentration is low. A conformational change in the carrier protein
(step 3) exposes the binding sites to the cytosol, where Na+
readily dissociates because of the low intracellular Na+ concentration (step 4). The release of Na+ decreases the affinity of the
carrier for solute and forces the release of the solute inside the
cell (step 5), where solute concentration is already high. The
free carrier then reverts to the conformation required for step 1,
and the cycle begins again.
several carrier mechanisms to transport one solute against
its concentration gradient by using the energy stored in the
favorable gradient of another solute. In mammals, most of
these mechanisms use sodium as the driver solute and use
the energy of the sodium gradient to carry out the “uphill”
transport of another important solute (Fig. 2.11). Because
the sodium gradient is maintained by the action of the Na+/
K+-ATPase, the function of these transport systems depends
on the function of the Na+/K+-ATPase. Although they do not
directly use metabolic energy for transport, these systems
ultimately depend on the proper supply of metabolic energy
to the sodium–potassium pump. They are called secondary
active transport mechanisms. Disabling the pump with
metabolic inhibitors or pharmacologic blockers causes these
transport systems to stop when the sodium gradient has been
dissipated.
Similar to passive carrier-mediated systems, secondary
active transport systems are integral membrane proteins;
they have specificity for the solute they transport and show
saturation kinetics and competitive inhibition. They differ, however, in two respects. First, they cannot function in
the absence of the driver ion, the ion that moves along its
electrochemical gradient and supplies energy. Second, they
transport the solute against its own concentration or electrochemical gradient. Functionally, the different secondary
active transport systems can be classified into two groups:
symport (cotransport) systems, in which the solute being
transported moves in the same direction as the sodium ion;
and antiport (exchange) systems, in which the sodium ion
and the solute move in opposite directions (Fig. 2.12).
Dr. Murtadha Al-Shareifi e-Library
●
35
Chapter 2 / Plasma Membrane, Membrane Transport, and Resting Membrane Potential
Examples of symport mechanisms are the sodiumcoupled sugar transport system and the several sodiumcoupled amino acid transport systems found in the small
intestine and the renal tubule. The symport systems allow
efficient absorption of nutrients even when the nutrients
are present at low concentrations. The sodium-dependent
glucose transporter-1 (SGLT1) in the human intestine contains 664 amino acids in a single polypeptide chain with
14 membrane-spanning segments (Fig. 2.13). One complete
cycle or turnover of a single SGLT1 protein, illustrated in
Figure 2.11, can occur 1,000 times/s at 37°C. In reality, this
cycle probably involves a coordinated trapping–release cycle
and/or tilt of membrane-spanning segments rather than
the simplistic view presented in Figure 2.11. Another example of a symport system is the family of sodium-coupled
phosphate transporters (termed NaPi, types I and II) in the
intestine and renal proximal tubule. These transporters have
six to eight membrane-spanning segments and contain 460
to 690 amino acids. Sodium-coupled chloride transporters
in the kidney are targets for inhibition by specific diuretics.
The Na+–Cl− cotransporter in the distal tubule, known as
NCC, is inhibited by thiazide diuretics, and the Na+–K+–2Cl−
cotransporter in the ascending limb of the loop of Henle,
referred to as NKCC, is inhibited by bumetanide.
The most important examples of antiporters are the Na+/
+
H exchange and Na+/Ca2+ exchange systems, found mainly
in the plasma membrane of many cells. The first uses the
sodium gradient to remove protons from the cell, controlling
Hexose Malabsorption in the Intestine
Malabsorption of hexoses in the intestine can be the indirect
result of a number of circumstances, such as an increase in
intestinal motility or defects in digestion because of pancreatic insufficiency. Although less common, malabsorption
may be a direct result of a specific defect in hexose transport. Regardless of the cause, the symptoms are common
and include diarrhea, abdominal pain, and gas. The challenge is to identify the cause so proper treatment can be
applied. Some infants develop a copious watery diarrhea
when fed milk that contains glucose or galactose or the
disaccharides lactose and sucrose. The latter are degraded
to glucose, galactose, and fructose by enzymes in the intestine. The dehydration can begin during the first day of life
and can lead to rapid death if not corrected. Fortunately,
the symptoms disappear when a carbohydrate-free formula
fortified with fructose is used instead of milk. This condition
is a rare inherited disease known as glucose–galactose
malabsorption (GGM), and about 200 severe cases have
been reported worldwide. At least 10% of the general population has glucose or lactose intolerance, however, and
it is possible that these people may have milder forms of
the disease. A specific defect in absorption of glucose and
galactose can be demonstrated by tolerance tests in which
oral administration of these monosaccharides produces
little or no increase in plasma glucose or galactose. The primary defect lies in the Na+–glucose cotransporter protein
Out
In
Symport
Na+
S
Na+
S
Na+
S
Na+
S
Antiport
● Figure 2.12 Secondary active transport systems. In
a symport system (top), the transported solute (S) is moved
in the same direction as the Na+ ion. In an antiport system
(bottom), the solute is moved in the opposite direction to Na+.
Boldfaced and lightfaced fonts indicate high and low concentrations, respectively, of Na+ ions and solute.
Clinical Focus / 2.2
(SGLT, Fig. 2.13), located in the apical plasma membrane
of intestinal epithelial cells (Fig. 2.14). Glucose and galactose have very similar structures, and both are substrates
for transport by SGLT. Fructose transport is not affected by
a defect in SGLT because a specific fructose transporter
named GLUT5 is present in the apical membrane. Human
SGLT was cloned in 1989, and almost 30 different mutations have been identified in GGM patients. Many of the
mutations produce premature cessation of SGLT protein
synthesis or disrupt the trafficking of mature SGLT to the
apical plasma membrane. In a few cases, the SGLT reaches
the apical membrane but is no longer capable of glucose
transport. The result in all cases is that functional SGLT proteins are not present in the apical membrane so glucose
and galactose remain in the lumen of the intestine. As these
solutes accumulate in the lumen, the osmolality of the fluids increases and retards absorption of water, leading to
diarrhea and severe water loss from the body. Identification
of specific changes in defective SGLT proteins in patients
has provided clues about the specific amino acids that are
essential for the normal function of SGLT. At the same time,
advances in molecular biology have allowed a better understanding of the genetic defect at the cellular level and how
this leads to the clinical symptoms. GGM is an example of
how information from a disease can further understanding
of physiology and vice versa.
●
Dr. Murtadha Al-Shareifi e-Library
36
Part I / Cellular Physiology
● Figure 2.13 A model of the
secondary structure of the Na+glucose cotransporter protein
(SGLT1) in the human intestine. The
polypeptide chain of 664 amino acids
passes back and forth across the
membrane 14 times. Each membranespanning segment consists of 21 amino
acids arranged in an a-helical conformation. Both the NH2 and the COOH ends
are located on the extracellular side of
the plasma membrane. In the functional protein, the membrane-spanning
segments are clustered together to
provide a hydrophilic pathway across
the plasma membrane. The N-terminal
portion of the protein, including helices
1 to 9, is required to couple Na+ binding
to glucose transport. The five helices
(10–14) at the C terminus may form the
transport pathway for glucose.
NH2
COOH
Out
1
2
3
4
5
6
7
8
9
10
12
11
13
14
In
the intracellular pH and counterbalancing the production of
protons in metabolic reactions. It is an electroneutral system because there is no net movement of charge. One Na+
enters the cell for each H+ that leaves. The second antiporter
removes calcium from the cell and, together with the different calcium pumps, helps maintain a low cytosolic calcium
concentration. It is an electrogenic system because there is
a net movement of charge. Three Na+ enter the cell and one
Ca2+ ion leaves during each cycle.
The structures of the symport and antiport protein
transporters that have been characterized (see Fig. 2.13)
share a common property with ion channels (see Fig. 2.9)
and equilibrating carriers, namely the presence of multiple
membrane-spanning segments within the polypeptide chain.
This supports the concept that, regardless of the mechanism,
the membrane-spanning regions of a transport protein form
a hydrophilic pathway for rapid transport of ions and solutes across the hydrophobic interior of the membrane lipid
bilayer.
An example is the absorption of glucose in the small
intestine. Glucose enters the intestinal epithelial cells by active
transport using the electrogenic Na+–glucose cotransporter
system (SGLT) in the apical membrane. This increases the
Transcellular transport
GLUT2
Epithelial cells occur in layers or sheets that allow the directional movement of solutes not only across the plasma membrane but also from one side of the cell layer to the other. Such
regulated movement is achieved because the plasma membranes of epithelial cells have two distinct regions with different morphologies and different transport systems. These
regions are the apical membrane, facing the lumen, and the
basolateral membrane, facing the blood supply (Fig. 2.14).
The specialized or polarized organization of the cells is maintained by the presence of tight junctions at the areas of contact
between adjacent cells. Tight junctions prevent proteins on
the apical membrane from migrating to the basolateral membrane and those on the basolateral membrane from migrating to the apical membrane. Thus, the entry and exit steps for
solutes can be localized to opposite sides of the cell. This is the
key to transcellular transport across epithelial cells.
Apical (luminal) side
Na+ Glucose
Amino
Na+ acid
Tight junctions
Lumen
SGLT
Cell
layer
Na+
Na
Glucose
+
K+
+
K
Amino
acid
Intercellular
spaces
Blood
Basolateral side
● Figure 2.14 The localization of transport systems to
different regions of the plasma membrane in epithelial cells
of the small intestine. A polarized cell is produced, in which
entry and exit of solutes, such as glucose, amino acids, and
Na+, occur at opposite sides of the cell. Active entry of glucose
and amino acids is restricted to the apical membrane, and exit
requires equilibrating carriers located only in the basolateral
membrane. For example, glucose enters on sodium-dependent
glucose transporter (SGLT) and exits on glucose transporter-2
(GLUT2). Na+ that enters via the apical symporters is pumped
out by the Na+/K+-ATPase on the basolateral membrane. The
result is a net movement of solutes from the luminal side of
the cell to the basolateral side, ensuring efficient absorption of
glucose, amino acids, and Na+ from the intestinal lumen.
Dr. Murtadha Al-Shareifi e-Library
●
37
Chapter 2 / Plasma Membrane, Membrane Transport, and Resting Membrane Potential
intracellular glucose concentration above the blood glucose
concentration, and the glucose molecules move passively out
of the cell and into the blood via an equilibrating carrier mechanism (GLUT2) in the basolateral membrane (see Fig. 2.14).
The intestinal GLUT2, like the erythrocyte GLUT1, is a
sodium-independent transporter that moves glucose down its
concentration gradient. Unlike GLUT1, the GLUT2 transporter can accept other sugars, such as galactose and fructose,
that are also absorbed in the intestine. The Na+/K+-ATPase
that is located in the basolateral membrane pumps out the
sodium ions that enter the cell with the glucose molecules on
SGLT. The polarized organization of the epithelial cells and
the integrated functions of the plasma membrane transporters form the basis by which cells accomplish transcellular
movement of both glucose and sodium ions.
●
WATER MOVEMENT ACROSS
THE PLASMA MEMBRANE
Water can move rapidly in and out of cells, but the partition
coefficient of water into lipids is low; therefore, the permeability of the membrane lipid bilayer for water is also low.
Specific membrane proteins that function as water channels explain the rapid movement of water across the plasma
membrane. These water channels are small (molecular
weight of about 30 kDa), integral membrane proteins known
as aquaporins. Ten different forms have been discovered
so far in mammals. At least six forms are expressed in cells
in the kidney and seven forms in the gastrointestinal tract,
tissues in which water movement across plasma membranes
is particularly rapid.
In the kidney, aquaporin-2 (AQP2) is abundant in
the collecting duct and is the target of the hormone arginine vasopressin, also known as antidiuretic hormone.
This hormone increases water transport in the collecting
duct by stimulating the insertion of AQP2 proteins into
the apical plasma membrane. Several studies have shown
that AQP2 has a critical role in inherited and acquired
disorders of water reabsorption by the kidney. For example, diabetes insipidus is a condition in which the kidney loses its ability to reabsorb water properly, resulting
in excessive loss of water and excretion of a large volume
of very dilute urine (polyuria). Although inherited forms
of diabetes insipidus are relatively rare, it can develop in
patients receiving chronic lithium therapy for psychiatric disorders, giving rise to the term lithium-induced
polyuria. Both of these conditions are associated with a
decrease in the number of AQP2 proteins in the collecting
ducts of the kidney.
Water movement across the plasma
membrane is driven by differences in
osmotic pressure.
The spontaneous movement of water across a membrane
driven by a gradient of water concentration is the process
known as osmosis. The water moves from an area of high concentration of water to an area of low concentration. Because
concentration is defined by the number of particles per unit
of volume, a solution with a high concentration of solutes has
a low concentration of water, and vice versa. Osmosis can,
therefore, be viewed as the movement of water from a solution of high water concentration (low concentration of solute) toward a solution with a lower concentration of water
(high solute concentration). Osmosis is a passive transport
mechanism that tends to equalize the total solute concentrations of the solutions on both sides of every membrane.
If a cell that is in osmotic equilibrium is transferred to a
more dilute solution, water will enter the cell, the cell volume
will increase, and the solute concentration of the cytoplasm
will be reduced. If the cell is transferred to a more concentrated solution, water will leave the cell, the cell volume will
decrease, and the solute concentration of the cytoplasm will
increase. As discussed below, many cells have regulatory
mechanisms that keep cell volume within a certain range.
Other cells such as mammalian erythrocytes do not have
volume regulatory mechanisms, and large volume changes
occur when the solute concentration of the ECF is changed.
The driving force for the movement of water across the
plasma membrane is the difference in water concentration
between the two sides of the membrane. For historical reasons, this driving force is not called the chemical gradient of
water but the difference in osmotic pressure. The osmotic
pressure of a solution is defined as the pressure necessary to
stop the net movement of water across a selectively permeable membrane that separates the solution from pure water.
When a membrane separates two solutions of different
osmotic pressure, water will move from the solution with
low osmotic pressure (high water and low solute concentrations) to the solution of high osmotic pressure (low water
and high solute concentrations). In this context, the term
selectively permeable means that the membrane is permeable
to water but not solutes. In reality, most biologic membranes
contain membrane transport proteins that permit solute
movement.
The osmotic pressure of a solution depends on the number of particles dissolved in it, the total concentration of all
solutes, regardless of the type of solutes present. Many solutes, such as salts, acids, and bases, dissociate in water, so
the number of particles is greater than the molar concentration. For example, NaCl dissociates in water to give Na+ and
Cl−, so one molecule of NaCl will produce two osmotically
active particles. In the case of CaCl2, there are three particles
per molecule. The equation giving the osmotic pressure of a
solution is
p = nRTC
(3)
where p is the osmotic pressure of the solution, n is the number of particles produced by the dissociation of one molecule
of solute (2 for NaCl, 3 for CaCl2), R is the universal gas constant (0.0821 L×atm/mol×K), T is the absolute temperature,
and C is the concentration of the solute in mol/L. Osmotic
pressure can be expressed in atmospheres (atm). Solutions
with the same osmotic pressure are called isosmotic. A solution is hyperosmotic with respect to another solution if it
has a higher osmotic pressure and hyposmotic if it has a
lower osmotic pressure.
●
Dr. Murtadha Al-Shareifi e-Library
38
Part I / Cellular Physiology
Equation 3, called the van’t Hoff equation, is valid only
when applied to very dilute solutions, in which the particles
of solutes are so far away from each other that no interactions occur between them. Generally, this is not the case at
physiologic concentrations. Interactions between dissolved
particles, mainly between ions, cause the solution to behave
as if the concentration of particles is less than the theoretical value (nC). A correction coefficient, called the osmotic
coefficient (j) of the solute, needs to be introduced in the
equation. Therefore, the osmotic pressure of a solution can
be written more accurately as
p = nRTFC
(4)
The osmotic coefficient varies with the specific solute
and its concentration. It has values between 0 and 1. For
example, the osmotic coefficient of NaCl is 1.00 in an infinitely dilute solution but changes to 0.93 at the physiologic
concentration of 0.15 mol/L.
At any given T, because R is constant, equation 4 shows
that the osmotic pressure of a solution is directly proportional to the term nfC. This term is known as the osmolality or osmotic concentration of a solution and is expressed
in Osm/kg H2O. Most physiologic solutions such as blood
plasma contain many different solutes, and each contributes
to the total osmolality of the solution. The osmolality of a
solution containing a complex mixture of solutes is usually
measured by freezing point depression. The freezing point
of an aqueous solution of solutes is lower than that of pure
water and depends on the total number of solute particles.
Compared with pure water, which freezes at 0°C, a solution
with an osmolality of 1 Osm/kg H2O will freeze at −1.86°C.
The ease with which osmolality can be measured has led to
the wide use of this parameter for comparing the osmotic
pressure of different solutions. The osmotic pressures of
physiologic solutions are not trivial. Consider blood plasma,
for example, which usually has an osmolality of 0.28 Osm/kg
H2O, determined by freezing point depression. Equation 4
shows that the osmotic pressure of plasma at 37°C is 7.1 atm,
about seven times greater than the atmospheric pressure.
Many cells can regulate their volume.
Cell volume changes can occur in response to changes in
the osmolality of ECF in both normal and pathophysiologic
situations. Accumulation of solutes also can produce volume
changes by increasing the intracellular osmolality. Many cells
can correct these volume changes.
Volume regulation is particularly important in the brain
where cell swelling can have serious consequences because
expansion is strictly limited by the rigid skull.
Tonicity
A solution’s osmolality is determined by the total concentration of all the solutes present. In contrast, the solution’s
tonicity is determined by the concentrations of only those
solutes that do not enter (“penetrate”) the cell. Tonicity
determines cell volume, as illustrated in the following examples. Na+ behaves as a nonpenetrating solute because it is
pumped out of cells by the Na+/K+-ATPase at the same rate
that it enters. A solution of NaCl at 0.2 Osm/kg H2O is hypoosmotic compared with cell cytosol at 0.3 Osm/kg H2O. The
NaCl solution is also hypotonic because cells will accumulate water and swell when placed in this solution. A solution
containing a mixture of NaCl (0.3 Osm/kg H2O) and urea
(0.1 Osm/kg H2O) has a total osmolality of 0.4 Osm/kg H2O
and will be hyperosmotic compared with cell cytosol. The
solution is isotonic, however, because it produces no permanent change in cell volume. The reason is that cells shrink
initially as a result of loss of water, but urea is a penetrating
solute that rapidly enters the cells. Urea entry increases the
intracellular osmolality, so water also enters and increases
the volume. Entry of water ceases when the urea concentration is the same inside and outside the cells. At this point, the
total osmolality both inside and outside the cells will be 0.4
Osm/kg H2O and the cell volume will be restored to normal.
By extension, it can be seen that normal blood plasma is an
isotonic solution because Na+ is the predominant plasma solute and is nonpenetrating. This stabilizes cell volume while
other plasma solutes (glucose, amino acids, phosphate, urea,
etc.) enter and leave the cells as needed.
Volume regulation mechanisms
When cell volume increases because of extracellular hypotonicity, the response of many cells is rapid activation of
transport mechanisms that tend to decrease the cell volume
(Fig. 2.15A). Different cells use different regulatory volume
decrease (RVD) mechanisms to move solutes out of the cell
and decrease the number of particles in the cytosol, causing water to leave the cell. Because cells have high intracellular concentrations of potassium, many RVD mechanisms
involve an increased efflux of K+, either by stimulating the
opening of potassium channels or by activating symport
mechanisms for KCl. Other cells activate the efflux of some
amino acids, such as taurine or proline. The net result is a
decrease in intracellular solute content and a reduction of
cell volume close to its original value (see Fig. 2.15A).
When placed in a hypertonic solution, cells rapidly lose
water and their volume decreases. In many cells, a decreased
volume triggers regulatory volume increase (RVI) mechanisms, which increase the number of intracellular particles,
bringing water back into the cells. Because Na+ is the main
extracellular ion, many RVI mechanisms involve an influx of
sodium into the cell. Na+–Cl− symport, Na+–K+–2Cl− symport, and Na+/H+ antiport are some of the mechanisms activated to increase the intracellular concentration of Na+ and
increase the cell volume toward its original value (Fig. 2.15B).
Mechanisms based on an increased Na+ influx are effective for only a short time because, eventually, the sodium
pump will increase its activity and reduce intracellular Na+
to its normal value. Cells that regularly encounter hypertonic
ECFs have developed additional mechanisms for maintaining
normal volume. These cells can synthesize specific organic
solutes, enabling them to increase intracellular osmolality
for a long time and avoiding altering the concentrations of
ions they must maintain within a narrow range of values. The
organic solutes are usually small molecules that do not interfere with normal cell function when they accumulate inside
Dr. Murtadha Al-Shareifi e-Library
●
39
Chapter 2 / Plasma Membrane, Membrane Transport, and Resting Membrane Potential
A
Relative volume
2.0
Rapid
swelling
phase
the gastrointestinal tract. The main ingredients of rehydration solutions are glucose, NaCl, and water. The glucose and
Na+ ions are reabsorbed by SGLT1 and other transporters in
the epithelial cells lining the lumen of the small intestine (see
Fig. 2.14). Deposition of these solutes on the basolateral side
of the epithelial cells increases the osmolality in that region
compared with the intestinal lumen and drives the osmotic
absorption of water. Absorption of glucose, and the obligatory increases in absorption of NaCl and water, helps to compensate for excessive diarrheal losses of salt and water.
Regulatory
volume
decrease
1.0
●
0
10
20
The different passive and active transport systems are coordinated in a living cell to maintain intracellular ions and other
solutes at concentrations compatible with life. Consequently,
the cell does not equilibrate with the ECF but rather exists
in a steady state with the extracellular solution. For example, intracellular Na+ concentration (10 mmol/L in a muscle
cell) is much lower than extracellular Na+ concentration (140
mmol/L), so Na+ enters the cell by passive transport through
nongated (always open) Na+ channels. The rate of Na+ entry
is matched, however, by the rate of active transport of Na+ out
of the cell via the sodium–potassium pump (Fig. 2.16). The
net result is that intracellular Na+ is maintained constant and
at a low level, even though Na+ continually enters and leaves
Time (min) in hypotonic solution
B
Relative volume
2.0
Regulatory
volume
increase
1.0
Rapid
shrinkage
0
10
RESTING MEMBRANE
POTENTIAL
20
Time (min) in hypertonic solution
+
Passive exit K
via nongated
channel
Figure 2.15 The effect of tonicity changes on cell
volume. Cell volume changes when a cell is placed in either a
hypotonic or a hypertonic solution. (A) In a hypotonic solution,
the reversal of the initial increase in cell volume is known as a
regulatory volume decrease. Transport systems for solute exit
are activated, and water follows movement of solute out of
the cell. (B) In a hypertonic solution, the reversal of the initial
decrease in cell volume is a regulatory volume increase. Transport systems for solute entry are activated, and water follows
solute into the cell.
Active transport by
+ +
Na /K -ATPase
●
the cell. For example, cells of the medulla of the mammalian
kidney can increase the level of the enzyme aldose reductase when subjected to elevated extracellular osmolality. This
enzyme converts glucose to an osmotically active solute,
sorbitol. Brain cells can synthesize and store inositol. Synthesis of sorbitol and synthesis of inositol represent different
answers to the problem of increasing the total intracellular
osmolality, allowing normal cell volume to be maintained in
the presence of hypertonic ECF.
Oral rehydration therapy is driven by solute
transport.
Oral administration of rehydration solutions has dramatically
reduced the mortality resulting from cholera and other diseases that involve excessive losses of water and solutes from
+ ATP
K
2K+
Na
+
3Na+
ADP
Passive entry
via nongated
channel
Na
+
● Figure 2.16 The concept of a steady state. Na+ enters
a cell through nongated Na+ channels, moving passively
down the electrochemical gradient. The rate of Na+ entry is
matched by the rate of active transport of Na+ out of the cell
via the Na+/K+-ATPase. The intracellular concentration of Na+
remains low and constant. Similarly, the rate of passive K+ exit
through nongated K+ channels is matched by the rate of active
transport of K+ into the cell via the pump. The intracellular K+
concentration remains high and constant. During each cycle of
the ATPase, two K+ are exchanged for three Na+, and one molecule of adenosine triphosphate (ATP) is hydrolyzed. Boldfaced
and lightfaced fonts indicate high and low ion concentrations,
respectively. ADP, adenosine diphosphate.
●
Dr. Murtadha Al-Shareifi e-Library
40
Part I / Cellular Physiology
the cell. The reverse is true for K+, which is maintained at a
high concentration inside the cell relative to the outside. The
passive exit of K+ through nongated K+ channels is matched
by active entry via the pump (see Fig. 2.16). Maintenance of
this steady state with ion concentrations inside the cell different from those outside the cell is the basis for the difference
in electrical potential across the plasma membrane or the
resting membrane potential.
will be Dm = 0. Substituting this condition into equation 5,
we obtain
æC ö
0 = RT ln ç i ÷ + zF(Ei - Eo )
è Co ø
RT æ C i ö
Ei - E o = ln
zF çè C o ÷ø
Ei - E o = +
Ion movement is driven by the
electrochemical potential.
(6)
RT æ C o ö
ln
zF çè C i ÷ø
If there are no differences in temperature or hydrostatic
pressure between the two sides of a plasma membrane,
two forces drive the movement of ions and other solutes
across the membrane. One force results from the difference in the concentration of a substance between the
inside and the outside of the cell and the tendency of
every substance to move from areas of high concentration to areas of low concentration. The other force results
from the difference in electrical potential between the two
sides of the membrane and applies only to ions and other
electrically charged solutes. When a difference in electrical potential exists, positive ions tend to move toward the
negative side, whereas negative ions tend to move toward
the positive side.
The sum of these two driving forces is called the gradient (or difference) of electrochemical potential across the
membrane for a specific solute. It measures the tendency of
that solute to cross the membrane. The expression of this
force is given by
Equation 6, known as the Nernst equation, gives the value
of the electrical potential difference (Ei − Eo) necessary for a
specific ion to be at equilibrium. This value is known as the
Nernst equilibrium potential for that particular ion and it is
expressed in millivolts (mV). At the equilibrium potential, the
tendency of an ion to move in one direction because of the difference in concentrations is exactly balanced by the tendency
to move in the opposite direction because of the difference in
electrical potential. At this point, the ion will be in equilibrium
and there will be no net movement. By converting to log10 and
assuming a physiologic temperature of 37°C and a value of +1
for z (for Na+ or K+), the Nernst equation can be expressed as
Dm = RT ln Ci/C0 + zF(E1-E0)
The resting membrane potential is the electrical potential
difference across the plasma membrane of a normal living
cell in its unstimulated state. It can be measured directly by
the insertion of a microelectrode into the cell with a reference electrode in the ECF. The resting membrane potential is
determined by those ions that can cross the membrane and
are prevented from attaining equilibrium by active transport
systems. Potassium, sodium, and chloride ions can cross the
membranes of every living cell, and each of these ions contributes to the resting membrane potential. By contrast, the
permeability of the membrane of most cells to divalent ions
is so low that it can be ignored in this context.
The Goldman equation gives the value of the membrane potential (in mV) when all the permeable ions are
accounted for
(5)
where m represents the electrochemical potential (Dm is
the difference in electrochemical potential between two
sides of the membrane); Ci and C o are the concentrations
of the solute inside and outside the cell, respectively; Ei
is the electrical potential inside the cell measured with
respect to the electrical potential outside the cell (Eo); R
is the universal gas constant (2 cal/mol×K); T is the absolute temperature (K); z is the valence of the ion; and F is
the Faraday constant (23 cal/mV×mol). By inserting these
units in equation 5 and simplifying, the electrochemical
potential will be expressed in cal/mol, which is the unit
of energy. If the solute is not an ion and has no electrical charge, then z = 0 and the last term of the equation
becomes zero. In this case, the electrochemical potential is defined only by the different concentrations of the
uncharged solute, called the chemical potential. The driving force for solute transport becomes solely the difference
in chemical potential.
Net ion movement is zero at the equilibrium
potential.
Net movement of an ion into or out of a cell continues as
long as the driving force exists. Net movement stops and
equilibrium is reached only when the driving force of
electrochemical potential across the membrane becomes
zero. The condition of equilibrium for any permeable ion
EI – EO = 61 LOG10 (C O /C I )
(7)
Because Na+ and K+ (and other ions) are present at different
concentrations inside and outside a cell, it follows from equation
7 that the equilibrium potential will be different for each ion.
Resting membrane potential is determined
by the passive movement of several ions.
Ei - E o =
+
+
RT æ PK [K ]o + PNa [Na ]o + PCl [Cl ]i ö
ln ç
÷
F è PK [K + ]i + PNa [Na + ]i + PCl [Cl - ]o ø
(8)
where PK, PNa, and PCl represent the permeability of the membrane to potassium, sodium, and chloride ions, respectively,
and brackets indicate the concentration of the ion inside (i)
and outside (o) the cell. If a specific cell is not permeable
to one of these ions, the contribution of the impermeable
ion to the membrane potential will be zero. For a cell that is
permeable to an ion other than the three considered in the
Goldman equation, that ion will contribute to the membrane
potential and must be included in equation 8.
Dr. Murtadha Al-Shareifi e-Library
●
41
Chapter 2 / Plasma Membrane, Membrane Transport, and Resting Membrane Potential
It can be seen from equation 8 that the contribution
of any ion to the membrane potential is determined by the
membrane’s permeability to that particular ion. The higher
the permeability of the membrane to one ion relative to the
others, the more that ion will contribute to the membrane
potential. The plasma membranes of most living cells are
much more permeable to potassium ions than to any other
ion. Making the assumption that PNa and PCl are zero relative
to PK, equation 8 can be simplified to
Ei - E o =
RT æ PK [K + ]o ö
ln
F çè PK [K + ]i ÷ø
(9)
RT æ [K + ]o ö
Ei - E o =
ln
F çè [K + ]i ÷ø
which is the Nernst equation for the equilibrium potential
for K+ (see equation 6). This illustrates two important points:
• In most cells, the resting membrane potential is close to
the equilibrium potential for K+.
• The resting membrane potential of most cells is dominated by K+ because the plasma membrane is more permeable to this ion compared with the others.
As a typical example, the K+ concentrations outside and
inside a muscle cell are 3.5 and 155 mmol/L, respectively.
Substituting these values in equation 7 gives a calculated
equilibrium potential for K+ of −100 mV, negative inside
the cell relative to the outside. Measurement of the resting
membrane potential in a muscle cell yields a value of −90
mV (negative inside). This value is close to, although not the
same as, the equilibrium potential for K+.
The reason the resting membrane potential in the muscle
cell is less negative than the equilibrium potential for K+ is as
follows. Under physiologic conditions, there is passive entry
of Na+ ions. This entry of positively charged ions has a small
but significant effect on the negative potential inside the cell.
Assuming intracellular Na+ to be 10 mmol/L and extracellular Na+ to be 140 mmol/L, the Nernst equation gives a value
of +70 mV for the Na+ equilibrium potential (positive inside
the cell). This is far from the resting membrane potential of
−90 mV. Na+ makes only a small contribution to the resting
membrane potential because membrane permeability to Na+
is low compared with that of K+.
The contribution of Cl− ions need not be considered
because the resting membrane potential in the muscle cell
is the same as the equilibrium potential for Cl−. Therefore,
there is no net movement of chloride ions.
In most cells, as shown above using a muscle cell as an
example, the equilibrium potentials of K+ and Na+ are different from the resting membrane potential, which indicates
that neither K+ ions nor Na+ ions are at equilibrium. Consequently, these ions continue to cross the plasma membrane via
specific nongated channels, and these passive ion movements
are directly responsible for the resting membrane potential.
The Na+/K+-ATPase is important indirectly for maintaining the resting membrane potential because it sets up the
gradients of K+ and Na+ that drive passive K+ exit and Na+
entry. During each cycle of the pump, two K+ ions are moved
into the cell in exchange for three Na+, which are moved out
(see Fig. 2.16). Because of the unequal exchange mechanism,
the pump’s activity contributes slightly (about −5 mV) to the
negative potential inside the cell.
Chapter Summary
• The plasma membrane consists of proteins in a phospholipid bilayer. Integral proteins are embedded in the
bilayer, whereas peripheral proteins are attached to
the outer surface.
• Macromolecules cross the plasma membrane by
endocytosis and exocytosis.
• Passive movement of a solute across a membrane
dissipates the gradient (driving force) and reaches an
equilibrium at which point there is no net movement
of solute.
• Simple diffusion is the passage of lipid-soluble solutes
across the plasma membrane by diffusion through the
lipid bilayer.
• Facilitated diffusion is the passage of water-soluble
solutes and ions through a hydrophilic pathway created by a membrane-spanning integral protein.
• Facilitated diffusion of small ions is mediated by specific pores and ion channel proteins.
• Active transport uses a metabolic energy source to
move solutes against gradients, and the process prevents a state of equilibrium.
• Polarized organization of epithelial cells ensures
directional movement of solutes and water across the
epithelial layer.
• Water crosses plasma membranes rapidly via channel
proteins termed aquaporins. Water movement is a
passive process driven by differences in osmotic
pressure.
• Cells regulate their volume by moving solutes in or out
to drive osmotic entry or exit of water, respectively.
• The driving force for ion transport is the sum of the
electrical and chemical gradients, known as the gradient of electrochemical potential across the membrane.
• The resting membrane potential is determined by the
passive movements of several ions through nongated
channels, which are always open. It is described most
accurately by the Goldman equation, which takes
into account the differences in membrane permeability of different ions. In a muscle cell, for example,
the membrane permeability to Na+ is low compared
with K+ and the resting membrane potential is a result
primarily of passive exit of K+.
Visit http://thePoint.lww.com for chapter review Q&A, case studies, animations, and more!