Добавил:
Опубликованный материал нарушает ваши авторские права? Сообщите нам.
Вуз: Предмет: Файл:

книги студ / color atlas of physiology 5th ed[1]. (a. despopoulos et al, thieme 2003)

.pdf
Скачиваний:
124
Добавлен:
03.09.2020
Размер:
31.56 Mб
Скачать

 

 

!

 

 

 

 

united to form a close electrical and metabolic

 

 

unit (syncytium), as is present in the

 

 

epithelium, many smooth muscles (single-

 

 

unit type, !p. 70), the myocardium, and the

 

 

glia of the central nervous system. Electric

 

 

coupling permits the transfer of excitation,

 

 

e.g., from excited muscle cells to their adjacent

Physiology

 

cells, making it possible to trigger a wave of ex-

 

citation across wide regions of an organ, such

 

as the stomach, intestine, biliary tract, uterus,

 

 

 

 

ureter, atrium, and ventricles of the heart. Cer-

 

 

tain neurons of the retina and CNS also com-

Cell

 

municate in this manner (electric synapses).

 

Gap junctions in the glia

(!p. 338) and

 

 

and

 

epithelia help to distribute the stresses that

 

occur in the course of transport and barrier ac-

Fundamentals

 

tivities (see below) throughout the entire cell

 

 

 

 

community. However, the connexons close

 

 

when the concentration of Ca2+ (in an extreme

 

 

case, due to a hole in cell membrane) or H+

 

 

concentration increases too rapidly ( !C3). In

1

 

other words, the individual (defective) cell is

 

left to deal with its own problems when neces-

 

 

 

 

sary to preserve the functionality of the cell

 

 

community.

 

 

 

Transport through Cell Layers

 

 

 

 

 

 

Multicellular organisms have cell layers that

 

 

are responsible for separating the “interior”

 

 

from the “exterior” of the organism and its

 

 

larger compartments. The epithelia of skin and

 

 

gastrointestinal, urogenital and respiratory

 

 

tracts, the endothelia of blood vessels, and neu-

 

 

roglia are examples of this type of extensive

 

 

barrier. They separate the immediate extra-

 

 

cellular space from other spaces that are

 

 

greatly different in composition, e.g., those

 

 

filled with air (skin, bronchial epithelia),

 

 

gastrointestinal contents, urine or bile

 

 

(tubules,

urinary bladder,

gallbladder),

 

 

aqueous humor of the eye, blood (endothelia)

 

 

and cerebrospinal fluid (blood–cerebrospinal

 

 

fluid barrier), and from the extracellular space

 

 

of the CNS (blood–brain barrier). Nonetheless,

 

 

certain substances must be able to pass

 

 

through these cell layers. This requires selec-

 

 

tive transcellular transport with import into

 

 

the cell followed by export from the cell. Un-

 

 

like cells with a completely uniform plasma

18

 

membrane

(e.g., blood cells), epiand en-

 

dothelial cells are polar cells,

as defined by

 

 

their structure (!p. 9A and B) and transport function. Hence, the apical membrane (facing exterior) of an epithelial cell has a different set of transport proteins from the basolateral membrane (facing the blood). Tight junctions (described below) at which the outer phospholipid layer of the membrane folds over, prevent lateral mixing of the two membranes (!D2).

Whereas the apical and basolateral membranes permit transcellular transport, paracellular transport takes place between cells. Certain epithelia (e.g., in the small intestinal and proximal renal tubules) are relatively permeable to small molecules (leaky), whereas others are less leaky (e.g., distal nephron, colon). The degree of permeability depends on the strength of the tight junctions (zonulae " occludentes) holding the cells together (!D). The paracellular pathway and the extent of its permeability (sometimes cation-specific) are essential functional elements of the various epithelia. Macromolecules can cross the barrier formed by the endothelium of the vessel wall by transcytosis (!p. 28), yet paracellular transport also plays an essential role, especially in the fenestrated endothelium. Anionic macromolecules like albumin, which must remain in the bloodstream because of its colloid osmotic action (!p. 208), are held back by the wall charges at the intercellular spaces and, in some cases, at the fenestra.

Long-distance transport between the various organs of the body and between the body and the outside world is also necessary. Convection is the most important transport mechanism involved in long-distance transport (!p. 24).

Despopoulos, Color Atlas of Physiology © 2003 Thieme

All rights reserved. Usage subject to terms and conditions of license.

C. Gap junction

 

 

 

 

 

 

 

 

II

COO

Cell 1

 

Channel

Cells

 

 

 

 

 

(1.2–1.5 nm)

 

 

 

and Between

NH3+

Cell 2

Cytosol 1

 

 

 

 

Cell membranes

 

 

 

 

 

Cytosol 2

In, Through

 

 

 

 

 

1

 

Ions, ATP, cAMP,

1.9 Transport

 

 

 

amino acids, etc.

2

3

 

 

Connexon of Cell1

 

 

Connexin

 

 

Connexon of Cell 2

 

 

Plate

(27kDa)

 

 

 

 

 

 

Channel open

Channel closed

 

D. Apical functional complex

 

 

 

1

Apical

Microvilli

 

 

 

 

 

Tight

Para-

 

 

junction

cellular

 

See (2)

 

transport

 

 

 

Cell 2

Actin-

 

Zonula

 

myosin

 

Cell 1

belt

 

adherens

 

 

 

 

Epithelial cells (e.g., enterocytes)

Basolateral

 

N

 

 

 

 

 

 

N

 

 

Occlusin

 

 

Myosin II

E-cadherin

 

 

 

 

Actin

 

Adapter proteins

Ca2+

 

 

 

2

 

 

19

 

 

 

Photos: H. Lodish. Reproduced with permission from Scientific American Books, New York, 1995.

Despopoulos, Color Atlas of Physiology © 2003 Thieme

All rights reserved. Usage subject to terms and conditions of license.

Passive Transport by Means of

Diffusion

 

Diffusion is movement of a substance owing to

 

the random thermal motion (brownian move-

 

ment) of its molecules or ions (!A1) in all

 

directions throughout a solvent. Net diffusion

Physiology

or selective transport can occur only when the

solute concentration at the starting point is

 

 

higher than at the target site. (Note: uni-

 

directional fluxes also occur in absence of a

 

concentration gradient—i.e., at equilibrium—

Cell

but net diffusion is zero because there is equal

flux in both directions.) The driving force of

and

diffusion is, therefore, a concentration gra-

 

Fundamentals

dient. Hence, diffusion equalizes concentra-

tion differences and requires a driving force:

 

 

passive transport (= downhill transport).

 

Example: When a layer of O2 gas is placed

 

on water, the O2 quickly diffuses into the water

 

along the initially high gas pressure gradient

1

(!A2). As a result, the partial pressure of O2

(Po2) rises, and O2 can diffuse further

 

 

downward into the next O2-poor layer of water

 

(!A1). (Note: with gases, partial pressure is

 

used in lieu of concentration.) However, the

 

steepness of the Po2 profile or gradient (dPo2/

 

dx) decreases (exponentially) in each sub-

 

sequent layer situated at distance x from the

 

O2 source (!A3). Therefore, diffusion is only

 

feasible for transport across short distances

 

within the body. Diffusion in liquids is slower

 

than in gases.

 

 

 

 

The diffusion rate, Jdiff (mol·s–1), is the

 

amount of substance that diffuses per unit of

 

time. It is proportional to the area available for

 

diffusion (A) and the absolute temperature (T)

 

and is inversely proportional to the viscosity

 

(η) of the solvent and the radius (r) of the dif-

 

fused particles.

 

 

 

 

According to the Stokes–Einstein equation,

 

the coefficient of diffusion (D) is derived from T,

 

η, and r as

 

 

 

 

D !

R ! T

 

[m2 ! s–1],

[1.1]

 

NA · 6π ! r

! η

 

 

 

 

 

where R is

the general

gas constant

 

(8.3144 J·K–1 · mol–1) and NA Avogadro’s con-

 

stant (6.022 · 1023 mol–1). In Fick’s first law of

20

diffusion (Adolf Fick, 1855), the diffusion rate

is expressed as

 

Jdiff ! A ! D ! !

dC

"[mol ! s–1]

[1.2]

dx

where C is the molar concentration and x is the distance traveled during diffusion. Since the driving “force”—i.e., the concentration gradient (dC/dx)—decreases with distance, as was explained above, the time required for diffusion increases exponentially with the distance traveled (t #x2). If, for example, a molecule travels the first µm in 0.5 ms, it will require 5 s to travel 100 µm and a whopping 14 h for 1 cm.

Returning to the previous example (!A2), if the above-water partial pressure of free O2 diffusion (!A2) is kept constant, the Po2 in the water and overlying gas layer will eventually equalize and net diffusion will cease (diffusion equilibrium). This process takes place within the body, for example, when O2 diffuses from the alveoli of the lungs into the bloodstream and when CO2 diffuses in the opposite direction (!p. 120).

Let us imagine two spaces, a and b (!B1) containing different concentrations (Ca "Cb) of an uncharged solute. The membrane separating the solutions has pores x in length and with total cross-sectional area of A. Since the pores are permeable to the molecules of the dissolved substance, the molecules will diffuse from a to b, with Ca –Cb = C representing the concentration gradient. If we consider only the spaces a and b (while ignoring the gradients dC/dx in the pore, as shown in B2, for the sake of simplicity), Fick’s first law of diffusion

(Eq. 1.2) can be modified as follows:

Jdiff ! A ! D !

C

[mol ! s–1].

[1.3]

x

 

 

 

In other words, the rate of diffusion increases as A, D, and C increase, and decreases as the thickness of the membrane ( x) decreases.

When diffusion occurs through the lipid membrane of a cell, one must consider that hydrophilic substances in the membrane are sparingly soluble (compare intramembrane gradient in C1 to C2) and, accordingly, have a hard time penetrating the membrane by means of “simple” diffusion. The oil-and-water partition coefficient (k) is a measure of the lipid solubility of a substance (!C).

!

Despopoulos, Color Atlas of Physiology © 2003 Thieme

All rights reserved. Usage subject to terms and conditions of license.

A. Diffusion in homogeneous media

 

 

 

 

 

1 Brownian particle

 

 

 

 

 

 

I

 

 

Gas

O2

 

 

Diffusion

movement (~T)

 

 

 

 

2

Passive transport

 

3

PO2 profile

 

 

 

 

 

 

 

 

 

PO2

 

 

 

 

of

 

 

 

 

 

 

 

 

O2

O2

O2

 

PO2

 

Means

 

 

 

 

0

 

 

 

 

 

 

 

Slope=gradient

by

 

 

 

 

 

 

 

 

 

 

X

P

=dP/dx

 

 

 

 

Transport

 

 

PO2

 

 

 

 

 

 

 

x

 

 

 

 

 

 

 

 

 

Water

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

P

 

 

 

 

 

 

 

x

 

 

 

 

 

 

0

Distance from O2 source (x)

Passive

B. Diffusion through porous membranes

 

 

 

1

Porous membrane

2

 

1.10

 

 

 

 

 

 

Space a

 

Space b

 

 

Plate

 

 

 

 

Ca

 

Ca

 

 

Cb

Pore

Gradient

Cb

 

 

 

 

 

 

 

x

 

 

Space a

Membrane

Space b

 

Ca–Cb= C

 

 

 

 

 

 

 

 

C. Diffusion through lipid membranes

 

 

 

 

 

 

1

2

 

 

 

5nm

 

Hydrophilic

Hydrophobic

 

 

 

 

 

substance X

substance Y

 

 

CXa

 

 

(k <1)

(k >1)

 

 

 

Gradient

 

CYa

 

Gradient

 

CXb

 

 

 

 

 

 

 

 

 

 

Lipid

 

 

 

Lipid

CYb

Water

Water

 

 

membrane

Water

membrane

Water

 

 

k =

Equilibrium concentration in olive oil

 

 

 

 

Equilibrium concentration in water

 

21

 

 

 

 

 

(Partly after S.G.Schultz)

Despopoulos, Color Atlas of Physiology © 2003 Thieme

All rights reserved. Usage subject to terms and conditions of license.

!

The higher the k value, the more quickly the substance will diffuse through a pure phospholipid bilayer membrane. Substitution into Eq. 1.3 gives

 

Jdiff ! k ! A ! D !

 

C

[mol ! s–1];

[1.4]

 

 

 

 

 

 

 

 

 

 

 

 

 

x

 

 

 

 

 

Whereas the molecular radius r (!Eq. 1.1) still

 

largely determines the magnitude of D when k re-

Physiology

mains constant (cf. diethylmalonamide with ethyl-

urea in D), k can vary by many powers of ten when r

 

 

remains constant (cf. urea with ethanol in D) and can

 

therefore have a decisive effect on the permeability

 

of the membrane.

 

 

 

 

 

 

 

Cell

 

Since the value of the variables k, D, and

x

within the body generally cannot be deter-

and

mined, they are usually summarized as the

 

Fundamentals

permeability coefficient P, where

 

 

P ! k !

D

[m ! s–1].

 

 

[1.5]

 

 

 

 

 

 

 

 

 

 

x

 

 

 

 

 

 

 

 

 

If the diffusion rate, Jdiff [mol!s–1], is related to

 

area A, Eq. 1.4 is transformed to yield

 

 

1

 

Jdiff

 

! P !

C [mol ! m–2 ! s–1].

[1.6]

 

 

 

 

 

A

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

The quantity of substance (net) diffused per

 

unit area and time is therefore proportional to

 

C and P (!E, blue line with slope P).

 

 

 

When considering the diffusion of gases,

C

 

in Eq. 1.4 is replaced by α·

P (solubility coeffi-

 

cient

 

times

partial

pressure. difference;

 

!p. 126) and Jdiff [mol !

s–1] by Vdiff

[m3! s–1].

 

k · α · D is then summarized as diffusion con-

 

ductance, or Krogh’s diffusion coefficient K [m2 !

 

s–1 ! Pa–1]. Substitution into Fick’s first diffusion

 

equation yields

 

 

 

 

 

 

 

 

.

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

Vdiff

 

! K !

 

P

 

[m ! s–1].

[1.7]

 

 

A

 

 

x

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

Since

A and

 

x of

alveolar gas exchange

 

(!p. 120) cannot be determined in living or-

 

ganisms, K · F/

x for O2 is often expressed as

 

the O2

diffusion capacity of the lung, DL:

 

 

.

 

 

 

 

 

PO2 [m3 ! s–1].

 

 

 

 

VO2diff ! DL !

[1.8]

 

Nonionic diffusion occurs when the uncharged

 

form of a weak base (e.g., ammonia = NH3) or

 

acid (e.g., formic acid, HCOOH) passes through

 

a membrane more readily than the charged

22

form (!F). In this case, the membrane would

be more permeable

to

NH3 than

to NH4+

(!p. 176 ff.). Since the pH of a solution determines whether these substances will be charged or not (pK value; !p. 378), the diffusion of weak acids and bases is clearly dependent on the pH.

The previous equations have not made allowances for the diffusion of electrically charged particles (ions). In their case, the electrical potential difference at cell membranes must also be taken into account. The electrical potential difference can be a driving force of diffusion (electrodiffusion). In that case, positively charged ions (cations) will then migrate to the negatively charged side of the membrane, and negatively charged ions (anions) will migrate to the positively charged side. The prerequisite for this type of transport is, of course, that the membrane contain ion channels (!p. 32 ff.) that make it permeable to the transported ions. Inversely, every ion diffusing along a concentration gradient carries a charge and thus creates an electric diffusion potential (!p. 32 ff.).

As a result of the electrical charge of an ion, the permeability coefficient of the ion x (= Px) can be transformed into the electrical conductance of the membrane for this ion, gx (!p. 32):

gx ! ! Px ! zx2 ! F2 R–1 ! T–1 !

 

x [S ! m–2]

[1.9]

c

where R and T have their usual meaning (explained above) and zx equals the charge of the ion, F equals the Faraday constant (9,65 ! 104 A ! s ! mol–1), and cx equals the mean ionic activity in the membrane. Furthermore,

c !

c1

–c2 .

[1.10]

lnc1

 

– lnc2

 

where index 1 = one side and index 2 = the other side of the membrane. Unlike P, g is concentration-depend- ent. If, for example, the extracellular K+ concentration rises from 4 to 8 mmol/kg H2O (cytosolic concentration remains constant at 160 mmol/kg H2O), c will rise, and g will increase by 20%.

Since most of the biologically important substances are so polar or lipophobic (small kvalue) that simple diffusion of the substances through the membrane would proceed much too slowly, other membrane transport proteins called carriers or transporters exist in addition to ion channels. Carriers bind the target molecule (e.g., glucose) on one side of the membrane and detach from it on the other side

Despopoulos, Color Atlas of Physiology © 2003 Thieme

All rights reserved. Usage subject to terms and conditions of license.

D. Permeability of lipid membranes

 

E. Facilitated diffusion

 

310–5

 

Methanol

Triethyl

 

 

 

 

II

 

 

 

 

 

 

citrate

 

 

Facilitated diffusion

Transport by Means of Diffusion

)

 

 

 

 

Ethanol

Trimethyl citrate

 

 

 

 

 

 

(see G. for carriers)

–1

 

 

 

 

 

 

 

 

 

 

 

Antipyrine

 

 

 

 

 

 

 

 

 

]

 

 

 

 

 

 

 

 

Valeramide

 

 

 

 

 

 

Cyanamide

 

–1

 

n

 

 

–6

 

 

s

 

 

 

 

 

 

Diacetin

 

ti

 

3 10

 

 

 

 

 

 

–2

a

 

 

 

 

 

 

Butyramide

 

r

 

 

 

 

Acetamide

 

 

 

u

 

 

 

 

 

 

 

 

 

t

 

 

 

 

 

Chlorohydrin

 

 

a

 

 

 

 

Ethylene

 

 

 

S

 

 

 

 

 

Succinamide

 

 

 

 

 

 

 

glycol

 

Dimethylurea

 

 

 

 

310–7

Methyl-

 

Ethylurea

 

 

 

 

 

 

 

 

urea

 

Diethylmalonamide

 

 

 

 

 

 

 

Urea

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

310–8

 

 

 

(Sphere diameter

 

Simple diffusion

 

 

 

 

 

 

coefficientPermeability(m•s

 

 

Glycerol

 

= molecular radius)

Transport[molratem•

 

 

 

 

 

 

 

Collander(Datafromal.)et

 

 

 

 

10–4

10–3

10–2

10–1

 

C[molm–3]

 

 

 

1

 

 

Passive

 

 

 

Distribution coefficient k for olive oil/water

 

 

 

F. Nonionic diffusion

 

G. Passive carrier transport

 

 

 

 

 

 

 

 

 

Carrier

 

1.11

H+

+ NH4+

 

NH4+ +

H+

 

 

protein

 

 

 

 

 

 

 

 

 

 

Plate

 

 

 

 

 

 

 

 

 

 

 

 

 

 

NH3

 

NH3

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

H+ + HCOO

 

HCOO+

H+

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

HCOOH

 

 

 

 

 

 

 

 

 

 

 

 

 

HCOOH

 

 

 

 

 

 

(after a conformational change) (!G). As in simple diffusion, a concentration gradient is necessary for such carrier-mediated transport (passive transport), e.g., with GLUT uniporters for glucose (!p. 158). On the other hand, this type of “facilitated diffusion” is subject to satu-

ration and is specific for structurally similar substances that may competitively inhibit one another. The carriers in both passive and active transport have the latter features in common (!p. 26).

23

Despopoulos, Color Atlas of Physiology © 2003 Thieme

All rights reserved. Usage subject to terms and conditions of license.

1 Fundamentals and Cell Physiology

24

Osmosis, Filtration and Convection

Water flow or volume flow (JV) across a membrane, in living organisms is achieved through osmosis (diffusion of water) or filtration. They can occur only if the membrane is water-per- meable. This allows osmotic and hydrostatic pressure differences (Δπ and P) across the membrane to drive the fluids through it.

Osmotic flow equals the hydraulic conductivity (Kf) times the osmotic pressure difference (Δπ) (!A):

JV ! Kf ! Δπ

[1.11]

The osmotic pressure difference (Δπ) can be calculated using van’t Hoff’s law, as modified by Staverman:

Δπ ! σ ! R ! T ! Cosm,

[1.12]

where σ is the reflection coefficient of the particles (see below), R is the universal gas constant (! p. 20), T is the absolute temperature,

and Cosm [osm ! kgH2O–1] is the difference between the lower and higher particle concen-

trations, Caosm –Cbosm (!A). Since Cosm, the driving force for osmosis, is a negative value, JV

is also negative (Eq. 1.11). The water therefore flows against the concentration gradient of the solute particles. In other words, the higher concentration, Cbosm, attracts the water. When the concentration of water is considered in osmosis, the H2O concentration in A,a, CaH2O, is greater than that in A,b, CbH2O. CaH2O –CbH2O is therefore the driving force for H2O diffusion

(!A). Osmosis also cannot occur unless the reflection coefficient is greater than zero (σ "0), that is, unless the membrane is less permeable to the solutes than to water.

Aquaporins (AQP) are water channels that permit the passage of water in many cell membranes. A chief cell in the renal collecting duct contains a total of ca. 107 water channels, comprising AQP2 (regulated) in the luminal membrane, and AQP3 and 4 (permanent?) in the basolateral membrane. The permeability of the epithelium of the renal collecting duct to water (!A, right panel) is controlled by the insertion and removal of AQP2, which is stored in the membrane of intracellular vesicles. In the presence of the antidiuretic hormone ADH (V2 receptors, cAMP; ! p. 274), water channels are inserted in the luminal membrane within minutes, thereby increasing the water permeability of the membrane to around 1.5 #10– 17 L s– 1 per channel.

In filtration (!B),

 

JV ! Kf ! P – Δπ

[1.13]

Filtration occurs through capillary walls, which allow the passage of small ions and molecules (σ = 0; see below), but not of plasma proteins (!B, molecule x). Their concentration difference leads to an oncotic pressure difference (Δπ) that opposes P. Therefore, filtration can occur only if P "Δπ (!B, p. 152, p. 208).

Solvent drag occurs when solute particles are carried along with the water flow of osmosis or filtration. The amount of solvent drag for solute X (JX) depends mainly on osmotic flow (JV) and the mean solute activity ax (! p. 376) at the site of penetration, but also on the degree of particle reflection from the membrane, which is described using the reflection coefficient (σ). Solvent drag for solute X (JX) is therefore calculated as

Jx ! JV (1 – σ) ax [mol ! s–1] [1.14] Larger molecules such as proteins are entirely

reflected, and σ = 1 (!B, molecule X). Reflection of smaller molecules is lower, and σ$ 1. When urea passes through the wall of the proximal renal tubule, for example, σ = 0.68. The value (1–σ) is also called the sieving coefficient (! p. 154).

Plasma protein binding occurs when smallmolecular substances in plasma bind to proteins (!C). This hinders the free penetration of the substances through the endothelium or the glomerular filter (! p. 154 ff.). At a glomerular filtration fraction of 20%, 20% of a freely filterable substance is filtered out. If, however, 9/10 of the substance is bound to plasma proteins, only 2% will be filtered during each renal pass.

Convection functions to transport solutes over long distances—e.g., in the circulation or urinary tract. The solute is then carried along like a piece of driftwood. The quantity of solute

transported over time (Jconv) is the product of volume flow JV (in m3 ! s–1) and the solute con-

centration C (mol ! m–3):

Jconv ! JV ! C [mol ! s–1]. [1.15] The flow of gases in the respiratory tract, the transmission of heat in the blood and the release of heat in the form of warmed air occurs through convection (! p. 222).

Despopoulos, Color Atlas of Physiology © 2003 Thieme

All rights reserved. Usage subject to terms and conditions of license.

A. Osmosis (water diffusion)

 

 

 

 

 

 

Caosm

π

Cbosm

 

b

Example

Lumen

Interstice

 

C

CaH2O

 

 

H2O

 

 

 

 

 

 

 

H2O

 

 

 

 

 

b

a

Aqua-

 

 

 

 

 

C osm > C osm,

porins

 

 

 

 

 

 

i.e.,

 

 

 

 

 

 

 

 

H2O

 

 

 

CaH2O > CbH2O

 

 

a

 

b

 

Water diffusion

 

 

 

 

 

 

from a to b

 

 

 

 

 

 

 

 

 

Epithelium

 

 

 

 

 

 

 

of renal

 

Water flux JV = Kf ·

π (~ Caosm– Cbosm)

collecting duct

 

 

 

B. Filtration

 

 

 

P

 

 

 

 

 

Pa

Example

 

 

 

 

 

 

 

 

 

 

 

Pb

 

 

 

Glomerular

 

 

 

 

 

 

 

 

 

 

 

 

 

capillary

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

Pa > Pb

 

 

 

 

 

 

 

 

 

 

 

 

and

πx

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

a

x

 

 

 

 

P >

 

Blood

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

b

 

 

 

 

 

 

 

 

 

 

 

 

 

 

Water filtration

 

 

 

 

 

 

 

 

 

 

 

 

from a to b

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

πx

 

 

πx)

 

 

P

 

 

π

 

Water flux JV = Kf · ( P

 

Primary

Filtrate

 

 

 

 

(= oncotic pressure

 

 

 

 

 

 

 

 

 

 

urine

 

 

of plasma proteins)

 

 

 

 

 

 

 

 

 

 

 

 

 

 

C. Plasma protein binding

Protein

Blood

side

 

a

H2O

Prevents excretion

(e.g., binding of heme by hemopexin)

Transports substances in blood

(e.g., binding of Fe3+ ions by apotransferrin)

Provides rapid access ion stores (e.g., of Ca2+ or Mg2+)

Helps to dissolve lipophilic substances in blood (e.g., unconjugated bilirubin)

bAffects certain medications (e.g., many sulfonamides): Protein-bound fraction

not pharmacologically active

not filtratable (delays renal excretion)

functions as an allergen (hapten)

Plate 1.12 Osmosis, Filtration and Convection

25

Despopoulos, Color Atlas of Physiology © 2003 Thieme

All rights reserved. Usage subject to terms and conditions of license.

1 Fundamentals and Cell Physiology

26

Active Transport

and, thus, for maintenance of the cell mem-

brane potential. During each transport cycle

 

Active transport occurs in many parts of the

(!A1, A2), 3 Na+ and 2 K+ are “pumped” out of

body when solutes are transported against

and into the cell, respectively, while 1 ATP

their concentration gradient (uphill transport)

molecule is used to phosphorylate the carrier

and/or, in the case of ions, against an electrical

protein (!A2b). Phosphorylation first

potential (! p. 22). All in all, active transport

changes the conformation of the protein and

occurs against the electrochemical gradient or

subsequently alters the affinities of the Na+

potential of the solute. Since passive transport

and K+ binding sites. The conformational

mechanisms represent “downhill” transport

change is the actual ion transport step since it

(! p. 20 ff.), they are not appropriate for this

moves the binding sites to the opposite side of

task. Active transport requires the expenditure

the membrane (!A2b – d). Dephosphoryla-

of energy. A large portion of chemical energy

tion restores the pump to its original state

provided by foodstuffs is utilized for active

(!A2e – f). The Na+/K+ pumping rate increases

transport once it has been made readily avail-

when the cytosolic Na+ concentration rises—

able in the form of ATP (! p. 41). The energy

due, for instance, to increased Na+ influx, or

created by ATP hydrolysis is used to drive the

when the extracellular K+ rises. Therefore,

transmembrane transport of numerous ions,

Na+,K+-activatable ATPase is the full name of

metabolites, and waste products. According to

the pump. Na-+K+-ATPase

is inhibited by

the laws of thermodynamics, the energy ex-

ouabain and cardiac glycosides.

 

pended in these reactions produces order in

Secondary active transport occurs when

cells and organelles—a prerequisite for sur-

uphill transport of a compound (e.g., glucose)

vival and normal function of cells and, there-

via a carrier protein (e.g., sodium glucose

fore, for the whole organism (! p. 38 ff.).

transporter type 2, SGLT2) is coupled with the

In primary active transport, the energy pro-

passive (downhill) transport of an ion (in this

duced by hydrolysis of ATP goes directly into

example Na+; !B1). In this case, the electro-

ion transport through an ion pump. This type

chemical Na+ gradient into the cell (created by

of ion pump is called an ATPase. They establish

Na+-K+-ATPase at another site on the cell mem-

the electrochemical gradients rather slowly,

brane; !A) provides the driving force needed

e.g., at a rate of around 1 µmol ! s–1 ! m–2 of

for secondary active uptake of glucose into the

membrane surface area in the case of Na+-K+-

cell. Coupling of the transport of two com-

ATPase. The gradient can be exploited to

pounds across a membrane is called cotrans-

achieve rapid ionic currents in the opposite

port, which may be in the form of symport or

direction after the permeability of ion chan-

antiport. Symport occurs when the two com-

nels has been increased (! p. 32 ff.). Na+ can,

pounds (i.e., compound and driving ion) are

for example, be driven into a nerve cell at a rate

transported across the membrane in the same

of up to 1000 µmol ! s–1 ! m–2 during an action

direction (!B1–3). Antiport (countertrans-

potential.

port) occurs when they are transported in op-

ATPases occur ubiquitously in cell mem-

posite directions. Antiport occurs, for example,

branes (Na+-K+-ATPase) and in the endo-

when an electrochemical Na+ gradient drives

plasmic reticulum and plasma membrane

H+ in the opposite direction by secondary ac-

(Ca2+-ATPase), renal collecting duct and stom-

tive transport (!B4). The resulting H+ gradient

ach glands (H+,K+ -ATPase), and in lysosomes

can then be exploited for tertiary active sym-

(H+-ATPase). They transport Na+, K+, Ca2+ and

port of molecules such as peptides (!B5).

H+, respectively, by primarily active mecha-

Electroneutral transport occurs when the

nisms. All except H+-ATPase consist of 2 α-sub-

net electrical charge remains balanced during

units and 2 "-subunits (P-type ATPases). The

transport, e.g., during Na+/H+ antiport (!B4)

α-subunits are phosphorylated and form the

and Na+-Cl symport (!B2). Small charge sep-

ion transport channel (!A1).

aration occurs in electrogenic (rheogenic)

Na+-K+-ATPase is responsible for main-

transport, e.g., in

Na+-glucose0

symport

tenance of intracellular Na+ and K+ homeostasis

(!B1), Na+-amino

acid0

symport

(!B3),

!

Despopoulos, Color Atlas of Physiology © 2003 Thieme

All rights reserved. Usage subject to terms and conditions of license.

+

+

 

 

 

 

A. Na , K - ATPase

 

 

 

 

 

 

3 Na+

 

 

 

 

1

Outside

[Na+]o

 

[K+]o

 

 

 

 

 

 

 

β

β

 

 

 

 

 

 

 

 

 

membrane

α

α

 

 

 

 

Cytosol Cell

[Na+]i

ADP

[K+]i

Active Transport I

 

 

 

ATP

 

 

 

 

2 K+

 

2

 

a

 

b

 

1.13

 

 

 

 

Plate

 

 

 

 

 

 

 

 

 

 

 

P

 

 

 

 

 

ATP

 

 

 

 

 

Na+ binding,

Phosphorylation

Confor-

 

 

 

 

K+ discharge

 

High affinity

 

 

 

 

 

mational

 

 

 

 

 

change

for K+

 

 

 

 

 

c

 

f

 

 

Conformation E1

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

Na+

 

 

K+

 

 

 

P

 

 

 

 

 

 

hochaffin

Conformation E2

 

 

 

d

 

Na+

 

 

 

 

Conformational

 

e

 

 

change

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

P

 

 

 

 

 

Na+ discharge,

 

 

 

 

 

Pi

K+ binding

27

 

 

 

Dephosphorylation

 

 

 

 

(Partly after P. Läuger)

Despopoulos, Color Atlas of Physiology © 2003 Thieme

All rights reserved. Usage subject to terms and conditions of license.

Соседние файлы в папке книги студ