13.09.2022 Views

Molecular Biology of the Cell by Bruce Alberts, Alexander Johnson, Julian Lewis, David Morgan, Martin Raff, Keith Roberts, Peter Walter by by Bruce Alberts, Alexander Johnson, Julian Lewis, David Morg

Create successful ePaper yourself

Turn your PDF publications into a flip-book with our unique Google optimized e-Paper software.

TRANSPORTERS AND ACTIVE MEMBRANE TRANSPORT

601

lipid

bilayer

solute

OUTWARD-

OPEN

OCCLUDED

INWARD-

OPEN

OUTSIDE

INSIDE

concentration

gradient

conformational changes that alternately expose the solute-binding site first on

one side of the membrane and then on the other—but never on both sides at the

same time. The transition occurs through an intermediate state in which the solute

is inaccessible, or occluded, from either side of the membrane (Figure 11–5).

MBoC6 m11.05/11.05

When the transporter is saturated (that is, when all solute-binding sites are occupied),

the rate of transport is maximal. This rate, referred to as V max (V for velocity),

is characteristic of the specific carrier. V max measures the rate at which the

carrier can flip between its conformational states. In addition, each transporter

has a characteristic affinity for its solute, reflected in the K m of the reaction, which

is equal to the concentration of solute when the transport rate is half its maximum

value (Figure 11–6). As with enzymes, the binding of solute can be blocked by

either competitive inhibitors (which compete for the same binding site and may

or may not be transported) or noncompetitive inhibitors (which bind elsewhere

and alter the structure of the transporter).

As we discuss shortly, it requires only a relatively minor modification of the

model shown in Figure 11–5 to link a transporter to a source of energy in order

to pump a solute uphill against its electrochemical gradient. Cells carry out such

active transport in three main ways (Figure 11–7):

1. Coupled transporters harness the energy stored in concentration gradients

to couple the uphill transport of one solute across the membrane to the

downhill transport of another.

2. ATP-driven pumps couple uphill transport to the hydrolysis of ATP.

3. Light- or redox-driven pumps, which are known in bacteria, archaea, mitochondria,

and chloroplasts, couple uphill transport to an input of energy

from light, as with bacteriorhodopsin (discussed in Chapter 10), or from a

redox reaction, as with cytochrome c oxidase (discussed in Chapter 14).

Amino acid sequence and three-dimensional structure comparisons suggest

that, in many cases, there are strong similarities in structure between transporters

that mediate active transport and those that mediate passive transport. Some

bacterial transporters, for example, that use the energy stored in the H + gradient

across the plasma membrane to drive the active uptake of various sugars are

structurally similar to the transporters that mediate passive glucose transport

into most animal cells. This suggests an evolutionary relationship between various

transporters. Given the importance of small metabolites and sugars as energy

sources, it is not surprising that the superfamily of transporters is an ancient one.

We begin our discussion of active membrane transport by considering a class

of coupled transporters that are driven by ion concentration gradients. These proteins

have a crucial role in the transport of small metabolites across membranes

in all cells. We then discuss ATP-driven pumps, including the Na + -K + pump that is

found in the plasma membrane of most animal cells. Examples of the third class

of active transport—light- or redox-driven pumps—are discussed in Chapter 14.

Active Transport Can Be Driven by Ion-Concentration Gradients

Some transporters simply passively mediate the movement of a single solute from

one side of the membrane to the other at a rate determined by their V max and

rate of transport

Figure 11–5 A model of how a

conformational change in a transporter

mediates the passive movement of a

solute. The transporter is shown in three

conformational states: in the outwardopen

state, the binding sites for solute are

exposed on the outside; in the occluded

state, the same sites are not accessible

from either side; and in the inward-open

state, the sites are exposed on the inside.

The transitions between the states occur

randomly. They are completely reversible

and do not depend on whether the solutebinding

site is occupied. Therefore, if

the solute concentration is higher on the

outside of the bilayer, more solute binds

to the transporter in the outward-open

conformation than in the inward-open

conformation, and there is a net transport

of solute down its concentration gradient

(or, if the solute is an ion, down its

electrochemical gradient).

V max

transporter-mediated

diffusion

1/2V max

K m

simple diffusion

and channel-mediated

transport

concentration of

transported molecule

Figure 11–6 The kinetics of simple

diffusion compared with transportermediated

diffusion. Whereas the rate

of diffusion and channel-mediated

transport is directly proportional to the

solute concentration (within the physical

limits imposed by total surface area

or total channels available), the rate of

transporter-mediated MBoC6 m11.06/11.06

diffusion reaches a

maximum (V max ) when the transporter is

saturated. The solute concentration when

the transport rate is at half its maximal

value approximates the binding constant

(K m ) of the transporter for the solute and

is analogous to the K m of an enzyme

for its substrate. The graph applies to a

transporter moving a single solute; the

kinetics of coupled transport of two or

more solutes is more complex and exhibits

cooperative behavior.

Hooray! Your file is uploaded and ready to be published.

Saved successfully!

Ooh no, something went wrong!