Sie sind auf Seite 1von 16

Biological Membrane

MARCH 13, 2017


Sameer Jareen
UOM Semester IV
Introduction
A membrane is a selective barrier; it allows some things to pass through but stops others. Such things may
be molecules, ions, or other small particles. Membranes control the flow of information between cells
and are involved in the capture and release of energy; photosynthesis and oxidative phosphorylation take
place on membranes.

Therefore Biological membranes are more than just an inert barrier or covering, they also play an active
part in the life of the cell.

Biological membranes are made of three important components: lipids, proteins and sugars. All
membranes have a common basic structure, in which two-layered sheets of lipid molecules have proteins
embedded between them. The structure is highly fluid and most of the lipid and protein molecules can
move about within the plane of the membrane. The lipid and protein molecules are held together
primarily by means of non-covalent interactions. Sugars are attached by covalent bonds to some of the
lipid and protein molecules. They are found on one side of the membrane only i.e.: outer surface of the
plasma membrane.

1. Components of Biological Membrane

I. Lipids
Lipids are biologically essential components that are insoluble in water but soluble in organic
solvents such as propanone (acetone), ethanol, chloroform, diethyl ether and light petroleum
(B.P. 4060 C). There are three predominant types of lipid found in biological membranes:
phospholipids, glycolipids and sterol. They each play unique roles within the membrane.

a) Phospholipids
The most common type of phospholipid consists of glycerol attached to two fatty acid chains,
phosphate and choline. The fatty acid chains generally contain between 14 and 24 carbon
atoms. One chain is usually unsaturated, containing from one to four Cis-double bonds. Each
double bond puts a bend in the fatty acid chain. Because they contain glycerol, lipids of this
type are called glycerophospholipids.

The three major glycerophospholipids contain choline, or serine, or ethanolamine attached


to the phosphate.

Serine - HOCH2CH(COO)NH3+.
Ethanolamine - HOCH2CH2NH3+.

The most common example, sphingomyelin, contains choline attached to the phosphate.
Fig 1.1 (a) Phosphatidyl choline, a glycerophospholipid; (b) sphingomyelin, a sphingo-phospholipid

b) Glycolipids

In common with phospholipids, glycolipid molecules contain either glycerol or sphingosine linked
to fatty acid chains. They differ from phospholipids in that glycolipids have a sugar, such as glucose
or galactose, instead of the phosphate-containing head. Glycolipids in animal membranes almost
always contain sphingosine, whereas those in bacterial and plant membranes contain glycerol. In
all cases, glycolipids are found on the outer surface of the plasma membrane with their sugars
exposed at the cell surface.

Fig 1.2 Glycolipid structures.


c) Sterols

The third type of membrane lipid is cholesterol, a molecule that is structurally quite different from
the phospholipids and glycolipids. Cholesterol contains a four-ring steroid structure together with
a short hydrocarbon side-chain and a hydroxyl group. Cholesterol is found in some mammalian
membranes, and also in one type of micro-organism the mycoplasmas. It is not usually found
in most bacterial membranes, nor in any plant membranes.

Fig 1.3. - Cholesterol


Amphipathic

All membrane lipids are amphipathicthat is, they contain both a hydrophilic region and a hydrophobic
region. Thus the most favorable environment for the hydrophilic head is an aqueous one, whereas the
hydrophobic tail is more stable in a lipid environment. The amphipathic nature of membrane lipids means
that they naturally form bilayers in which the hydrophilic heads point outward towards the aqueous
environment and the hydrophobic tails point inward towards each other (Figure 1.4a). When placed in
water, membrane lipids will spontaneously form liposomes, which are spheres formed of a bilayer with
water inside and outside, resembling a tiny cell (Figure 1.4b). This is the most favorable configuration for
these lipids, as it means that all of the hydrophilic heads are in contact with water and all of the
hydrophobic tails are in a lipid environment.

Fig 1.4. (a) A membrane bilayer (b) Section through a liposome


II. Membrane Protein

In addition to the lipid bilayer, the cell membrane also contains a number of proteins. While
the lipid bilayer provides the structure for the cell membrane, membrane proteins allow for
many of the interactions that occur between cells. As membrane proteins are free to move
within the lipid bilayer as a result of its fluidity Membrane proteins perform various functions,
and this diversity is reflected in the significantly different types of proteins associated with
the lipid bilayer.
Proteins are generally broken down into the smaller classifications of integral proteins,
peripheral proteins, and transmembrane proteins.

a) Integral Protein
An integral membrane protein (IMP) is a type of membrane protein that is permanently
attached to the biological membrane. All transmembrane proteins are IMPs, but not all
IMPs are transmembrane proteins. IMPs comprise a significant fraction of the proteins
encoded in an organism's genome. Proteins that cross the membrane are surrounded by
"annular" lipids, which are defined as lipids that are in direct contact with a membrane
protein. Such proteins can be separated from the biological membranes only using
detergents, nonpolar solvents, or sometimes denaturing agents.

b) Transmembrane proteins
A transmembrane protein (TP) is a type of integral membrane protein that spans the
entirety of the biological membrane to which it is permanently attached. Many
transmembrane proteins function as gateways to permit the transport of specific
substances across the biological membrane. They frequently undergo significant
conformational changes to move a substance through the membrane.
Transmembrane proteins are polytopic proteins that aggregate and precipitate in water.
They require detergents or nonpolar solvents for extraction, although some of them
(beta-barrels) can be also extracted using denaturing agents.

c)Peripheral Proteins
Peripheral membrane proteins are membrane proteins that adhere only temporarily to
the biological membrane with which they are associated. These proteins attach to integral
membrane proteins, or penetrate the peripheral regions of the lipid bilayer. The
regulatory protein subunits of many ion channels and transmembrane receptors, for
example, may be defined as peripheral membrane proteins. In contrast to integral
membrane proteins, peripheral membrane proteins tend to collect in the water-soluble
component, or fraction, of all the proteins extracted during a protein purification
procedure.

Fig 1.5 a) Transmembrane protein attached to bilayer by fatty acid chain; (b) integral membrane protein (e.g. cholinesterase)
attached by fatty acid chain; (c) peripheral membrane protein (e.g. cytochrome c) attached to a transmembrane protein

Glycoproteins

Fig 1.6. - Diagrammatic representation of a glycoprotein embedded in a membrane bilayer

Most of the proteins of the plasma membrane that are exposed to the cell surface have
covalently linked sugars. The sugars may be linked either to the CONH2 side-chain of the amino
acid asparagine, or to the OH in the side-chains of serine or threonine. The sugars are present
as short, branched chains containing from about 4 to 12.
III. Membrane Carbohydrate.

Fig1.7. - Carbohydrates in the


plasma membrane.

Membrane carbohydrates are mostly covalently bound to glycolipids and glycoproteins.


However, some membrane carbohydrates are part of proteoglycans, molecules that insert their
amino acid chain among the lipid fatty acids. Although some carbohydrates can be found
associated to intracellular membranes, the majority are located in the outer monolayer of the
plasma membrane, facing the extracellular space. Membrane carbohydrates are mostly
synthesized in the Golgi complex, and at less extent in the endoplasmic reticulum.

2. Fluid Mosaic Model


The fluid mosaic model proposed by Jonathan Singer and Garth Nicolson in 1972 describes the
dynamic and fluid nature of biological membranes. Lipids and proteins can diffuse laterally
through the membrane. Phospholipids can diffuse relatively quickly in the leaflet of the bilayer
in which they are located. A phospholipid can travel around the perimeter of a red blood cell in
around 12 sec, or move the length of a bacterial cell within 1 sec. Phospholipids can also spin
around on their head-to-tail axis, and their lipid tails are very flexible. These different types of
movements create a dynamic, fluid membrane which surrounds cells and organelles.
Membrane proteins can also move laterally in the bilayer, but their rates of movement vary and
are generally slower than those of lipids. In some cases, membrane proteins are held in particular
areas of the membrane in order to polarize the cell and enable different ends of the cell to have
different functions. One example of this is the attachment of a glycosyl-phosphatidylinositol (GPI)
anchor to proteins to target them to the apical membrane of epithelial cells and exclude them
from the basolateral membrane.
Despite all this movement of lipids and proteins in the bilayer, vertical movement, or flip-flop,
of lipids and proteins from one leaflet to another occurs at an extremely low rate. This is due to
the energetic barrier encountered when forcing the hydrophilic head (in the case of lipids) or
hydrophilic regions (in the case of proteins) through the hydrophobic environment of the inside
of the membrane. This near absence of vertical movement allows the inner and outer leaflets of
the bilayer to maintain different lipid compositions, and enables membrane proteins to be
inserted in the correct orientation for them to function. However, some enzymes facilitate the
process of lipid flip-flop from one leaflet to another. These flippases, or phospholipid trans
locators, use ATP to move lipids across the bilayer to the other leaflet. In eukaryotic cells,
flippases are located in various organelles, including the endoplasmic reticulum (ER), where they
flip-flop newly synthesized lipids.

3. Functions of Plasma Membrane


In all cells, the plasma membrane has several essential functions. These include transporting
nutrients into and metabolic wastes out of the cell; preventing unwanted materials in the
extracellular milieu from entering the cell; preventing loss of needed metabolites and
maintaining the proper ionic composition, pH (7.2), and osmotic pressure of the cytosol. To
carry out these functions, the plasma membrane contains specific transport proteins that permit
the passage of certain small molecules but not others. Several of these proteins, use the energy
released by ATP hydrolysis to pump ions and other molecules into or out of the cell against their
concentration gradients. Small charged molecules such as ATP and amino acids can diffuse freely
within the cytosol but are restricted in their ability to leave or enter it across the plasma
membrane.
In addition to these universal functions, the plasma membrane has other crucial roles in
multicellular organisms. Few of the cells in multicellular plants and animals exist as isolated
entities; rather, groups of cells with related specializations combine to form tissues. Specialized
areas of the plasma membrane contain proteins and glycolipids that form specific contacts and
junctions between cells to strengthen tissues and to allow the exchange of metabolites between
cells. Other proteins in the plasma membrane act as anchoring points for many of the cytoskeletal
fibers that permeate the cytosol, imparting shape and strength to cells. Surrounding most animal
cells is a mixture of fibrous proteins and polysaccharides collectively called the extracellular
matrix. This viscous, water-filled matrix provides a bedding on which most sheets of epithelial
cells or small glands lie. Proteins in the plasma membrane anchor cells to many of the matrix
components, adding to the strength and rigidity of many tissues (Figure 5-40). In addition,
enzymes bound to the plasma membrane catalyze reactions that would occur with difficulty in
an aqueous environment. The plasma membrane of many types of eukaryotic cells also contains
receptor proteins that bind specific signaling molecules (e.g., hormones, growth factors,
neurotransmitters), leading to various cellular responses.
4. How small molecules cross membranes

I. Membrane proteins control what enters and leaves the cell


A vital class of membrane proteins are those involved in active or passive transport of
materials across the cell membrane or other subcellular membranes surrounding
organelles. For a cell or an organism to survive, it is crucial that the right substances enter
cells (e.g. nutrients) and the right substances are transported out of them (e.g. toxins).

II. Passive and active transport


Molecules can cross biological membranes in several different ways depending on their
concentration on either side of the membrane, their size and their charge. Some molecules,
including water, can simply diffuse through the membrane without assistance. However,
large molecules or charged molecules cannot cross membranes by simple diffusion.
Charged molecules such as ions can move through channels passively, down
electrochemical gradients. This movement is described as downhill, as the ions or
molecules travel from an area of high concentration to an area of low concentration. This
requires channel proteins but no energy input. Passive transport can also be mediated by
carrier proteins that carry specific molecules such as amino acids down concentration
gradients, again without any requirement for energy. Active transport moves species
against concentration gradients and requires energy, which is obtained from ATP, from
light, or from the downhill movement of a second type of molecule or ion within the same
transporter.

Fig 4.1. Passive and Active transport across a biological membrane


III. Passive transport
Passive transport is the movement of molecules across biological membranes down
concentration gradients. This type of transport does not require energy. Channels form
water-filled pores and thus create a hydrophilic path that enables ions to travel through the
hydrophobic membrane. These channels allow downhill movement of ions, down an
electrochemical gradient. Both the size and charge of the channel pore determine its
selectivity. Different channels have pores of different diameters to allow the selection of
ions on the basis of size. The amino acids that line the pore will be hydrophilic, and their
charge will determine whether positive or negative ions travel through it. For example, Ca2+
is positively charged, so the amino acids lining the pores of Ca2+ channels are generally basic
(i.e. they carry a negative charge).
Channels are not always open. They can be gated by ligands which bind to some part of the
protein, either by a change in membrane potential (voltage gated) or by mechanical stress
(mechanosensitive).
Carrier proteins are the other class of membrane proteins, apart from channels, which can
facilitate passive transport of substances down concentration gradients. Carrier proteins
transport molecules much more slowly than channels, as a number of conformational
changes in the carrier are required for the transport of the solute across the membrane. A
molecule such as a sugar binds to the carrier protein on one side of the membrane where
it is present at a high concentration. Upon binding, the carrier changes conformation so
that the sugar molecule then faces towards the opposite side of the membrane. The
concentration of sugar on this side is lower, so dissociation occurs and the sugar is released.
This is downhill movement and, although slower than movement through channels, it
requires no energy.

IV. Active transport


The transport of molecules across a membrane against a concentration gradient requires
energy, and is referred to as active transport. This energy can be obtained from ATP
hydrolysis (primary active transport), from light or from an electrochemical gradient of an
ion such as Na+ or H+ (secondary active transport).
Calcium ions signal many events, including muscle contraction, neurotransmitter release
and cellular motility. However, high cytoplasmic concentrations of Ca2+ are toxic to the cell.
Therefore Ca2+ must be tightly regulated and removed from the cytoplasm either into
internal stores (the ER, and the SR in muscle cells) or into the extracellular space. This Ca 2+
removal is carried out by a family of Ca2+-ATPases, including the Sarco/endoplasmic
reticulum Ca2+-ATPase (SERCA), which hydrolyse ATP to move Ca2+ against its
electrochemical gradient into the ER and SR. There are Ca2+-ATPases in the ER, Golgi and
plasma membrane, and despite their sequence similarity, these proteins are differentially
targeted to the appropriate membrane. These Ca2+ pumps are primary active transporters.
SERCA moves two Ca2+ ions into the ER or SR for every ATP molecule that is hydrolysed. The
pump undergoes a cycle of binding ATP and phosphorylation, and undergoes large
conformational changes every time it transports a pair of Ca2+ ions. SERCA is a P-type
ATPase (so called because it is phosphorylated during ion transport). There are many P-type
ATPases, and they are conserved in evolution across many species. The Na +/K+-ATPase is
one of these P-type ATPases, and it works in a similar way to SERCA to pump Na+ out of the
cell and K+ into the cell using energy derived from the hydrolysis of ATP. We have now
obtained three-dimensional structures of SERCA in a number of conformational states,
which allow scientists to visualize the transport process.
Secondary active transport requires an ion electrochemical gradient to drive the uphill
transport of another solute. The downhill movement of one species drives the uphill
movement of the other. This can be symport (in which both types of molecule or ion travel
across the membrane in the same direction) or antiport (in which the two species travel in
opposite directions)

Fig 4.2. - The two types of co-transport are shown, with examples.

V. Symport and Antiport.


In order to transport glucose into cells, the Na+glucose symporter uses the electrochemical
gradient of Na+ across the plasma membrane. The concentration of Na+ is much higher
outside the cell, and the inside of the cell is negatively charged relative to the outside, so
by allowing Na+ to travel down its electrochemical gradient, these transporters can move
glucose uphill, into the cell and against its concentration gradient. This is referred to as
Symport, as both Na+ and glucose travel in the same directionin this case into the cell. In
order for this Symport to be sustainable, the Na+ gradient must be maintained. This is done
by the Na+/K+-ATPase, which uses ATP to pump the Na+ back into the extracellular space,
thus maintaining a low intracellular Na+ concentration.
Both Na+ and Ca2+ are present at much higher concentrations outside the cell than inside it.
Like the Na+ glucose symporter, the Na+ Ca2+ exchanger uses the electrochemical gradient
of Na+ across the plasma membrane to move a second species (Ca2+) against its
electrochemical gradient. However, in this case the transporter is an antiporter, as it uses
the concentration gradient of one substance moving in (Na+) to move another (Ca2+) out of
the cell. This antiporter has an exchange rate of three Na + ions in to two Ca2+ ions out. It
moves Ca2+ out of the cell faster than the plasma membrane equivalents of SERCA, but has
a lower affinity for Ca2+ than these P-type ATPases. Again this transporter relies on the
Na+/K+-ATPase to maintain the low intracellular Na+ concentration.

5. How large particles cross membranes

I. Endocytosis

There are two types of endocytosis, which


differ in the size of the vesicles formed. In
pinocytosis (or cell drinking), fluid or small
particles are taken into small vesicles about
150 nm in diameter. In phagocytosis (or cell
eating), large particles such as micro-
organisms and cell debris are taken into large
vesicles (or vacuoles) about 250 nm in
diameter. Most cells carry out pinocytosis;
only specialized phagocytic cells carry out
Fig 5.1. Endocytosis and Exocytosis across cell
phagocytosis.
membrane

Fig 5.2. - The different types of endocytosis


Many of the particles taken up by either process end up in lysosomes. Phagocytic vesicles
(phagosomes) fuse directly with lysosomes. Pinocytic vesicles are initially transferred to
endosomes, from where they are taken to lysosomes, unless retrieved specifically. The
breakdown products of lysosomal enzyme action are transported across the lysosomal
membrane into the cytosol for reuse. Many of the membrane constituents of endosomes are
recycled to the plasma membrane by exocytosis.

Receptor-mediated endocytosis
Specific macromolecules are taken into cells by receptor-mediated endocytosis. The method
is used, for example, on hormones bound to plasma membrane receptors. There,
receptor/hormone complexes cluster in special regions of the plasma membrane called
coated pits. These are dents in the membrane which have on their cytosolic side a coating of
a protein called Clathrin.
Endocytosis begins with the invagination of the coated pit containing a cluster of receptors
to be engulfed. The clathrin forms a lattice around it, which causes the formation of a coated
vesicle about 80 nm in diameter. The vesicle rapidly loses its clathrin coat, which returns to
the plasma membrane to form a new coated pit. The bare vesicle fuses with an endosome
whose acidic interior (maintained by ATP-driven proton pumps) causes most protein
receptor complexes to dissociate. This sorting enables the proteins and the receptors to be
directed to different destinations.

6. Cell Membrane Receptors


Receptors are protein molecules in the target cell or on its surface that bind ligands. There are
two types of receptors: internal receptors and cell-surface receptors.

I. Internal receptors
Internal receptors, also known as intracellular or cytoplasmic receptors, are found in the
cytoplasm of the cell and respond to hydrophobic ligand molecules that are able to travel
across the plasma membrane. Once inside the cell, many of these molecules bind to
proteins that act as regulators of mRNA synthesis to mediate gene expression. Gene
expression is the cellular process of transforming the information in a cell's DNA into a
sequence of amino acids that ultimately forms a protein. When the ligand binds to the
internal receptor, a conformational change exposes a DNA-binding site on the protein. The
ligand-receptor complex moves into the nucleus, binds to specific regulatory regions of the
chromosomal DNA, and promotes the initiation of transcription. Internal receptors can
directly influence gene expression without having to pass the signal on to other receptors
or messengers.

II. Intracellular Receptors


Hydrophobic signaling molecules typically diffuse across the plasma membrane and interact
with intracellular receptors in the cytoplasm. Many intracellular receptors are transcription
factors that interact with DNA in the nucleus and regulate gene expression.

III. Cell-Surface Receptors


Cell-surface receptors, also known as transmembrane receptors, are cell surface,
membrane-anchored, or integral proteins that bind to external ligand molecules. This type
of receptor spans the plasma membrane and performs signal transduction, converting an
extracellular signal into an intracellular signal. Ligands that interact with cell-surface
receptors do not have to enter the cell that they affect. Cell-surface receptors are also called
cell-specific proteins or markers because they are specific to individual cell types.
Each cell-surface receptor has three main components: an external ligand-binding domain
(extracellular domain), a hydrophobic membrane-spanning region, and an intracellular
domain inside the cell. The size and extent of each of these domains vary widely, depending
on the type of receptor.
Cell-surface receptors are involved in most of the signaling in multicellular organisms. There
are three general categories of cell-surface receptors: ion channel-linked receptors, G-
protein-linked receptors, and enzyme-linked receptors.

IV. Ion Channel-Linked Receptors


Ion channel-linked receptors bind a ligand and open a channel through the membrane that
allows specific ions to pass through. To form a channel, this type of cell-surface receptor
has an extensive membrane-spanning region. In order to interact with the phospholipid
fatty acid tails that form the center of the plasma membrane, many of the amino acids in
the membrane-spanning region are hydrophobic in nature. Conversely, the amino acids
that line the inside of the channel are hydrophilic to allow for the passage of water or ions.
When a ligand binds to the extracellular region of the channel, there is a conformational
change in the protein's structure that allows ions such as sodium, calcium, magnesium, and
hydrogen to pass through.
Gated ion channels form a pore through the plasma membrane that opens when the
signaling molecule binds. The open pore then allows ions to flow into or out of the cell.

V. G-Protein Linked Receptors


G-protein-linked receptors bind a ligand and activate a membrane protein called a G-
protein. The activated G-protein then interacts with either an ion channel or an enzyme in
the membrane. All G-protein-linked receptors have seven transmembrane domains, but
each receptor has its own specific extracellular domain and G-protein-binding site.
Cell signaling using G-protein-linked receptors occurs as a cyclic series of events. Before the
ligand binds, the inactive G-protein can bind to a newly-revealed site on the receptor
specific for its binding. Once the G-protein binds to the receptor, the resultant shape change
activates the G-protein, which releases GDP and picks up GTP. The subunits of the G-protein
then split into the subunit and the subunit. One or both of these G-protein fragments
may be able to activate other proteins as a result. Later, the GTP on the active subunit of
the G-protein is hydrolyzed to GDP and the subunit is deactivated. The subunits
reassociate to form the inactive G-protein, and the cycle starts over.

VI. Enzyme-Linked Receptors


Enzyme-linked receptors are cell-surface receptors with intracellular domains that are
associated with an enzyme. In some cases, the intracellular domain of the receptor itself is
an enzyme or the enzyme-linked receptor has an intracellular domain that interacts directly
with an enzyme. The enzyme-linked receptors normally have large extracellular and
intracellular domains, but the membrane-spanning region consists of a single alpha-helical
region of the peptide strand. When a ligand binds to the extracellular domain, a signal is
transferred through the membrane and activates the enzyme, which sets off a chain of
events within the cell that eventually leads to a response. An example of this type of
enzyme-linked receptor is the tyrosine kinase receptor. The tyrosine kinase receptor
transfers phosphate groups to tyrosine molecules. Signaling molecules bind to the
extracellular domain of two nearby tyrosine kinase receptors, which then dimerize.
Phosphates are then added to tyrosine residues on the intracellular domain of the receptors
and can then transmit the signal to the next messenger within the cytoplasm.
References

1. Molecular Cell Biology. 4th edition. Lodish H, Berk A, Zipursky SL, et al. New York: W. H.
Freeman; 2000.

2. The Cell: A Molecular Approach. 2nd edition. Cooper GM.

3. Medical cell biology. Academic Press. Steven R. Goodman (2008).pp. 37. ISBN 978-0-12-
370458-0. Retrieved 24 November 2010.

4. Singer SJ, Nicolson GL (Feb 1972). "The fluid mosaic model of the structure of cell
membranes". Science. 175 (4023): 72031. doi:10.1126/science.175.4023.720. PMID
4333397.

5. Fundamentals of Biochemistry: Life at the Molecular Level (4 Ed.). Wiley. ISBN 978-
1118129180.

6. Membrane Receptors - Annual Review of Biochemistry, Vol. 43:169-214 (Volume publication


date July 1974) DOI: 10.1146/annurev.bi.43.070174.001125

7. Milsom, D.W. (1994) Biological Membranes. Biological Sciences Review 6(4), 1620

8. "The structural era of endocytosis". Marsh, M.; McMahon, HT (July 1999). Science. 285
(5425): 21520. doi:10.1126/science.285.5425.215. PMID 10398591. Retrieved 2009-06-19.

Das könnte Ihnen auch gefallen