Transformer protocol Student’s guide THE

advertisement
THE
Transformer
protocol
Student’s guide
INTRODUCTION
Bacteria everywhere
Bacteria are the commonest living things on Earth. Despite the
fact that they are microscopic, together they weigh more than all
the planet’s plants and animals combined. Because some bacteria
can cause disease they tend to receive a bad press. However,
most bacteria are harmless and they perform a vital role by
helping to recycle elements, breaking down dead organisms to
simpler materials so that they can be used again. They also make
nutrients available to other living things by, for example, extracting
nitrogen from the air for protein production by plants. Even
some of the bacteria that inhabit our intestines are useful — for
example, Escherichia coli (E. coli for short) provides us with essential
vitamins.
Inside a living cell, plasmids can be copied, but they rely upon
the molecular machinery of the cell carrying them to do this
job. So, in the right conditions, a test tube of bacteria will grow
and multiply, but a test tube of plasmids will just sit there. Neither
plasmids nor the DNA from which they are made are alive.
Plasmids are duplicated separately from the bacterial
chromosome, and not just when the cell divides.
Consequently one bacterial cell may hold many identical
copies of plasmids.
old
strand
new strand
Cells express themselves
Bacterial cells are enclosed by cell membranes containing the
cytoplasm. Most bacteria also have a cell wall surrounding the
membranes. Some have one or more thread-like hairs (flagella)
to propel themselves along.
Inside the cell there’s just one chromosome, in the form of a
ring. The chromosome is made of DNA, which carries the
instructions needed for making proteins. A set of instructions
for making all or part of a protein is called a gene. The genes in
bacteria are arranged one after the other, like beads on a string,
along the chromosome. The genetic ‘recipe’ of E. coli is 4.6 million
base pairs long — that’s enough information for making about
3 000 proteins. Fortunately, every E. coli has about 15 000 proteinmaking ‘factories’ (ribosomes) where the genetic instructions
can be translated and the genes ‘expressed’.
Just before bacteria divide to form two new cells (which, in
favourable conditions, can happen every 30 minutes or so),
the whole chromosome is copied so that each new cell ends
up with a full set of instructions.
Chromosome
Plasmids
cytoplasm
Ribosomes
sites of protein
synthesis
Cell
wall
Not to scale
Cell membranes
Some of the main components of a bacterial cell.
2
old
strand
new strand
Nucleotide
5'
3'
Phosphate
Deoxyribose sugar
3'
5'
Complementary
base pairs
DNA structure and function. The structure of DNA can be
likened to that of a twisted rope ladder. At the sides are chains of
sugar and phosphate molecules. In the centre are the bases: adenine
(A), guanine (G), cytosine (C) and thymine (T). Hydrogen bonds
between the bases hold the two helices together—A always pairing
withT, and C always pairing with G.Through the double helix, genetic
information is copied from one cell or generation to the next.When
DNA is duplicated, the two complementary strands are untwisted,
and nucleotides are brought into place alongside each of the old
strands to form new ones.The new copies are checked to ensure
accurate copying. Several different enzymes perform these tasks.
The order of bases in the DNA specifies the sequence of the
amino acids in proteins. Three bases in a row (a triplet) specify
each amino acid. A particular gene — a length of DNA —
determines the structure of all or part of a specific protein.
A little extra DNA helps
Bacterial ‘sex’
Sometimes other smaller rings of DNA, called plasmids, are also
found in bacteria. Plasmids carry just a few genes. They are not essential
for the bacteria, but they may help them to survive in some
environments. For instance, some plasmids help the bacteria that
carry them to resist the toxic effects of heavy metals, or to live on
particular nutrients.
Plasmids are sometimes transferred between bacterial cells in a
natural ‘mating’ process called conjugation. This is as close as
bacteria get to sex (bacteria don’t actually reproduce by sexual
means). Some bacteria can also pick up extra DNA from their
surroundings. This is not as easy as it sounds, as DNA molecules
can be very large and they have to cross the bacterial cell
membrane somehow. Special proteins are needed to help ferry
the DNA across and since only a few species have these proteins,
this sort of DNA transfer (called ‘transformation’) is thought to
be fairly rare in nature.
Entry restrictions
Even if DNA (a plasmid or smaller fragments) makes it across
the cell membrane, many bacteria can destroy the incoming
genetic message. This is particularly important if the DNA is
from a bacteriophage — a type of virus that preys upon bacteria.
The bacterial defence mechanism consists of enzymes —
restriction enzymes — whose job is to ‘restrict’ the invasion of
viruses. There are many hundreds of different restriction
enzymes that ‘recognise’ particular sequences of DNA and cut
it up, like very precise molecular scissors. The bacteria’s own
DNA escapes damage by ‘disguising’ the sites where the
restriction enzymes cut — by adding methyl (–CH3) groups to
the DNA bases.
A computer model of restriction enzyme EcoRI.
This enzyme is obtained from Escherichia coli, strain RI. The
DNA is shown at the top of the picture, as a ball-and-stick
model, looking down the axis of the double helix. The enzyme
wraps around the DNA and moves along it,‘searching’ for a site
where the enzyme can cut. EcoRI cuts at G AATTC.
➔
Bacteria perform useful tasks
It’s not only bacteria that benefit from restriction enzymes.
They’re now an important tool used by medical researchers and
biologists of all kinds.
Parts of the enzyme’s structure are shown here: flat sheets (betapleated sheets) in yellow; coils (alpha helices) in pink and turns
in blue and white.
With restriction enzymes almost any section of DNA, and
consequently any single gene, may be cut out at will. The end of
one DNA molecule can be joined to another that’s been cut
with the same enzyme. The combined DNA can then be put
into a cell in which it may be expressed and duplicated so that it
passes from one cell division to the next. For microorganisms,
one of the most successful methods for transferring genes is to
‘paste’ them into plasmids. The result is a ring of ‘recombinant’
DNA that can be put into a bacterium. Specialised plasmids can
be used to ferry genes from bacteria into yeast cells or even plants.
The data for this image was obtained from the Nucleic Acid
Database (http://ndbserver.rutgers.edu/NDB/ndb.html) where
you can view the computer model in three dimensions. (Search
for the protein with the ID code PD0055.)
A cause for concern?
Genetic modification of this type has already given us new and
improved medicines and vaccines, and altered crop plants with
new characteristics. As a research tool, genetic modification has
helped us to study our own genes and those of other species,
and this work will have major consequences for human health.
In this way, microbes can be ‘re-programmed’ to make a wide variety
of useful substances including vaccines, insulin and other hormones
and materials such as plastics, helping to reduce the consumption of
valuable oil reserves. Microbes can be altered to clean up toxic wastes,
protect us from food poisoning and help in biological research.
However, many people are concerned about the potential dangers
of ‘meddling with genes’ and doubts have been raised about the
wider economic, social and environmental impact that such work
may have.
How bacteria are genetically modified. First, the gene
of interest is cut out, using carefully selected restriction enzymes.
Next, some plasmid DNA is cut with the same enzymes. The
two fragments are joined using DNA ligase to form a recombinant
plasmid. The plasmid DNA is put into a suitable host strain of
bacteria.The transformed (genetically-modified) cells are selected.
The plasmid DNA is copied within the bacteria, and the proteins
it encodes are produced.
DNA
Restriction
enzyme
cuts DNA
at specific
sites
Gene
isolated
So that you can better understand and assess this technology,
this kit provides a practical activity which will allow you to
investigate some of the key techniques, coupled with a discussion
task so that you can begin to think about and critically assess
some of the wider implications of genetic modification.
DNA ligase bonds
DNA fragments
New protein made
by bacteria
Plasmid put into
bacterium
Plasmid
copied
Gene spliced
into plasmid
Cell divides
Transformed cell
Bacterial
chromosome
3
‹Ž‘’
“
”­
‹
­
Œ
…
†
•‚–ƒα€
­
†
—˜Œ­
†™
­
†
‚
‡
†
†
ˆ‰†
Š†
‚ƒ
‚ƒ
 ­
€‚ƒ„…
…
†

†

†
…‰
†


š
™
…

 ­€‚€ƒ‚„…†
‡ƒ
ˆ
™š›
•ˆˆ†…ˆ
•
†
ˆ•

ˆ
•
ˆ
‰€Š†­‹„…†ˆ
Œ ˆ
Ž
­‹„…ˆ

—˜

‰€Š
‘†‡€Š
ˆ’“
‚€Šˆ”
­
’‡€Š”
‰€Š
—˜

‰€Š
­
’‡€Š”

†
ˆ
ˆŽ††­‹„…ˆ
…ˆ
•
†

ˆ
–






†


…
“…

‰
‚
­
β
“
“

“
“
•
œ
ADVANCED INFORMATION
Transformation
Transformation is the uptake and expression of DNA by cells.
Although some bacterial species can take up DNA from their
environment naturally, most cannot. Cells that can take up
DNA in this way are termed ‘competent’. However, natural
transformation is a relatively rare event. Most bacteria have
not evolved the membrane proteins that allow foreign DNA
to be ‘recognised’ and absorbed. Without these special
mechanisms, gene-sized DNA molecules are too large to
diffuse or be transported through the cell membranes.
Chemical transformation
Escherichia coli (the common gut bacterium) is the best-understood
and most studied organism on Earth. It is not naturally
competent. However, in 1970 a method was developed to
artificially transform this and other bacteria. Today, such
transformation is a key process used in the genetic modification
of microbes.
The original method involved suspending rapidly-dividing cells
in a ‘transformation buffer’ of cold calcium chloride solution
then subjecting them to a brief heat shock in the presence of
the DNA to be taken up. Later it was found that other divalent
cations (such as magnesium, Mg2+, manganese Mn2+, and barium
Ba2+) all had a similar effect to that of calcium ions. The exact
mechanism of DNA uptake is poorly understood. One
hypothesis is as follows:
●
●
●
●
Cooling the cells to 0 °C stabilizes the normally fluid bacterial
cell membranes.
Positively-charged ions in the transformation buffer (e.g., Ca2+)
are then able to bind to the negatively-charged phosphate
groups in DNA and the phospholipids of the cell
membranes, shielding their negative charges.
This allows the plasmids to approach the membrane and
the channels through it that are formed where the outer and
inner cell membranes meet.
The heat shock helps to force the plasmids through the
channels by creating a thermal imbalance on either side of
the cell membranes.
Other transformation methods
In the last decade, other methods of transforming bacteria have been
introduced. These include, for example, subjecting the bacteria to
brief pulses of very high voltages. This technique — called
electroporation — punches holes in the bacteria through which DNA
can enter. This is the most efficient method devised so far.
A procedure called particle bombardment or ballistic impregnation
is also sometimes used to introduce DNA into
cells. With this method, the DNA is first
stuck onto minute tungsten or gold
beads. Using a ‘gene gun’, the DNAcoated particles are fired into the cells.
This technique is often used for
transforming plant cells.
‘Chemical transformation’ remains popular
however, because it is simple, inexpensive and does
not require specialist equipment or materials.
6
Plasmid
DNA
Plasmid DNA entering a bacterial cell. Both DNA and
the bacterial cell membrane normally have a negative charge,
so they repel one another. However, solutions of positive ions
can be used to neutralise both the DNA and the cell surface,
allowing the plasmid DNA to enter cells through pores in the
bacterial membranes.This is a schematic diagram, showing the
general principle involved. In reality, the relative sizes of
components and the distribution of charges will differ from that
shown here.
Selectable marker genes
The transformation process is very inefficient and only a small
proportion of the cells treated will take up plasmid DNA.
Therefore a means of selecting those cells that have been
transformed is needed. Antibiotic resistance markers are often
used for this purpose. The p2k plasmid includes a kanamycin
resistance gene (called APH(3')-I) put there specifically to act as
a genetic marker.
The strain of Escherichia coli used here is incapable of
hydrolysing the sugar lactose, because it lacks the gene for
the enzyme lactase (β-galactosidase). However, the p2k
plasmid carries this gene, and if a bacterium takes up the
plasmid it gains the ability to hydrolyse lactose.
The colourless compound X-Gal is hydrolysed by lactase,
yielding galactose and an insoluble indigo dye. The dye is
precipitated within the bacteria, enabling X-Gal to be used
as an indicator of lactase activity (transformed cells are blue).
Why can’t lactase action and this blue colour alone therefore be
used to select transformed cells?
Cells with resistance plasmids are normally disadvantaged
compared to their neighbours without them. In the presence of
appropriate antibiotics however, such plasmid-bearing cells thrive
while their less well-endowed neighbours perish. In this way,
selection pressure is applied to maintain the plasmid in the
population of cells.
Without that pressure, the few transformed cells would be swamped
by their untransformed neighbours, and the plates would be covered
by a uniform ‘lawn’ of ordinary bacterial cells rather than individual
blue colonies.
How kanamycin acts on bacteria
amino
acid
Kanamycin kills bacteria by stopping protein synthesis at the
ribosomes. Unlike several other antibiotics (e.g., penicillin or
ampicillin) kanamycin kills all cells, rather than just those that are
actively growing. Consequently, bacteria transformed with p2k
require a short ‘recovery period’ before they are transferred onto
kanamycin-containing plates. This allows the resistance marker
gene to be expressed, and ensures that the transformed cells are
not killed on contact with kanamycin-containing growth medium.
Transfer
RNA
small (30S)
subunit
AUG
UAC
large (50S)
subunit
While it would be more practical to use a marker (like ampicillin
resistance) that did not need a recovery period, there are several
compelling reasons for using kanamycin and kanamycin resistance
(see below).
How the resistance mechanism works
There are very many different mechanisms that confer resistance
to the effects of kanamycin and related antibiotics. At least seven
of these mechanisms work by transferring a phosphate group
onto the kanamycin, altering its structure.
The APH(3')-I gene encodes an enzyme that catalyses the transfer
of a phosphate group from ATP onto a hydroxyl (–OH) group
of kanamycin. The modified antibiotic which results is unable
to bind to the bacterial ribosome, so the antibiotic is inactivated.
This particular resistance mechanism has been found to occur
in about 50% of gram-negative bacteria.
The enzyme which confers the resistance is relatively unstable,
and it is inactivated by increased temperatures or pH changes.
Its requirement for ATP means that this enzyme can only
function in environments where that compound is abundant
(e.g., inside cells).
Use of kanamycin and the resistance gene
Many transformation experiments use plasmids that confer
resistance to the antibiotic ampicillin. However, in the
construction of p2k we have chosen to incorporate a kanamycin
resistance marker for several important reasons:
●
●
●
●
●
unlike ampicillin, kanamycin is very seldom used to treat human
disease, having been superseded by other drugs;
it is needed in small amounts in culture plates (a tenth of the
concentration normally used for ampicillin);
unlike ampicillin, kanamycin is not absorbed by the gut (in
clinical use, it has to be injected). Therefore the safety hazard
posed by accidental ingestion is reduced;
ampicillin resistance enzymes (β-lactamases) often provide
resistance to many other similar antibiotics whereas this
particular kanamycin resistance gene affects a lesser range of
antibiotics of limited use. APH(3')-I confers resistance mainly
to kanamycin and neomycin;
for several other technical reasons, the use of kanamycin
resistance markers is now widely accepted as safe, even in food.
Scientists disagree, however, about the wisdom of using
ampicillin resistance markers in such products. This is mainly
because of the slight risk that the marker gene will pass into
other organisms, giving them the ability to withstand not only
ampicillin, but other antibiotics too.
anticodon
CCA
5'
GGC
AUGCCGGGUUACUUA
AUGCCGGGUUACU
AUGCCGGGUUAC
3'
Messengerr RNA
NA
codon
Bacterial ribosome
Protein synthesis at a bacterial ribosome. Successive
transfer RNA (tRNA) molecules, each carrying an amino acid, are
brought to the ribosome according to the genetic code of the
messenger RNA (mRNA). The amino acid residues are strung
together to make a protein. Kanamycin interferes with this process
by binding irreversibly to the 30S sub-unit of the bacterial ribosomes.
The kanamycin/ribosome complex is able to start protein synthesis
by binding to mRNA and the first tRNA. However, the second
tRNA cannot bind, and the mRNA/ribosome complex dissociates.
Further investigations
There are numerous variations on the technique described
here which can be attempted to try to improve the
efficiency of transformation.
You could investigate the effect of changing:
•
•
•
•
the age of the host cells used;
the amount of plasmid DNA used;
the duration of the heat shock;
the intensity of the heat shock
(i.e., its temperature);
• the duration and/or temperature of the
recovery period.
To determine the effect of altering these factors, it is useful
to calculate the transformation efficiency. This is expressed
as the number of transformed colonies produced per µg
of plasmid DNA. The transformation efficiency can be
calculated as follows:
1. Calculate the mass, in µg, of plasmid DNA used in Step 4.
Concentration of the plasmid DNA x Volume of plasmid
DNA solution used = Mass of plasmid.
2. Determine how much of the cell suspension you spread,
in µL, onto the LB/antibiotic/X-Gal plate. Volume of
suspension spread / Total volume of suspension = Fraction
of cell suspension spread.
3. Calculate the mass of plasmid contained in the cell
suspension spread onto the LB/antibiotic/X-Gal plate. Mass
of plasmid x Fraction of cell suspension spread = Mass of
plasmid spread.
4. Determine the number of colonies per µg of plasmid
DNA. Colonies counted / Mass of plasmid spread (µg)
= Transformation efficiency.
7
SAFETY PRECAUTIONS
Resistance to the effects of some important
antibiotics is now widespread in several species
of disease-causing microbes. Special steps have
been taken to ensure that the procedures
followed in this practical exercise do not
contribute to this problem.
Antibiotic resistance has evolved to give
bacteria the ability to thrive in environments
containing antibiotics secreted by other
microorganisms. Resistant bacteria often
produce proteins that inactivate specific
antibiotics or stop them from working in some
way (e.g., by preventing their transport into
bacterial cells). Resistance genes are often
carried on plasmids, which can pass from one
bacterial cell to another of the same or a related
species by a natural ‘mating’ process called
conjugation.
During conjugation, a tube or pilus is formed
between adjacent cells, through which the
plasmid passes. The genes required for the
formation of the pilus are also carried on a
plasmid (an F or fertility plasmid).
The bacterial strain provided in this kit
does not carry an F plasmid. Nor does it
carry any bacteriophages which may pick
up DNA and transfer it to other cells.
Deleted genes
For a plasmid to travel through a pilus, two
additional requirements must be met. The
plasmid must possess a gene encoding a
mobility protein (mob) and have a nic site. The
mobility protein nicks the plasmid at the nic
site, attaches to it there and conducts the
plasmid through the pilus. p2k has neither a
nic site nor the mob gene.
This means that once it has been
introduced into a bacterial cell by artificial
means (transformation) the plasmid
cannot naturally transfer (by conjugation)
into other cells that do not possess it.
1 F plasmid unwound
and replicated
The F plasmid
has about 30 genes.
Bacterial
chromosome
Sex
pilus
These include genes
for making the pilus
and for a mobility
(mob) protein.
The mob protein
conducts the plasmid
through the pilus.
F plasmid
2 Single-stranded DNA
passes into recipient cell
through pilus
Recipient
F- strain
3 Complementary
DNA synthesised,
forming new F plasmid
Conjugation between bacteria
Conjugation involves one-way transfer of DNA from a donor (‘male’) to a recipient
(‘female’) strain, through a tube called a sex pilus.The pilus is made by the donor
cell using genes encoded by a specialised F plasmid. The F plasmid can also
temporarily become part of the bacterial chromosome.There it can pick up extra
genes that are carried with it when it later passes into another cell by conjugation.
F plasmids with these extra genes are called F' plasmids.
Biological containment
The bacterial strain used in this procedure is non-pathogenic. It has been selected
for its suitability for work of this type and its inability to survive outside the
laboratory.
Over many years of laboratory use and millions of generations, changes have also
occurred in the lipopolysaccharides on the outer membrane of this bacterial strain.
These changes mean that it is not possible for this strain of E. coli to colonise
the mammalian gut.
Physical containment
In addition to the biological containment measures described above, the practical
procedure requires that good microbiological practice is followed to ensure that
the microorganisms are physically contained during the investigation and destroyed
afterwards.
Mass
Volume
1 gram (g) = 1 000 milligrams (mg)
1 milligram (mg) = 1 000 micrograms (µg)
1 microgram (µg) = 1 000 nanograms (ng)
1 litre (L) = 1 000 millilitres (mL)*
1 millilitre (mL) = 1 000 microlitres (µL)
Nucleic acid
1 kilobase (kb) = 1 000 bases
1 megabase (Mb) = 1 000 kilobases (kb)
8
Donor
F+ strain
Transfer of antibiotic resistance
NOTE
Some people prefer to use the cubic decimetre (dm3) and cubic
centimetre (cm3) in preference to the litre and millilitre, as S.I.
units for volume are derived from those for length.
National Centre for Biotechnology Education,The University of Reading, PO Box 228,Whiteknights, Reading, RG6 6AJ.
Telephone: 0118 987 3743 Fax: 0118 975 0140 eMail: NCBE@reading.ac.uk Web: www.ncbe.reading.ac.uk
Copyright © Dean Madden, 2000 ISBN: 0 7049 1372 0
Download