Добавил:
Опубликованный материал нарушает ваши авторские права? Сообщите нам.
Вуз: Предмет: Файл:
Скачиваний:
46
Добавлен:
17.08.2013
Размер:
343.56 Кб
Скачать

74 ANALYSIS OF DNA SEQUENCES BY HYBRIDIZATION

BOX 3.3

STACKING PARAMETERS AT OTHER TEMPERATURES

In practice, it turns out to be a sufficiently accurate approximation to assume that the

various

H s0are'sindependent of temperature. This is a considerable simplification. It

allows direct integration of the van’t Hoff relationship to compute thermodynamic parameters at any desired temperature from the parameters measured at 298 K. Starting from the usual relationship for the dependence of the equilibrium constant on temperature,

 

d (ln K )

 

d (1 /T )

we can integrate this directly:

 

 

 

 

R

 

d (ln K )

H

H 0

R

0 d T1

With

limits of integration

from

T 0 , which is 298 K, to

 

any other

melting temperature,

T m

, the result is

 

 

 

 

 

 

 

 

 

 

 

 

 

 

ln K (T m

) ln K (T 0 )

H 0

1

 

1

 

 

 

 

 

 

 

 

 

 

 

 

 

R

 

T m

T 0

 

 

 

 

 

 

 

 

However, the first term of the left-hand side of the equation, for a pair of complemen-

 

 

 

 

tary

oligonucleotides,

is

just 4/

C T . The second term on the

left-hand side is equal to

G

(T 0 )/RT 0 . Inserting these values and rearranging gives a final, useful expression for

computing the

T m

of a duplex at any temperature from data measured at 298 K. All of

the necessary data are summarized in Table 3.1.

 

 

 

 

 

 

 

 

 

T m

T 0

H

0

G

H 0

0 (T 0 ) RT 0 ln( C T / 4)

THERMODYNAMICS OF IMPERFECTLY PAIRED DUPLEXES

 

 

 

 

In contrast to the small number of discrete interactions that must be considered in calcu-

 

lating the energetics of perfectly paired DNA duplexes, there is a plethora of ways that

 

 

duplexes can pair imperfectly. We have available model compound data on most

of

these

 

 

so that estimates of the energetics can be made. However, the large number of possibili-

 

 

ties precludes a complete analysis of imperfections in the context of

all

possible

se-

 

quences, at least for the present.

 

 

 

 

 

 

The simplest imperfection is a dangling end, as shown in Figure 3.7

 

 

a.

If both ends are

dangling, their contributions can be treated separately. From available data

it

appears that

 

a dangling

end contributes

8 kcal/mol on average to

the overall

H of

duplex forma-

tion, and

1 kcal/mol to the

overall

G of duplex

formation. The

large enthalpy arises

 

 

THERMODYNAMICS OF IMPERFECTLY PAIRED DUPLEXES

75

TABLE

3.2 Predicted and Observed Stabilities

 

 

 

 

(free energy of formation) of Various

 

 

 

 

 

 

Oligonucleotide Duplexes

 

 

 

 

 

 

 

 

 

 

 

 

 

Comparison of Calculated and Observed

 

 

G

(kcal/mol)

 

 

 

 

 

 

 

 

 

 

Oligomeric Duplex

 

G

pred

G obs

 

 

 

 

 

 

 

 

 

1

GCGCGC

 

11.1

 

11.1

 

 

 

CGCGCG

 

 

 

 

 

 

2

CGTCGACG

 

11.2

11.9

 

 

 

GCAGCTGC

 

 

 

 

 

 

3

GAAGCTTC

 

7.9

 

8.7

 

 

 

CTTCGAAG

 

 

 

 

 

 

4

GGAATTCC

 

9.3

 

9.4

 

 

 

CCTTAAGG

 

 

 

 

 

 

5

GGTATACC

 

6.7

 

7.4

 

 

 

CCATATGG

 

 

 

 

 

 

6

GCGAATTCGC

 

16.5

15.5

 

 

 

CGCTTAAGCG

 

 

 

 

 

 

7

CAAAAAG

 

6.1

 

6.1

 

 

 

GTTTTTC

 

 

 

 

 

 

8

CAAACAAAG

 

9.3

10.1

 

 

 

 

GTTTGTTTC

 

 

 

 

 

 

9

CAAAAAAAG

 

9.9

9.6

 

 

 

 

GTTTTTTTC

 

 

 

 

 

 

10

CAAATAAAG

 

8.5

8.5

 

 

 

GTTTATTTC

 

 

 

 

 

 

11

CAAAGAAAC

 

9.3

9.5

 

 

 

 

GTTTCTTTC

 

 

 

 

 

 

12

CGCGTACGCGTACGCG

32.9

34.1

 

 

 

 

 

GCGCATGCGCATGCGC

 

 

 

 

 

 

 

 

 

 

 

Note: Calculations use the equations given in the text and the mea-

 

 

 

 

sured thermodynamic values given in Table 3.1.

 

 

 

 

 

 

Source:

Adapted from Breslauer et al. (1986).

 

 

 

 

because the first base of the dangling end can still stack on the

last base pair of the

du-

 

 

plex. Note that there are two distinct types of dangling ends: a 3

 

 

 

 

 

-overhang and a 5

-over-

hang. At the current level of available information, we treat these as equivalent. Simple

 

 

dangling ends will arise whenever a target and a probe are different in size.

 

 

 

 

 

The next imperfection, which leads to considerable destabilization, is an internal mis-

 

 

match. As

shown

in Figure

3.7

b, this leads to the

loss

of two internal

and two inH- s0 's

 

ternal

G

s0 'sApparently. this is empirically compensated by some

residual

stacking

ei-

 

 

ther between the bases that are mispaired and the duplex borders or between the bases

 

 

 

themselves. Whatever the detailed mechanism, the result is to

gain

back

about

 

 

8

kcal/mol

in

H.

There is no effect on the

G.

A larger internal mismatch is called an

 

internal

loop.

Considerable data on the thermodynamics of

such loops exist for RNA,

 

 

and much

less

for DNA. In general, such structures will be

far less stable than the

 

 

perfect duplex

because of

the loss

of additional free energies and

enthalpies of

stacking.

 

 

76 ANALYSIS OF DNA SEQUENCES BY HYBRIDIZATION

Figure 3.7

Stacking interactions

in various types of imperfectly matched duplexes.

(a)Dangling

ends.

(b)Internal mismatch.

(c)Terminal mismatch.

 

A related imperfection is a bulge loop in which two strands of unequal size come together

 

 

 

so that one is perfectly paired while the other has an internal loop of unpaired residues.

 

 

 

There

can

also

be internal loops in which the single-stranded

regions are of different

 

 

 

length on the two strands. Again the data describing the thermodynamics of such struc-

 

 

 

tures are available mostly just for RNAs (Cantor and Schimmel, 1981). There are many

 

 

 

complications. See, for example, Schroeder et al. (1996).

 

 

 

 

A key imperfection that needs to be

considered

to understand the specificity of

 

 

 

hybridization is

a terminal

mismatch. As shown

in Figure

3.7

c, this

results in

the

loss of

one

H

0

one

0

an external mismatch

should be

less destabilizing than an

 

 

 

and

. SuchG

 

 

 

 

 

s

 

s

 

 

 

 

 

 

 

internal mismatch because one less stacking interaction is disrupted. It is not clear, at the

 

 

 

present time, how much any stacking of the mismatch to the end of the duplex partially

 

 

 

compensates for the lost duplex stacking. It may be adequate to model an end-mismatch

 

 

 

like

a dangling

end. Alternatively, one might consider it like an internal mismatch with

 

 

 

only one lost set of duplex stacking interactions. Both of these cases require no correction

 

 

 

to the predicted

G of duplex formation, once the lost stacking interactions have been ac-

 

 

 

counted for. Therefore at present it is simplest to concentrate on

G

estimates

and

ignore

H

estimates. More studies are needed in this area to clear up these uncertainties. In

 

 

 

Chapter 12 we will discuss attempts to use hybridization of short oligonucleotides to infer

 

 

 

KINETICS OF THE MELTING OF SHORT DUPLEXES

77

Figure 3.8 Equilibrium melting behavior of a perfectly matched oligonucleotide duplex and a duplex with a single internal mismatch.

sequence information about a larger

DNA target. As interest in this approach becomes

 

more developed, the necessary model

compound data will doubtlessly be accumulated.

 

Note that there are 12 distinct terminal mismatches, and their energetics could depend not

 

only on the sequence of the neighboring duplex stack but also on whether one or both

 

strands of

the

duplex are further elongated. A large number of model compounds will

 

need to be investigated.

 

 

 

 

 

Despite the limitations and complications just described, our current ability to predict

the effect of a mismatch on hybridization is quite good. A typical example is shown in

 

Figure 3.8. Note

that the

melting

transitions of oligonucleotide

duplexes with 8 or more

base pairs

are quite

sharp. This reflects the large

and

0

of the

the all-or-Hnone 'snature

 

 

 

 

 

 

 

s

 

reaction. The key practical concern is finding a temperature where there is good discrimi-

 

nation between the amount of perfect duplex formed and the amount of mismatched du-

 

plex. The results shown in Figure 3.8 indicate that a single internal mismatch in a 14-mer

leads to a 7°C decrease in

T m . Because of

the sharpness of the melting

transitions, this

can easily be translated into a discrimination of about a factor of ten in amount of duplex

 

present.

 

 

 

 

 

 

 

 

KINETICS OF THE MELTING OF SHORT DUPLEXES

 

 

The major issue we want to address here is whether, by kinetic studies of duplex dissocia-

 

tion, one can achieve a comparable or greater discrimination between the stability of dif-

 

ferent duplexes than by equilibrium melting experiments. In practice, equilibrium melting

 

experiments

are

technically

difficult

because one must discriminate between free and

 

bound probe or target. If the analysis is done spectroscopically, it can be done in homoge-

 

neous solution, but this usually requires large amounts of sample. If it is done with a ra-

dioisotopic tag (or a fluorescent label that shows no marked change on duplex formation),

 

it is usually done by physical isolation of the duplex. This assumes that the rate of duplex

 

dissociation is slow compared with the rate of physical separation, a constraint that usu-

 

ally can be met.

 

 

 

 

 

 

 

In typical hybridization formats a complex is formed between probe and target, and

then the immobilized complex is washed to remove excess

unbound

probe (Fig. 3.9).

 

78 ANALYSIS OF DNA SEQUENCES BY HYBRIDIZATION

Figure 3.9 Schematic illustration of the sort of experiment used to measure the kinetics of duplex melting.

The complex is stored under conditions where little further dissociation occurs, and af-

 

terward

the amount of

probe

bound to the target is measured. A variation on this theme

 

is to allow the complex to

dissociate for a fixed time period and then to measure the

 

amount of probe remaining bound. It turns out that this procedure gives results very sim-

 

ilar to equilibrium duplex formation. The reason is that the temperature dependence for

 

the reaction between two long single strands to form a duplex is very slight. The activa-

 

tion energy for this reaction has been estimated to be only about 4 kcal/mol. Figure 3.10

 

shows

a

hypothetical

reaction profile for interconversion between duplex and single

 

strands. If we assume that the forward and reverse reaction pathways are the same, then

 

the

activation energy

for

the

dissociation of the double strand will be just

H 4

kcal/mol. Since

H

is much larger than 4 kcal/mol for most duplexes, the temperature

 

dependence of the reaction kinetics will mirror the equilibrium melting of the duplex.

 

Figure 3.10 Thermodynamic reaction profile for oligonucleotide melting. Shown is the enthalpy,

H, as a function of strand separation.

It is simple to estimate the effect of temperature on the dissociation rate. At

T m we ex-

pect an equation of the form

 

 

 

 

 

 

km

A exp

 

 

[ H

4]

 

RT

m

 

 

 

 

 

 

 

 

 

 

 

 

KINETICS OF MELTING OF LONG DNA

79

to apply, where

A is a constant, and the

exponential term reflects the number of duplexes

 

that have enough energy to exceed the activation energy. At any other temperature,

 

 

 

 

 

 

 

 

T d ,

 

 

kd

A exp

 

 

[ H

4]

 

 

 

 

 

 

 

RT

d

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

The rate constants

km and kd are

for duplex melting

at temperatures

 

 

 

 

T m and

T d , respec-

tively. We define the time of half release at

 

 

 

T

d

to be

twash

 

, and the time of half release at

T m

to be t1/2 . Then simple algebraic manipulation of the above equations yields

 

 

 

 

 

 

 

 

 

 

ln

t

 

 

( H 4)(1T/

d

1 / T

m

)

 

 

wash

 

 

 

 

 

 

 

 

 

 

 

 

 

 

t1/2

 

 

 

 

 

R

 

 

 

 

 

 

 

 

 

This expression allows us to calculate kinetic results for particular duplexes once measurements are made at any particular temperature.

KINETICS OF MELTING OF LONG DNA

 

 

 

 

 

 

 

The kinetic behavior of the melting of long DNA is very different from that of short

 

duplexes in several respects. Melting near

 

 

T m

cannot

normally be considered

an all-or-

none reaction. Melting experiments at temperatures far above

 

 

 

 

T m

reveal an interesting

complication. The base pairs break rapidly, and if the

sample

is immediately

returned

 

to temperatures below

T m ,

they

reform

again.

However, if

 

a sufficient time at high

temperature is allowed to elapse, now the renaturation rate

can be many orders of

 

magnitude slower. What is happening is shown schematically in Figure 3.11. Once the

 

DNA base pairs are broken, the two single strands are

still twisted around

each

other.

 

The twist represented by each turn of the double helix must still be unwound. As the

 

untwisting begins, the process greatly slows because the coiled single strands must ro-

 

tate around each other. Experiments show that the untwisting time scales as the square

 

of the length of the DNA. For

molecules

10 kb

or longer, these unwinding

times

can

 

be very long. This is probably one of the reasons why people have trouble performing

 

conventional polymerase chain reaction (PCR)

amplifications

on

high

molecular

 

weight DNA samples. The usual PCR protocols allow only 30 to 60 seconds for DNA

 

 

melting (as discussed in the next chapter.) This is far too short for strand untwisting of

 

large DNAs.

 

 

 

 

 

 

 

 

 

 

Figure 3.11 Kinetic steps in the melting of very high molecular weight DNA.

80

ANALYSIS OF DNA SEQUENCES BY HYBRIDIZATION

 

KINETICS OF DOUBLE-STRAND FORMATION

 

The kinetics of renaturation of short duplexes are simple and straightforward. The reac-

 

tion is essentially all or none; nucleation of

the duplex is the rate-limiting step, and

it

comes about by intermolecular collision of the separated strands. Duplex renaturation ki-

 

netic measurements can be carried out at any temperature sufficiently below the melting

 

temperature so that the reverse process, melting, does not interfere. The situation is much

 

more complicated when longer DNAs are considered. This is illustrated in Figure 3.12

a,

where the absorbance increase upon melting is used to measure the relative amounts of

 

singleand double-stranded structures. If a melted sample, once the strands have physi-

 

cally unwound, is slowly cooled, full recovery of the original duplex can be achieved.

 

However,

depending on the sample, the rate of

duplex formation can be astoundingly

 

slow, as we will see. If the melted sample is cooled too rapidly, the renaturation is not

complete. In fact, as shown in Figure 3.12

b, true renaturation has

not really occurred at

all. Instead, the separated single strands start to fold on themselves to make hairpin duplexes, analogous to the secondary structures seen in RNA. If the temperature falls too rapidly, these structures become stable; they are kinetically trapped and, effectively, can never dissociate to allow formation of the more stable perfect double-stranded duplex.

To achieve effective duplex renaturation, one needs to hold the sample at a temperature where true duplex formation is favored at the expense of hairpin structures so that these rapidly melt, and the unfolded strands are free to continue to search for their correct partners. Alternatively, hairpin formation is kept sufficiently low so that there remains a sufficiently unhindered DNA sequence to allow nucleation of complexes with the correct partners.

Figure 3.12

Melting and renaturation behavior of DNA as a function of the rate

of cooling.

(a)

Fraction of initial base pairing as a function of temperature.

(b)Molecular structures

in DNA after

rapid heating and cooling.

 

 

 

 

 

 

 

 

 

 

 

 

 

 

KINETICS OF DOUBLE-STRAND FORMATION

81

In practice, renaturations are carried out most successfully when the renaturation tem-

 

 

 

 

perature is (

T m 25)°C. In free solution the simple model, shown in Figure 3.5, is suffi-

 

 

cient to account for renaturation kinetics under such conditions. It is identical to the pic-

 

 

 

 

ture used for short duplexes. The rate-limiting step is

 

nucleation, when

two

strands

 

 

 

 

collide in an orientation that allows the first base pair to be formed. After this, rapid du-

 

 

 

 

plex growth occurs in both directions by a

rapid

zippering

 

of the two strands together.

 

 

 

 

This model predicts that the kinetics of renaturation should be a pure second-order reac-

 

 

 

 

tion. Indeed,

they are. If

we call the two strands

 

 

 

 

 

 

 

 

A

and

B we can

describe

the

reaction

as

 

 

 

 

 

 

 

 

 

 

A

 

B

:AB

 

 

 

 

 

 

 

The kinetics will be governed by a second-order rate constant

 

 

 

 

 

 

 

 

 

 

 

k2:

 

 

 

 

 

 

 

 

 

 

d [AB ] k2[A ][B ]

 

 

 

 

 

 

 

 

 

 

 

 

 

 

dt

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

d [A ]

d [B ] k2[A ][B ]

 

 

 

 

 

 

 

 

 

 

dt

 

 

 

 

 

dt

 

 

 

 

 

 

 

In a typical renaturation experiment, the two strands have the same initial concentra-

 

 

 

 

tion: [A ] [B ] C

s

, where

C

s

is 1 the initial total concentration of DNA strands,

 

 

 

o

o

 

 

2

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

[A ] [B

]

 

 

 

 

 

 

 

 

 

 

 

 

 

C s

 

 

 

o

o

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

2

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

At any subsequent time in the reaction, simple mass conservation indicates that the con-

 

 

 

 

centrations of single strands remaining will be [

 

 

 

 

 

 

 

 

 

 

 

 

 

A ] [B

] C,

where

C

is the total instan-

taneous concentration of each strand at time

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

t. This allows us to write the rate equation

 

for the renaturation reaction as in terms of the disappearance of single strands as

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

dC

 

 

 

k2C 2

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

dt

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

We can integrate this directly from the initial concentrations to those at any time

 

 

 

 

 

 

t,

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

t

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

dC

 

k2dt

 

 

 

 

 

 

 

 

 

 

 

C 2

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

0

 

 

 

 

 

 

 

which becomes

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

1

 

 

1

k2t

 

 

 

 

 

 

 

 

 

 

 

 

 

 

C

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

C

s

 

 

 

 

 

 

 

and can be rearranged to

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

C s

 

1

k2tC

s

 

 

 

 

 

 

 

 

 

 

 

 

 

C

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

0 . This is obtainable

82 ANALYSIS OF DNA SEQUENCES BY HYBRIDIZATION

It is convenient to rearrange this further to

 

 

C

 

fs

 

 

1

 

 

 

 

 

 

C

 

(k2tC

s 1)

 

 

s

 

where

fs is the fraction of original single strands remaining.

 

 

 

 

 

 

 

When

fs 0.5, we call the time of the reaction the half time,

 

 

 

 

t1/2 . Thus

 

k2t1/2 C

 

s 1

or

t1/2

 

1

 

 

k2C s

 

 

 

 

 

 

 

 

 

The key result, for a second-order reaction, is that we can adjust the halftime to any value we choose by selecting an appropriate concentration.

The concentration that enters the equations just derived is the concentration of unique, complementary DNA strands. For almost all complex DNA samples, this is not a concen-

tration we can measure directly. Usually the only concentration information readily available about a DNA sample is the total concentration of nucleotides, C

by direct UV absorbance measurements or by less direct but more sensitive fluorescence

measurements on bound dyes. To convert

C

0

 

to

C s , we have to know something about the

sequence of the DNA. Suppose that our DNA was the intact genome of an organism, with

N base pairs, as shown schematically in Figure 3.13. The concentration of each of the two

strands at a fixed nucleotide concentration,

 

C

0 , is just

C

s

 

 

C

0

 

 

2N .

 

 

 

 

 

Now suppose that the genome is divided into chromosomes, or the chromosomes into unique fragments, as shown in Figure 3.13. This does not change the concentration of any of the strands of these fragments. (An important exception occurs if the genome is ran-

domly broken instead of uniquely broken; see Cantor and Schimmel, 1980.) Thus we can

write the product of

k2, the concentration and half-time, as

 

 

 

k2t1/2 C s

 

k2t1/2 C

0

1

 

2N

 

 

 

 

 

 

and this can be rearranged to

C 0 t1/2 2N

k2

Figure 3.13 Schematic illustration of intact genomic DNA, or genomic DNA digested into a set of nonoverlapping fragments.

 

 

 

 

 

 

 

 

KINETICS OF DOUBLE-STRAND FORMATION

83

This key result means that the rate of genomic DNA reassembly depends linearly on the

 

 

 

genome size. Since genome sizes vary across many orders of magnitude, the resulting re-

 

 

 

naturation rates will also. Here we have profound implications for the design and execu-

 

 

 

tion of experiments that use genomic DNA. Molecular biologists have tended to lump the

 

 

 

two variables

C 0 and

t1/2 together, and they talk about the parameter

C

0 t1/2

as “a half cot.”

We will use this term, but don’t try to sleep on it.

 

 

 

 

 

 

 

 

 

 

Some predicted kinetic results for the renaturation of genomic DNA that illustrate the

 

 

renaturation behavior of the different samples are shown in Figure 3.14. These results are

 

 

 

calculated

for

typical

renaturation

conditions

with

a

nucleotide

concentration

of

 

 

1.5 10

4 M. Note

that

under

these conditions

a

3-kb plasmid will renature so

quickly

 

 

that it is almost impossible to work with the separate strands under conditions that allow

 

 

 

renaturation. In contrast, the 3

10

9

bp human genome is predicted to require 58 days to

 

renature under the very same conditions. In practice, this means never; much higher con-

 

 

centrations must be used to achieve the renaturation of total human DNA.

 

 

 

 

 

The actual renaturation results seen for human DNA are very different from the simple

 

 

 

prediction we just made. The reason is a significant fraction of human DNA is highly re-

 

 

 

peated sequences. This changes the renaturation rate of these sequences. The concentra-

 

 

 

tion of strands of a sequence repeated

 

 

m

times in

the genome is

m C

0 /2N

. Thus this se-

quence will renature much faster:

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

C 0 t1/2

2N

 

 

 

 

 

 

 

 

 

 

 

 

mk 2

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

The same equation also handles the case of heterogeneity in which a particular sequence

 

 

 

occurs in only a fraction of the genomic DNA under study. In this case

 

 

 

m 1, and the re-

naturation proceeds slower than expected. Typical renaturation results for human (or any

 

 

 

mammalian)

DNA

are shown

in

Figure

3.15. There

are

three

clearly

resolved kinetic

 

 

Figure 3.14

Renaturation kinetics expected for different DNA samples at a total initial nucleotide

concentration of 1.5

10 4 M at typical salt and temperature

conditions. The fraction of single

strand remaining, f

s , is plotted as a function of the log of the time,

t. Actual experiments conform to

expectations except for the human DNA sample. See Figure 3.15.

Соседние файлы в папке genomics1-10