29.12.2013 Views

Hydrophobic (interaction) chromatography c*). - Bashan Foundation

Hydrophobic (interaction) chromatography c*). - Bashan Foundation

Hydrophobic (interaction) chromatography c*). - Bashan Foundation

SHOW MORE
SHOW LESS

Create successful ePaper yourself

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

BIOCHIMIE, 1978, 60, 1-1_5.<br />

Revue<br />

<strong>Hydrophobic</strong> (<strong>interaction</strong>) <strong>chromatography</strong> <strong>c*</strong>).<br />

J.-L. OCHOA * *<br />

lnstitnlo de Quimica.<br />

Universidad Nacional Autonoma<br />

de Mexico (UNAM),<br />

Ciudad Universitaria,<br />

Mexico, D.F.<br />

Introduction.<br />

The isolation and purification of macromolecules<br />

by biochemical fractionation techniques,<br />

like ion-exchange <strong>chromatography</strong>, gel filtration<br />

(molecular-sieve <strong>chromatography</strong>), affinity chro-<br />

Inatography, electrophoresis, etc., is primarily<br />

dependent on their biological and physicochcmical<br />

properties [1-5].<br />

Based on biospecific <strong>interaction</strong>s [6], affinity<br />

<strong>chromatography</strong> has been considered as one of<br />

the most effective separating methods. However,<br />

serious disadvantages are found upon its application,<br />

due mainly to undesirable non-biospecific<br />

adsorption [7] attributed to the characteristics of<br />

the matrix ov support, and the nature of the<br />

ligand and spacer-arm introduced to b~idge the<br />

ligand from the matrix backbone. Moreover, once<br />

a protein is attached to an immobilized ligand,<br />

for which it sho~vs affinity, the properties of the<br />

support may change and become like those of an<br />

ion-exchanger, interfering with the chromatographic<br />

process.<br />

The belief that such interferences could be<br />

adequately controlled by eliminating ionic groups<br />

in the matrix and/or in the spacer-arm, brought<br />

as a consequence other types of undesirable<br />

effects closely related to the extent of hydrophobicity<br />

of the spacer-arm employed [57]. Yon [93,<br />

and Er-eI et aL [10], reported that by coupling<br />

different types of spacer-arms (varying in hydrophobicity)<br />

to an inert matrix, potent adsorbents<br />

for proteins were obtained. The adsorption mechanism<br />

seems to be based, fundamentally, on<br />

hydrophobic <strong>interaction</strong>s between the protein and<br />

the adsorbent, and it has been demonstrated that<br />

it can be positively exploited for separation purposes<br />

[11, 105]. In this way, a novel technique<br />

which takes advantage of the hydrophobicity of<br />

* This article is a complementary part to my contribution<br />

at the Forum des Jeunes, in Lyon, France<br />

(July 6-8, •977).<br />

** Present address : Facultd de Mddecine de Strasbourg,<br />

Institut de Clfimie Biologique, 11 rue<br />

Humann, 67000 Strasbourg, France.<br />

biomolecules, has been introduced as a complement<br />

to the other, routinely employed, separating<br />

methods mentioned above.<br />

Since the separation is based on hydrophobic<br />

<strong>interaction</strong>s, this technique receives a number of<br />

names 'which intend to describe the principle and<br />

the parameters involved in the separation process,<br />

though all of them are entirely or partly dealing<br />

with the concept of hydrophobicity :<br />

<strong>Hydrophobic</strong> <strong>chromatography</strong> [11] ; <strong>Hydrophobic</strong><br />

(<strong>interaction</strong>) <strong>chromatography</strong> [51]; <strong>Hydrophobic</strong><br />

salting-out <strong>chromatography</strong> [93]; Phosphate-induced<br />

protein <strong>chromatography</strong> [54] ; Repulsion<br />

controlled <strong>chromatography</strong> [94] ; Detergent<br />

protein-<strong>interaction</strong> [95, 961 ; <strong>Hydrophobic</strong><br />

affinity <strong>chromatography</strong> [128] ; etc.<br />

It has been shown [12] that a large part of the<br />

non-polar residues of the amino acids in proteins<br />

are exposed to ~vater interface, as opposed to the<br />

expected preferential location of the hydrophobic<br />

amino acids in the interior of the biomolecule.<br />

These non-polar amino acids are found in the<br />

protein surface forming ) of distinct<br />

hydrophobic character which Can account for the<br />

biospecific conformation of the protein [13] as<br />

well as for its ability to complex or aggregate to<br />

other types of molecules for instance, lipids. Presumably,<br />

these hydrophobic (< patches >> are randomly<br />

distributed on the surface of the biomolecule.<br />

Their number (possible number of interacting<br />

sites) and the extent of their hydrophobicity<br />

(type and distribution of the non-polar amino<br />

acids) should be a characteristic of each macromolecule.<br />

Therefore, their specific separation<br />

should be possible with an adequate hydrophobically<br />

coated support or matrix.<br />

I. THE BIOLOGICAL ROLE OF THE HYDROPHOBICITY<br />

IN BIOMOLECULES.<br />

There is no doubt that the hydrophobie <strong>interaction</strong>s<br />

play an important role in biological systems.<br />

The membranes in the living cell are made<br />

1


- - An<br />

2 J.-L. Ochoa.<br />

up of mainly hydrophobically interacting lipidlipids<br />

and lipid-proteins. The sub-units of large<br />

proteins are often held together by hydrophobic<br />

bonds, and it is well accepted that they are important<br />

in supporting the tertiary protein structure<br />

[13].<br />

Functionally, hydrophobic <strong>interaction</strong>s seem to<br />

be involved in recognition processes, as for example<br />

in the case of some enzyme-substrate complexes<br />

and antigen-antibodies associations, etc.<br />

[14-15, 27-30]. Except for the case of membranes,<br />

in which an exclusive structural function has<br />

been attributed [31], there is not much work in<br />

correlating the hydrophobic properties of the<br />

biomolecules with their functions [32, 99-100].<br />

Therefore, it could be interesting to find out<br />

whether these hydrophobic regions in the biomolecules<br />

are related to their intrinsic biological<br />

properties, and/or to their location in the cell.<br />

Perhaps one of the reasons why fe"w systematic<br />

studies of hydrophobic effects have been made,<br />

is that the hydrophobic compounds possess lo"w<br />

solubility in water. If they are made soluble by<br />

the introduction of polar groups, they tend to<br />

form micelles as exemplified by the case of soaps.<br />

Although molecular dispersion may be obtained<br />

through the addition of polarity reducing agents<br />

to the medium, such agents would also reduce the<br />

<strong>interaction</strong> of the hydrophobic compounds with<br />

proteins and even alter the protein structure,<br />

which generally depends on the integrity of the<br />

hydrophobic core [13].<br />

A means to obtain molecular dispersion of a<br />

hydrophobic ligand in aqueous milieu, without the<br />

addition of polarity reducing agents, is to attach<br />

the ligand to a hydrophilic but insoluble polymer<br />

such as agarose. Consequently, hydrophobic supports<br />

can be used not only for separation purposes,<br />

but also to study the hydrophobicity of the<br />

biomolecules and their modes of <strong>interaction</strong>. In<br />

turn, this may lead to a clearer understanding of<br />

their functions.<br />

II. THE PHYSICOCHEMICAL CONCEPT<br />

OF HYDROPHOBICITY.<br />

Perrin [16] `was the first to utilize the terms<br />

> and ¢ hydrophobic >> when studying<br />

colloids. He considered them hydrophilic<br />

if their stability was relatively insensitive to the<br />

addition of electrolytes, or hydrophobic if they<br />

exhibit extreme sensitivity to added electrolytes.<br />

Later, Langmuir [17], in a classical experiment,<br />

discussed polar molecules (fatty acids) in terms of<br />

BIOCHIMIE, 1978, 60, n ° 1.<br />

hydrophilic and hydrophobic groups. Since then,<br />

these terms have been in every day use.<br />

In general, "when an apolar group (or hydrophobic)<br />

is inserted into `water, several effects can be<br />

observed (for review see ref. 18) :<br />

-- A negative unitary entropy change (--AS),<br />

`which implies an overall increase in the degree<br />

of order. The entropy distribution becomes relatively<br />

important as the molecular weight of the<br />

apolar group increases.<br />

--A small enthalpy change, usually negative<br />

(--AH), which reflects an energetic component<br />

in the <strong>interaction</strong> between 'water and an apolar<br />

group and opposes the negative unitary effect<br />

favouring mixing of the hydrocarbon in water.<br />

increase in heat capacity (+ ACp), which<br />

implies either an increased degree of freedom of<br />

the vibrational and rotational motion within the<br />

existing structure, or a progressive alteration of<br />

an existing structure as the temperature is raised.<br />

-- A decrease in volume (--AV), which indicates<br />

that the apolar groups and the water molecules<br />

have packed together with a contraction of<br />

some structure, or that the apolar groups are<br />

accomodated into open spaces or voids preexisting<br />

`within the water structure itself. The former<br />

is unlikely in view of the small enthalpy change<br />

involved.<br />

--Finally, the presence of apolar groups in<br />

water results in an increase in the number<br />

or strength of hydrogen bonds within the water<br />

molecules, as has been demonstrated by spin lattice<br />

relaxation studies and further supported by<br />

Raman measurements [60]. Its probable reason is<br />

that the hydrocarbon chains restrict the mobility<br />

of water molecules in such a `way that the covalent<br />

character of the hydrogen bound is increased<br />

[92].<br />

On the other hand, the molecular picture of the<br />

hydrophobic <strong>interaction</strong> is the reverse of that obtained<br />

when introducing apolar groups into water<br />

(see mechanism below). The withdrawal of the<br />

apolar groups from the aqueous phase removes<br />

restrictions on hydrogen-bond bending and thus,<br />

achieves a positive unitary entropy. This supports<br />

the observation that the hydrophobic <strong>interaction</strong><br />

is a spontaneous process. This assumption was<br />

substantially demonstrated by Frank and Evans<br />

in 1945 [20]. Some years later, it has been shown<br />

[19] that the free energy values of the transfer of<br />

aliphatic hydrocarbons from an apolar medium<br />

to a polar one, like water, increases linearly with<br />

increasing numbers of (CH 2) methylene groups,<br />

and thus the hydrophobicity of the molecule.


<strong>Hydrophobic</strong> (<strong>interaction</strong>) <strong>chromatography</strong>. 3<br />

III. DETERMINATION OF THE HYDROPHOBICITY OF<br />

BIOMOLECULES.<br />

It is well kno~vn that most proteins contain a<br />

relatively high proportion of amino acids ~dth<br />

non-polar chains (table I). Tanford [21], and<br />

TABLE I.<br />

Non-polar (hydrophobic) amino acids commonly<br />

found in proteins.<br />

AI anine CH 3 -R<br />

Leucine<br />

Isoleucine<br />

Valine<br />

Proline<br />

Phenylalanine<br />

Tryptophane<br />

Methionine<br />

CH3x<br />

/ CH-CH2-R<br />

CH 3<br />

CH3-CH2-CH-R<br />

I<br />

CH 3<br />

CH3 x<br />

CH-R<br />

CH3 /<br />

CH CH2\ CH_CO0tl<br />

CH , /<br />

2XNH<br />

d CH:CH\<br />

CH ,,,CH-CH _ -R<br />

"~ CH-CH >" z<br />

/1 CH-CH \\<br />

C -- CH-CH ~-R<br />

CH % CH=C ~ Ixrfl "~CI-I "<br />

CH3-S-CH2-CH2-R<br />

R<br />

-CH-CO0-<br />

!<br />

+NH 3<br />

Nozaki and Tanford [22], formulated an experimental<br />

procedure based more or less on the Kauzmann<br />

[13] conception which permits the estimation<br />

of the amino acid hydrophobicity. The method<br />

consisted, essentially, in determining the<br />

solubilities of the amino acids in water as ~vell<br />

as in progressively increasing concentration of<br />

some organic solvents, such as ethanol in water.<br />

The solubilities of the amino acids were extrapolated<br />

to pure organic solvents and then the free<br />

BIOCHIMIE, 1978, 60, n ° 1.<br />

energy value of the transfer for the amino acid<br />

from pure organic solvent to ~vater was calculated.<br />

Using as a reference glycine, and substracring<br />

its free energy transfer value from that of all<br />

the other amino acids, it was possible to formulate<br />

a hydrophobicity scale for amino acid residues<br />

~vhere their free energy transfer values<br />

become more and more positive as the hydrophobic<br />

character of the compounds increases [18].<br />

Other attempts in scaling the amino acid hydrophobicity<br />

can be illustrated by the ~vorks of Bull<br />

and Breese [23] and Bigelov¢ [24]. The former<br />

studied the effect of the amino acid on the surface<br />

tension of water and the latter calculated the<br />

average hydrophobicity of several proteins by<br />

their amino acid composition according to data<br />

of Tanford and of Bull and Breese.<br />

A different approach ~vas done in terms of the<br />

frequency of non-polar side chains in proteins<br />

[2~], in which yeas found a variation from 0.21 to<br />

0.47. Separately, Fisher [26] employed the ratio<br />

of the volume of the polar groups to those of the<br />

non-polar as a hydrophobicity degree, but this<br />

idea has the inconvenience, like the one mentioned<br />

above [25], to consider that a group is<br />

either polar or non-polar, ~without any gradation<br />

between these t~o extremes.<br />

Recently, a method for studying the magnitude<br />

of the <strong>interaction</strong> bet~veen protein and aliphatic<br />

hydrocarbon chains, and thus indirectly the hydrophobicity<br />

, ~vas reported [~0]. It is based on<br />

the partition of proteins in an aqueous two phase<br />

system containing dextran and polyethylene glycol<br />

and different:fatty esters of polyethylene glycol.<br />

However, the measurements depend largely<br />

on a critical chain length which, by this technique,<br />

should be greater than 8 carbons.<br />

Finally, an approach to localize the hydrophobic<br />

sites on the surface of the proteins by means<br />

of interacting with small molecules, has been<br />

attempted using fluorescent probes [27]. It is not<br />

difficult to speculate that ~vith this set of information,<br />

a better comprehension of many biological<br />

phenomena is near. We do not know yet<br />

how the complex enzymatic systems are organized,<br />

nor how the recognition bet~veen proteins<br />

occurs to constitute such enzymatic complexes<br />

after protein synthesis. Neither do we have a good<br />

explanation for the transport of many substances,<br />

including proteins, through the hydrophobic core<br />

of the membrane. And furthermore, it is possible<br />

to believe that hydrophobic ((patches >> in the<br />

membrane (either out or inside) are needed to<br />

make possible many of the most common biolo-


4 J.-L. Ochoa.<br />

gical phenomena like the recognition of non-polar<br />

substances by membrane receptors and even cellcell<br />

<strong>interaction</strong>s.<br />

IV. THE MECHANISM OF THE HYDROPHOBIC INTER-<br />

ACTION.<br />

The hydrophobic <strong>interaction</strong> is the result of the<br />

adherence of two non-polar groups. The case of<br />

detergents can be considered as an example where<br />

negative enthalpy changes are observed during<br />

the micelle formation in aqueous solvents [33!.<br />

If the adsorption of proteins to hydrophobic matrices<br />

is considered to be a process of limited micelle<br />

formation, a negative change in the enthalpy<br />

value ~vould not preclude the hydrophobic nature<br />

of the binding. In addition, this change must be<br />

negligible as compared to the value of increasing<br />

entropy of the system I20]. Furthermore, before<br />

the <strong>interaction</strong> the water molecules are forced to<br />

keep in orde~ around the hydrophobic entities<br />

(fig. 1) as compared to the order in the bulk. When<br />

the hydrophobic sticks come in contact with each<br />

ooooooooooo<br />

o o o<br />

oeOOooeeoeooo o~ o • •<br />

o o o o<br />

::~ii........oj_<br />

oooooooooooo<br />

•<br />

ooooooooooo O O0000QeO<br />

0 0 0 000 O • 0<br />

D •<br />

a) b)<br />

F[~. 1. -- The mechanism of hydrophobic <strong>interaction</strong>.<br />

The water molecules around the hydrophobie<br />

> are doted, for simplicity, in order to<br />

distinguish them from those in the bul~. Notice. that<br />

the difference bet'ween a) and b) is only the improved<br />

degree of disorder (-I-AS) of vcater molecules.<br />

other, the ordered water molecules will be excluded<br />

and 'will adopt the less ordered buIk water<br />

state which is equivalent to an increase in entropy.<br />

This hypothesis is supported experimentally<br />

by the study of antigen-antibody complex formation<br />

[27] where a relative insensitivity to dissolution<br />

of a preformed antigen-antibody precipitate<br />

is observed, suggesting that once the complex is<br />

formed, the solvent is largely excluded in lhe<br />

regions of contact.<br />

Thermodynamically, the free energy value AG<br />

of a hydrophobic <strong>interaction</strong> is a function of AH<br />

and AS, according to equation :<br />

AG ~ AH -- TAS ..................... (N ° 1)<br />

Since AH is small as compared to TAS value,<br />

the process is fundamentally determined by the<br />

BIOCHIMIE, 1978, 60, n ° 1.<br />

change in entropy. In these conditions the reaction<br />

proceeds spontaneously [51]. In other words,<br />

the input of energy or chemical work is not necessary<br />

to make possible the <strong>interaction</strong> between two<br />

hydrophobic molecules in aqueous solutions.<br />

The contact between two different molecules,<br />

like the sub-units in the case of some proteins,<br />

is largely dependent on the surface areas occupied<br />

by the residues which participate in the <strong>interaction</strong>.<br />

The concept of accessible surface area [80j<br />

describes the extent to ~vhich protein atoms can<br />

form contacts with water, and is related to hydrophobic<br />

free energies [81]. In any case, the association<br />

of protein sub-units, whether by van der<br />

Waals contacts, electrostatic forces and/or hydrophobic<br />

<strong>interaction</strong>s, leads to a reduction of the<br />

surface area accessible to the solvent ~vhen the<br />

two molecules associate. Evidence against the<br />

hydrogen bond as a major contribution to the<br />

free energy of the protein-protein <strong>interaction</strong> has<br />

been obtained lhermodynamically [82]. On the<br />

other hand, van der Waals <strong>interaction</strong>s, though<br />

they are more numerous as they involve all the<br />

pair of neighbouring atoms, are much less energetic<br />

and their overall contribution is small. The<br />

hydrophobic contribution is largely dependent<br />

upon the entropy gained by water due to the<br />

smaller accessible protein surface area when the<br />

protein molecules form a complex. The hydrophobic<br />

<strong>interaction</strong> seems to be entirely unspecific<br />

as compared to the complementarity of the surfaces<br />

involving hydrogen bonds and van der<br />

Waals contacts ; ho~,ever, they decide which proteins<br />

can recognize each other.<br />

V. THE SPECIFICITY OF THE ADSORPTION<br />

OF BIOMOLECULES ON HYDROPHOBIC SUPPORTS.<br />

Biospecific affinity, whether involving an<br />

> or not, presumably depends to a<br />

large extent on the complementarity of the contours<br />

of the interacting molecules. In the absence<br />

of any specific effect, the binding of proteins to<br />

substituted agaroses is greatly affected by the<br />

overall charge of the biomolecule [105].<br />

The lack of activity in the adsorbed state of<br />

some proteolytic enzymes [90], indicates that the<br />

binding site is occupied by the hydrophobic<br />

group of the substituted agarose. Whether the<br />

hydrophobic group interacts with the hydrophobic<br />

poc1~ets of the active sites of the enzymes or<br />

acts by a less specific ~way can be questioned. In<br />

any event, both cases may contribute to the<br />

adsorption phenomena.


<strong>Hydrophobic</strong> (<strong>interaction</strong>) <strong>chromatography</strong>.<br />

In other examples, the addition of the specific<br />

substrate accompanying the enzyme during the<br />

<strong>chromatography</strong> (to mask the specific binding<br />

sites) did not alter the binding of the enzyme to<br />

the hydrophobic matrix, indicating that the<br />

adsorption takes place through sites other than<br />

the specific substrate binding site [971.<br />

It is interesting to note that glutamine synthetase<br />

and other three proteins involved in the regulation<br />

of glutamine metabolism are all retained by<br />

the same amino-aIkyl agarose derivatives [99~ (see<br />

table II). Although, as Shaltiel and coworkers<br />

have signaled, this could be fortuitous, it might<br />

reflect a mutual hiospecific affinity among these<br />

proteins since they must interact with each other<br />

in order to effect their regulatory functions in the<br />

highly integrated glutamine synthetase cascade<br />

system.<br />

The models proposed by Shaltiel [11] and<br />

Jennissen [78] explain in a different manner the<br />

mechanism by which the <strong>interaction</strong> between the<br />

protein and the adsorbent occurs. For neutral<br />

supports, Jennissen [86~ has presented evidence<br />

in favor of the idea that the adsorption of proteins<br />

to aIkyl-agarose derivatives takes place at a<br />

critical alkyl group density and is a function of<br />

its hydrophobicity. In other ~vords, this hypothesis<br />

considers that the protein needs to present<br />

multiple attachment points in order to be adsorbed<br />

by a determinate member of the series of<br />

alkyl-agarose derivatives. This is in agreement<br />

with an earlier observation made by Hjert6n el<br />

al. [521. Shaltiel Ill on the other hand, suggests<br />

that the adsorption of the protein is due to the<br />

inter.action of an allkyl residue of specific length<br />

(¢ yard-stick ~) with a hydrophobic pocket of the<br />

protein. Evidently, and as compared to Jennissen's<br />

model, the tatter implies a more specific mechanism.<br />

Ho~vever, it has been demonstrated [583 that<br />

at high chain length, when the <strong>interaction</strong>s are<br />

stronger, the binding is also less specific.<br />

The [841<br />

through the multivalent binding, may explain the<br />

increased free energy value of adsorption of a<br />

determinate member of the alkyl-agarose series.<br />

That is, a given protein may be adsorbed by a<br />

certain member of agarose-substituted series only<br />

if the number of ¢ contacts ~> is big enough to<br />

effect its retention, and this can be done by variation<br />

of the degree of substitution. Comparatively,<br />

amino-alkyl agaroses present a > effect to,wards the adsorption of phosphorylase<br />

b, possibly due to a different mechanism of<br />

<strong>interaction</strong> 'which does not necessarily exclude<br />

the multivalent attachment E84].<br />

BIOCHIMIE, 1978, 60, n ° 1.<br />

It has been observed, in the case of some hydrophobic<br />

aIkyl- and alkylamino supports, that certain<br />

> binding sites are occupied first,<br />

and that others of decreasing affinity become<br />

occupied when more protein is fed to the column.<br />

This is shown by the fact that only part of the<br />

protein can be eluted by increasing the salt concentration,<br />

whereas the remained is dislodged by<br />

the addition to the eluant of a polarity reducing<br />

{a)<br />

(c)<br />

(b)<br />

(d)<br />

Fro. 2. -- Models of adsorption o[ proteins on hydrophobic<br />

matrices : a} The model suggested by Shaltiel<br />

[11] ; b) and c) Represents the two possibilities of multi-attachment<br />

adsorption : on the matrix surface (b)<br />

and on a cavity in the matrix (c) ; d) The irregularity<br />

of the surface (attributed to the nature of the agarose<br />

structure) is the probable cause for 'which binding<br />

sites are not identical, and consequently the forces<br />

involved in the binding are different.<br />

agent such as ethylene-glycol. It should be emphasized<br />

that this problem is mostly related to those<br />

hydrophobic supports which possess both hydrophobic<br />

and ionic groups, resulting from the coupling<br />

[70, 471 procedure or the nature of the<br />

ligand [96], like in the case of amino-alkyl derivatives.<br />

This apparent inhomogeneity of both<br />

matrix and protein can lead to possible wrong<br />

interpretations about the purity and characteristics<br />

of many proteins [98, 105]. For instance,<br />

the recovery of rechromatographed materials is<br />

improved up to 95 per cent, as compared to that<br />

of 70 per cent obtained when the crude extract<br />

is applied into the column [98]. In other cases,<br />

desorption of purified material requires the variation<br />

of the eluant, as pointed above. It seems


6 J.-L. Ochoa.<br />

TABLE II.<br />

<strong>Hydrophobic</strong> matrices.<br />

Type Coupling procedure Bepresentation References<br />

I. Agarose derivatives<br />

1. Un-substituted<br />

2. Amino-alkyl-<br />

3, Alkyl-<br />

4. Amino acid<br />

5. Other derivatives<br />

Fatty acids<br />

AcetyI-N-amino-alkyl<br />

Aniline<br />

Benzyl-ether<br />

Phcnyl-N-butyl-amine-<br />

NAD +-<br />

Hydroxy- alky 1<br />

Diverse aromatic-alkyl<br />

derivatives<br />

BrCN<br />

BrCN<br />

glycidyl ether<br />

BrCN<br />

BrCN (azide)<br />

BrCN (cvrbodiimide)<br />

(acetylation of aminoalkyl-(A))<br />

BrCN and glycidyl ether<br />

glycidyl ether<br />

(A)<br />

(A)-CH-N H-(CH~)n-N H,_,<br />

I<br />

NH.~<br />

+<br />

(A)-CH-NH-( CH.~)-CH:~<br />

I<br />

NH~<br />

+<br />

(A)-O-CH~-CH-CH~-O-<br />

[ (CH'2)n-CH3<br />

OH<br />

(A)-glycine<br />

(A)-valine<br />

(A)-leucine<br />

(A)-tyrosine<br />

(A)-phenylalanine<br />

(A)-tryptophane<br />

(A)-(CHs)n-OH<br />

41, 47, li3]<br />

[9-11, 94, li3, Ill,<br />

li4, 58]<br />

[II,94, I13, li4]<br />

[52, li2l<br />

[97, I15]<br />

[54, 68, 97, li6, 77]<br />

[77, 97]<br />

[97]<br />

[35, 40, 42]<br />

[I051<br />

llo6]<br />

179, 58]<br />

[57, 94, li7]<br />

[40]<br />

[93, 90]<br />

[105l<br />

[57, 58]<br />

[58, li2]<br />

[I12, 55]<br />

II. Dextran derivatives<br />

(Sephadex)<br />

(s)<br />

1. Acetyl-Sephadex<br />

2. Methyl ether-(S)<br />

3. Hydroxy-propyl- ether-<br />

(S)<br />

4. Hydroxy-alkyL(S)<br />

5. LH-20-(S)<br />

acetylation<br />

acetylation<br />

from commercial sources<br />

CH3-CO-(S)<br />

CH3-O-(S)<br />

HO-CHs-CHs-CH.2-O-(S)<br />

HO-(CH~)n-O-(S)<br />

H3ol<br />

[91]<br />

[9t]<br />

[91]<br />

[9i]<br />

III. Cellulose derivatives<br />

(C)<br />

1. Diethyl amino ethyl<br />

cellulose<br />

2. Carboxy methyl-(C)<br />

3. Benzoylated-DEAE-(C)<br />

4. Esters of alkyl and aryl-<br />

(C) (paper, cotton)<br />

from commercial sources<br />

from commercial sources<br />

benzoyl chloride<br />

phenoxy acetylation<br />

DEAE-(C)<br />

CM-(C)<br />

BD-(C)<br />

[47, 71]<br />

[47, 71]<br />

[I03]<br />

[101]<br />

IV. Glass derivatives<br />

(G)<br />

1. Alkyl-silated-(G)<br />

2. Propyl_lipoamide-(G)<br />

BIOCHIMIE, 1978, 60, n ° 1.<br />

amino-silanization and<br />

amide bond formation<br />

with the corresponding<br />

alkyl chloride<br />

[ii8]<br />

[i07]


<strong>Hydrophobic</strong> (inleraction) <strong>chromatography</strong>.<br />

TABLE II.<br />

Type Coupling procedure Representation Relerences<br />

V. Others supports<br />

1. Polyamino methylstyrene<br />

(polystyrene,<br />

polyamide, Dowex<br />

I-X8)<br />

2. Alkyl-amine of polyacrylic<br />

resins (butyl,<br />

capryl, lauryl, palmityl,<br />

steatyl, oleyl, linoleyl)<br />

from commercial sources<br />

17t1<br />

chlorination for amide [761<br />

bond formation<br />

that this discrepancy depends on the hydrophobicity<br />

of the protein and the adsorbent, and moreover,<br />

that the binding sites are not identical in<br />

a given carrier (which is true for most of the<br />

adsorbents employed in <strong>chromatography</strong> in general).<br />

The multiple point attachment binding is most<br />

likely to occur in a cavity of the adsorbent than<br />

on its surface, particularly when the protein molecule<br />

fits into the cavity (fig. 2). Conversely, at<br />

points on the matrix ~vhere the bound ligands are<br />

distributed over a protruding area, binding would<br />

he less strong. Since the surface of the adsorbent<br />

may be assumed to be irregular, many different<br />

situations in addition to these two hypothetical<br />

cases ~vould be obtained. Additionally, small<br />

amounts of protein bind more homogeneously on<br />

a particular column than a large amount. This<br />

could mean that by reducing the amount of<br />

applied protein, the stronger binding sites might<br />

predominate in the binding.<br />

If multiple point attachment were one of the<br />

reasons for strong non-specific adsorption, one<br />

~vay to reduce it ~vould be to louver the degree of<br />

substitution to the point where the distance<br />

between the substitution groups is larger than the<br />

diameter of the protein molecule. This ~vould not<br />

affect the specific ¢ one-to-one >> <strong>interaction</strong> as<br />

that between an enzyme active site and an immobilized<br />

subslrate analogue.<br />

VI. THE FACTORS INVOLVED IN HYDROPHOBIC<br />

(INTERACTION) CHROMATOGRAPHY.<br />

A. The matrix hydrophobicity.<br />

1. Matrices and supports.<br />

Proteins differ in their hydrophobicity as a<br />

function of their primary structure, that is, in the<br />

BIOCHIMIE, 1978, 60, n ° 1.<br />

sequence and amino acid composition. Hence, it<br />

is not surprising that the first type of hydrophobic<br />

supports ever employed were derivated from<br />

coupling various non-polar amino acids to an<br />

inert support or matrix like agarose [54]. Other<br />

supports have also been employed: cellulose,<br />

glass, dextran, etc. (see table II), but the ideal<br />

matrix 'without secondary non-biospecific adsorption<br />

effects has not been found.<br />

All the different types of supports and matrices<br />

utilized at the present possess undesirable interferences<br />

attributed to their chemical composition.<br />

Additional effects arc obtained in such away that<br />

is difficult to obtain a pure hydrophobic <strong>interaction</strong><br />

<strong>chromatography</strong> after the introduction of the<br />

spacer-arm or hydrophobic ligand.<br />

In table II, the various types of matrices<br />

employed in HIC, as well as the different types<br />

of ligands coupled, have been summarized. References<br />

are given to illustrate specific applications.<br />

As has been demonstrated, un-substituted agarose<br />

is sufficiently non-polar presumably due to<br />

the 3-6 methylene diether bridges present in every<br />

second galactose residue of the polysaccharide<br />

chain [67] with respect to lhe retention of halophilie<br />

proteins [47] and nucleic acids [41, 119], at<br />

high salt concentrations (2.5 M ammonium sulfate).<br />

Though the halophilic proteins are highly<br />

negatively charged [48], as a result of an excess<br />

of acidic groups [49], it is believed that the high<br />

salt concentration eliminates the possible ionic<br />

<strong>interaction</strong>s.<br />

When DEAE- and CM-derivatives, either of<br />

cellulose or agarose, have been used, the elution<br />

pattern shifts to higher or louver salt concentrations<br />

than those required 'when tile neutral derivative<br />

is used as a support. This is explained by


8 J.-L. Ochoa.<br />

electrostatic attraction-repulsion forces between<br />

the negatively charged haloproteins and the nature<br />

of the charge on the matrices.<br />

With few exceptions, agarose is the type of<br />

matrix most commonly used (table II), and the<br />

kind of ligands or spacer-arm introduced are in<br />

a wide range of hydrophobicities and structures.<br />

Consequently, the hydrophobicity of the nmtrix<br />

depends on the type of the ligand E78], and on the<br />

degree of substitution [5,6, 86]. In this respect, the<br />

properties of the ligand should be discussed first.<br />

2. The role of the ligand in the hydrophobicity<br />

of the matrix.<br />

As pointed out above, HIC was born (in many<br />

cases) as a consequence of non-biospecific adsorption<br />

found in affinity <strong>chromatography</strong> systems<br />

and attributed to the type of spacer-arm used to<br />

bridge the ligand to the matrix. Nevertheless, evidences<br />

exist that prove that HIC was conceived<br />

by Hjerl6n and his group in a different way,<br />

while studying the solubilization of membrane<br />

proteins. They attempted their separation using<br />

a hydrophobic support to 'which a detergent<br />

(SDS) was coupled. In this case, however, the<br />

adsorption was so strong that the proteins could<br />

not be desorbed by any non-denaturing system<br />

(unpublished results, personal communication).<br />

When Cuatrecasas [87] sho~,ed that ~-galactosidase<br />

could only be fractionated if the ligand was<br />

sufficiently separated from the matrix, and thus<br />

avoiding sterical hindrance, the idea of separating<br />

ligands from the matrix backbone by use of<br />

spacer-arms rapidly spread. Very soon, it became<br />

clear, that many proteins were adsorbed nonspecifically<br />

because the presence of the arm introduced<br />

non-biospecific adsorption centers. For<br />

instance, a spacer-arm carrying charged groups<br />

gives origin to electrostatic <strong>interaction</strong>s [106]. The<br />

substitution of the charged arms by other neutral<br />

chemical analogs was thought to be an excellent<br />

alternative but then, other type of non-specific<br />

<strong>interaction</strong>s, hydrophobic in nature, turned out<br />

to be important [7].<br />

Yon [9], Er-el et al. [10] found that protein<br />

could bind substituted agarose matrices without<br />

any specific ligand. The adsorption ~ras attributed<br />

to the presence of the diamino-alkyl group<br />

employed as a spacer. This observations were<br />

further supported by O'Carra and coworkers [57,<br />

88], Who found that the presence of the ligand<br />

did not al'ways account for the ratardation effect<br />

and again the spacer-arm was considered as<br />

responsible. Yon [9] suggested that it was possible<br />

to take advantage of such non-biospecific adsorp-<br />

BIOCHIMIE, 1978, 60, n ° 1.<br />

tions and showed that certain proteins could be<br />

selectively adsorbed on decyl-agarose derivatives<br />

differing in the ionic character of the spacer-arm.<br />

Later, Shaltiel and his group [10, 11] reported<br />

purification of several enzymes on a series of<br />

alkyl and amino-alkyl agarose derivatives. Almost<br />

simultaneously, Hjert6n and his group [521 reported,<br />

for the first time, the preparation of different<br />

substituted agaroses 'with uncharged groups<br />

that could illustrate an exclusive hydrophobic<br />

mechanism of adsorption.<br />

The effect of the hydrophobicity of the spacerarm<br />

in the purification of various proteins [8-11]<br />

has been possible thanks to the development of<br />

the (< mock affinity systems >>. Steers [87 demonstrated<br />

that the adsorption of ~-galactosidase<br />

was strongly related to the length of the spacerarm.<br />

The series of aIkyl and amino-alkyl derivatives<br />

of agarose prepared by Shaltiel have also<br />

been successfully applied in many cases [11].<br />

The difference of a single C-atom in agarose<br />

bound N-aIkyl groups may have a large effect in<br />

the hydrophobic binding of a particular protein<br />

[75]. Therefore intermediary hydrophobic compounds<br />

between two consecutive members of a<br />

homologous series of ligauds are needed. Such<br />

intermediate hydrophobicities can be obtained,<br />

for instance, through the introduction of a<br />

charged or other hydrophilic group, e.g., hydroxyl<br />

group. The introduction of double bonds or<br />

of branching of the chain, reduces also the hydrophobicity<br />

as compared to the corresponding saturated<br />

straight chains [13]. Weiss and Bucher<br />

['76] have prepared some alkyl agarose derivatives<br />

of increasing insaturation for this purpose. In<br />

addition aromatic derivatives may possess intermediate<br />

hydrophobicities between t~wo consecutive<br />

members of the alkyl-agarose series. Benzene,<br />

for example, is equivalent to that of 3-4 straight<br />

chain hydrocarbon [19].<br />

Although Tanford [1O] has showed that the<br />

hydrophobicity of a linear aliphatic carbon<br />

increases linearly with increasing number of CH 2-<br />

groups, Shanbhag [50] considers that the effective<br />

hydrophobicity depends also on the flexibility<br />

of the hydrocarbon chain and consequently<br />

on the degree of <strong>interaction</strong> within such chains,<br />

specially for long chains. Hofstee [75], on the<br />

other hand, assures that neither the difference in<br />

molecular shape of the N-alkyl ligands and the<br />

side chain of aromatic compounds like phenylalanine<br />

or tryptophan, nor the difference in net<br />

charge of the adsorbent is a determining factor<br />

for protein binding. It seems then, that the me-


<strong>Hydrophobic</strong> (<strong>interaction</strong>) <strong>chromatography</strong>.<br />

chanism of the adsorption of proteins to hydrophobic<br />

matrices follows some very complex rules.<br />

This problem can be exemplified by the case of<br />

the adsorption of erythrocytes on a series of<br />

aIkyl-agarose derivatives [110] in ~which case a<br />

decrease in adsorption bet'ween C6-C s is noticed<br />

to occur without a reasonable explanation.<br />

The combination of the principles of ionicexchange<br />

<strong>chromatography</strong> and HIC has been successfully<br />

applied in some cases [9, 41]. The use of<br />

ligands containing both ionic and hydrophobic<br />

characters have been recommended in order to<br />

avoid the denaturing effect detected in the use of<br />

alkyl-agaroses [108]. It was also argued that the<br />

elution procedure might be milder than in the<br />

case of pure hydrophobic spacer-arms or ligands<br />

[9] and furthermore, that the use of charged<br />

arms was capable of finer discrimination between<br />

lipophilic proteins. The protein, in this case,<br />

interacts hydrophobically, involving the alkyl<br />

chains, and electrostatically, involving the terminal<br />

ionic groups. With this type of adsorbents,<br />

if the <strong>chromatography</strong> is carried out at a pH<br />

equivalent to the isoelectric point (IP) of the protein,<br />

the adsorption will be due to hydrophobic<br />

binding alone. By changing the pH to introduce<br />

in the protein a net charge of the same sign as the<br />

charge of the adsorbent, the repulsion effect will<br />

decrease the adsorptive force due to hydrophobie<br />

bonding, making possible the desorption. Obviously,<br />

a dra~vback in using this type of adsorbent<br />

arises from the problem that frequently the IP of<br />

the biomolecule of interest is unknown, and that<br />

many proteins precipitate at their IP. Moreover,<br />

the simultaneous presence of extraneous ionic<br />

and hydrophobic groups in affinity adsorbents,<br />

has been found to cause a substantial non-specific<br />

protein binding thus resulting in a reduced biospecificity<br />

of these materials [106].<br />

3. The influence of the degree of substitution<br />

on the hydrophobicity of the matrix.<br />

The degree of substitution on Sepharose 4B<br />

with aIkyl-amines of different hydrophobicities<br />

is a critical parameter in the adsorption of proteins.<br />

About 1012 alkyl residues per Sepharose<br />

sphere appears to be a critical degree of substitution<br />

for the adsorption of enzymes like phosphorylase<br />

kinase, phosphorylase phosphatase,<br />

3'5'-cAMP dependent protein kinase, glycogen<br />

synthetase, and phosphorylase b which are successively<br />

adsorbed when the hydrophobicity of<br />

the Sepharose is increased. In addition, the<br />

degree of substitution determines the capacity of<br />

the gel [78].<br />

The critical hydrophobicity needed to adsorb<br />

proteins can be obtained by either increasing the<br />

degree of substitution or by elongating the<br />

employed alkyl-amine chain at a constant degree<br />

of substitution. Consequently, as the hydrophobicity<br />

of the gel is increased, higher binding affinities<br />

result and the desorption requires more<br />

and more severe conditions. In the case of neutral<br />

alkyl-agaroses [3~1], elution of proteins from the<br />

hydrophobic matrix can be described in terms of<br />

salting- in phenomena, since desorption is dependent<br />

on the type of salt employed and not on the<br />

ionic strength alone.<br />

In all cases, a minimal chain length seems to be<br />

required in order to obtain the adsorption of a<br />

given protein. This fact has been interpreted as<br />

fitting a hydrophobic group into a hydrophobic<br />

pocket of the protein. By variation of the concentration<br />

of cyanogen bromide in the activation<br />

mixture, the amount of hydrophobic residues<br />

may be varied. Thus, the corresponding hydrophobicity<br />

may be increased such that the amount<br />

of adsorbed material increases exponentially<br />

when the degree of substitution of the gel is<br />

enhanced. Control experiments with methyl-agarose<br />

show that similarly increased degrees of<br />

substitution, and thus similar numbers of positive<br />

charges introduced by the coupling procedure<br />

[117], do not affect the adsorption of the assayed<br />

protein. Therefore, the adsorption of larger<br />

amounts of material, determined by increasing<br />

the degree of substitution, is not a function of<br />

the additional number of charges [78]. One may<br />

conclude that, if in a series of gels of different<br />

hydrophobicities a crude extract containing hydrophobically<br />

differing proteins is applied, the<br />

one vcith the highest hydrophobicity will be adsorbed<br />

by the gel of the lowest degree of substitution.<br />

Then as the number of alkyl residues increases<br />

on the matrix, proteins of lower hydrophobicities<br />

are adsorbed.<br />

A direct method of determining the degree of<br />

substitution for charged hydrophobie groups has<br />

been proposed by Hofstee [89]. The method describes<br />

the use of Ponceau S, a dye which carries<br />

hydrophobic groups in conjunction with an overall<br />

negative charge bound in an irreversible<br />

fashion to the ligand. Under a given set of conditions,<br />

and after application of a saturating amount<br />

of dye, a certain amount will remain bound even<br />

after extensive washing. Eighty-five percent of the<br />

Ponceau S binding capacity is lost after a period<br />

of almost 5 months indicating that the degree of<br />

substitution decreases gradually upon storage. In<br />

some cases, 40 per cent was lost in only 40 days.<br />

BIOCHIMIE, 1978, 60, n ° 1.


1 0 J.-L. Ochoa.<br />

The data also suggest that the adsorbents with the<br />

highest degree of substitution are the least stable<br />

[89].<br />

An estimation of the relative degree of substitution<br />

can be obtained from the acid-base treatment<br />

of the adsorbent [89], by a nucleophilic attack<br />

of N-substituted isourea (derivated from the<br />

coupling procedure 'with BrCN) with an active<br />

chromogenic substance [70] ; or more indirectly,<br />

by measuring the capacity of the adsorbent with<br />

a coloured protein like cytochrome C [90] or phycoerythrin<br />

[56]. A inuch more sophisticated but<br />

accurate method utilizes NMR spectra [56].<br />

Finally, it is important to mention that a consequence<br />

of the degree of substitution is the shrinkage<br />

of the gel owing to a decrease in hydrophilicity<br />

due to the hydrophobic character of the<br />

ligand. Depending on the special structure of the<br />

agarose gel, the shrinkage seems to be much less<br />

important than with other gels like Sephadex or<br />

polyacrylamide [90].<br />

B. The influence of salt.<br />

The influence of salt on the adsorption of proteins<br />

to hydrophobic supports is probably due<br />

to a number of factors acting on the protein as<br />

well as on the matrix. It has been demonstrated<br />

that neutral salts induce conformational and<br />

structural changes in biomolecules [34]. These<br />

studies have been carried out by circular dichroism<br />

spectra of the protein at cons,taut ionic<br />

strength. It has been concluded that salting-out<br />

ions cause conformational but not structural changes<br />

whereas salting-in ions cause sometimes<br />

severe structural changes which may be one reason<br />

of their denaturing effect [34].<br />

The > properties of saltingout<br />

ions enhance intramolecular, as well as intermolecular,<br />

hydrophobic bonding as reflected by a<br />

stabilization of the hydrophobic core of the biomolecule<br />

[35, 43']. The effect of specific ions on<br />

macromolecules was first noticed by Hofmeister<br />

[36]. He found that salts differ greatly in their<br />

ability to salt-out proteins at a given salt concentration.<br />

At high concentrations of salting-out ions,<br />

the solubility of the protein is adversely affected<br />

by decreasing the availability of ~vater molecules<br />

in the bulk and increasing the surface tension of<br />

water, resulting in an enhancement of the hydrophobic<br />

<strong>interaction</strong>s [31, 371. Accordingly, the influence<br />

of neutral salts on the adsorption phenomena<br />

determines to a large extent the degree of<br />

adsorption and correlates closely with the Hofmeister<br />

series [63, 58]. Salting-in ions, on the other<br />

BIOCHIMIE, 1978, 60, n ° 1.<br />

hand, are > ions, and thus<br />

do not favour hydrophobic <strong>interaction</strong>s. This<br />

effect can be regarded as an inevitable consequence<br />

of the new order imposed by the ion<br />

orienting water molecules in such a way that<br />

water cannot undergo further positive entropy<br />

change, as it is required for the formation of hydrophobic<br />

bonds [18]. These ions have also been<br />

termed ¢ chaotropic >> [44] because they provoke<br />

unfolding, extension and dissociation of the macromolecules.<br />

In this respect, they might be used<br />

in the elution of strongly adsorbed materials on<br />

hydrophobic supports.<br />

Already in 1948, Tiselius [38] noticed that proteins<br />

and other substances (e.g., dyes) that could<br />

be precipitated by high concentrations of neutral<br />

salts could be adsorbed (at much lo'wer salt concentration)<br />

to common adsorbents, whereas in the<br />

absence of salts those adsorbents showed no affinity<br />

for the substance. In the years afler~vards,<br />

some attempts have been made to purify proteins<br />

on solid supports using gradients of ammonium<br />

sulfate [39, 40, 54]. In general, proteins which precipitate<br />

at low ammonium sulfate concentrations<br />

should likewise be retained on hydrophobic supports<br />

at low salt concentrations ; and similarly,<br />

proteins ~vhich precipitate at high concentrations<br />

of ammonium sulfate would require relatively<br />

higher concentrations of salt to obtain their retention<br />

[40].<br />

The purification of nucleic acids [41] has also<br />

been reported using unsubstituted agarose with<br />

high concentrations of ammonium sulfate. It<br />

should be emphasized that tbe adsorption occurs<br />

at concentrations below which the macromolecules<br />

precipitate out of solution. Since the adsorption<br />

is controlled by salt concentration, rather<br />

than by the hydrophobicity of the column, the<br />

names of > and hydrophobic (salting-out) <strong>chromatography</strong><br />

have been employed elsewhere [42, 54.<br />

67]. Hovccver, the hydrophobicity of the gel and<br />

the effect of the salt concentration may be combined<br />

efficiently to improve the adsorption phenomena<br />

and/or the purification [64J, especially of<br />

proteins which may be affected by the high salt<br />

concentration required for its adsorption on the<br />

given gel [46]. If the use of high salt concentrations<br />

is limited by the enzyme activity or stability<br />

for instance, a comparatively longer alkyl-chain<br />

must be employed in order to obtain sufficient<br />

capacity. Because a similar salting-out/salting-in<br />

effect has been noticed in the case of high concentrations<br />

of sugars, namely sugaring-out and<br />

sugaring-in effect [45], the possibility in using


<strong>Hydrophobic</strong> (<strong>interaction</strong>) <strong>chromatography</strong>. 11<br />

sugars instead of salts should be considered. However,<br />

this effect is probably not general.<br />

C. Effect of temperature and pH.<br />

From the equation N ° 1, it is clear that the temperature<br />

may influence, in a positive way, the<br />

adsorption of proteins on hydrophobic matrices<br />

~52, 551. However, temperature may also provoke<br />

other effects, such as increased solvation and decreased<br />

surface tension, ~vhich may withstand the<br />

ability of molecules like proteins to interact xvith<br />

hydrophobic supports [53J.<br />

The protein molecules retain their biological<br />

ac.fivity or capacity to function only within a<br />

limited range of temperature and pH. Their exposure<br />

to extremes of pH and temperature causes<br />

them to undergo denaturation ; the most visible<br />

effect is a decreased solubility of globular proteins.<br />

Since the covalent chemical bonds in the<br />

peptide backbone of the protein are not broken<br />

during denaturation, it has been concluded that<br />

it is due to the unfolding of the characteristical<br />

conformation of the native form of the protein<br />

molecule. The refolding of a denatured protein<br />

does not require the input of chemical work from<br />

outside. It proceeds spontaneously, provided that<br />

the conditions of temperature and pH are adjusted<br />

to be compatible with the stability of the native<br />

conformation of the protein. In this respect, the<br />

process is similar to the one in hydrophobic bond<br />

formation.<br />

The strength of hydrophobic bonds should increase<br />

with rising temperature up to above 60°C,<br />

at which point the additional stability arising<br />

from hydrogen bonding, electrostatic forces between<br />

charges or dipoles, van der Waals <strong>interaction</strong>s<br />

and disulfide bridges disappear, and turn<br />

against the favoured hydrophobic <strong>interaction</strong> [18,<br />

30, 50] of the protein.<br />

A pH effect on hydrophobic bonding is observed<br />

in the case, for example, of bovine serum albumin<br />

binding alkanes at pH 4. It 'was observed<br />

that at low pH the site of hydrophobic bonding is<br />

disrupted [61J, and that the degree of aggregation<br />

of the globulin does not affect the degree of binding.<br />

The conclusion ~vas that the sites associated<br />

with aggregation are relatively non-hydrophobic<br />

and that any conformational changes, resulting<br />

from the polymerization do not exert any effect<br />

at the hydrophobic binding site of alkanes [62].<br />

Decreasing pH belo'w pK value of the amino<br />

group of adenylic and cytidylic acid residues of<br />

tRNA changes the negative charge of the molecule,<br />

BIOCHIMIE, 1978, 60, n ° 1.<br />

altering its tertiary structure [65] and provoking<br />

its desorption from the charged alkyl amine agarose<br />

columns. On the other hand, binding at<br />

pH 7.5 does not occur probably due to other conformational<br />

changes in the molecule, which could<br />

mask key sites involved in the binding E41_~.<br />

Lysozyme, a basic protein, becomes more and<br />

more retarded on Clo (aIkyl Sepharose) as the pH<br />

of the eluant is lowered, and binds to this column<br />

at pH 1.5 [661. Cytochrome C is also highly pHdependent<br />

in its adsorption to neutral ethers of<br />

agarose derivatives [67J.<br />

No influence of pH has been observed on the<br />

adsorption of polysaccharides to substituted agaroses<br />

[68~. In this case, the retention is mostly<br />

determined by the hydrophobicity of the column<br />

and by the size of the polymer.<br />

Naturally, pH plays an important role in the<br />

adsorption of biomolecules on hydrophobic matrices<br />

carrying charged groups. With this type of<br />

adsorbents, pH may modify the charge either<br />

of the biomolecule or of the adsorbent. By<br />

changing the pH, a net charge may be introduced<br />

in protein, which can be of the same sign,<br />

or opposite of the charge on the adsorbent. In<br />

the first case, the repulsion effect ~vill affect negatively<br />

the adsorption phenomena. As was pointed<br />

out previously, a maximal hydrophobic adsorption<br />

will thus occur at the isoelectric point of the<br />

protein E9!.<br />

VII. METHODS OF DESORPTIOX on ELUTION<br />

OF ADSORBED MATERIALS<br />

ON HYDROPHOBICS MATRICES.<br />

The elution of adsorbed materials from hydrophobic<br />

gels can be obtained in a number of ~vays<br />

depending on the type of adsorbent employed,<br />

the conditions in 'which the adsorption occurred,<br />

and the properties of the biomolecule.<br />

When proteins have been adsorbed at high salt<br />

concentrations, a decreasing ionic strength of the<br />

eluant ~vill usually result in the removal of the<br />

material from neutral adsorbents [71, 791. If the<br />

support carries charges introduced either by the<br />

coupling procedure [70], or by the nature of the<br />

ligand, electrostatic <strong>interaction</strong>s will become<br />

important at low ionic strength, and the elution<br />

will not be possible [47, 1031. In such cases,<br />

changes in the pH [9, 65], buffer composition<br />

[47], temperature, or the addition of nonpolar<br />

organie substances may be quite useful<br />

procedures.


12 J.-L. Ochoa.<br />

The fact that in many cases elution is possible<br />

by raising salt concentration, demonstrates the<br />

presence of cooperative electrostatic <strong>interaction</strong>s<br />

in the overall adsorption phenomena in the case<br />

of charged-hydrophobic supports. Nevertheless, it<br />

has been shaw that in those situations, the mechanism<br />

of hydrophobic adsorption at high salt<br />

concentration prevents the electrostatic <strong>interaction</strong>s<br />

of being the main driving force involved<br />

in the process.<br />

The use of denaturing agents, such as urea,<br />

guanidine-HC1 etc, has been successfully applied<br />

in desorption procedures. Their mechanism of<br />

action appears to be involved in both direct <strong>interaction</strong><br />

'with the side-chains and peptide groups<br />

and disruption of hydrophobic <strong>interaction</strong>s between<br />

side chains [23, 69, 77, 109].<br />

Chaotropie ions, on the other hand, are known<br />

for their ability to impart lipophilic properties to<br />

water [72] and to alter the structure of biomolecules<br />

[44]. Their application for the elution of<br />

proteins from high members of alkyl-agarose derivatives<br />

has been reported to be successful [73] ;<br />

ho'wever, it is important to bear in mind their<br />

property of dissolving agarose gels and denaturating<br />

proteins. In this case, cross-linked agaroses,<br />

which are kno'~vn to support more severe conditions<br />

(pH, temperature, ionic strength etc) than<br />

the normal agarose gel, are to be employed.<br />

As has been mentionned before, by increasing<br />

the hydrophobicity of the gel tighter binding resuits,<br />

and desorption of proteins requires ever<br />

more drastic conditions [78]. The combination of<br />

salt gradients with > agents<br />

(e.g., ethylene glycol):, variation in pH or temperature,<br />

etc., may give an increased selectivity of<br />

the elution. As an illustration, it can be mentionned<br />

the case of serum albumin, where elution<br />

occurs at pH 3 with 50 per cent ethanol in the<br />

buffer [791. Similarly, the use of mixed gradients<br />

of buffer solution with high salt concentration,<br />

and buffer without salt but containing 50 per cent<br />

of ethylene glycol, have been successfully applied<br />

to desorb enzymes such as chymotrypsin and<br />

trypsin from hydrophobic matrices [90].<br />

A simple pure salt gradient may result in a<br />

strongly diluted peak in the elution because the<br />

cond,itions in 'which the equilibrium for the<br />

hydrophobic <strong>interaction</strong> is attained are slob,.<br />

Only lower flow rates could provide more favourable<br />

results. In addition, it should be emphasized<br />

that in some cases the elution with ethylene<br />

glycol, in the absence of salts, may be inefficient<br />

due to the presence of electroslatie <strong>interaction</strong>s<br />

BIOCHIMIE, 1978, 60, n ° 1.<br />

[98] when adsorbents involving mixed effects are<br />

employed.<br />

The use of detergents [127] is usually the last<br />

resource, when the elution by other milder conditions<br />

has not been possible. The problem in using<br />

detergents is their intrinsic denaturing effect and<br />

the difficulty of their removal from the columns<br />

during the regeneration step [127]. A wide range<br />

of different detergents has been used varying in<br />

their efficiency to eliminate strongly adsorbed<br />

materials E127]. In this respect, the more ionic<br />

ones are to be preferred for the reasons discussed<br />

above.<br />

Finally, flat curves are obtained by gradient<br />

elution [75], presumably due to a postulated<br />

irregularity of the matrix (see fig. 2) and the<br />

occurence of a wide range of binding sites of<br />

varied strengths [74]. For this reason, attempts<br />

should be made to fractionate through > [123] (as opposed to differential-elulion)<br />

on a series of adsorbents of<br />

increasing hydrophobicities. In this way, each<br />

protein tends to be adsorbed or bound to the<br />

column that provides the minimum degree of<br />

hydrophobicity required for binding, and complete<br />

elution of the materials can be accomplished<br />

with mild eluants. Another alternative consists<br />

of the use of a high member of the alkyl agarose<br />

series, which may retain a high percentage of the<br />

protein content of a particular mixture, and<br />

allows the elution of the molecule of interest<br />

under mild conditions nvhile most of the other<br />

protein remain adsorbed on the column [123].<br />

Obviously, this possibility can be utilized only<br />

when the biomolecule in question has a lo~,-<br />

hydrophobicity. One should note that in principle<br />

it should be possible to achieve purification by<br />

HIC, not only by a selective retention of a given<br />

protein, but also by exclusion of a desired protein<br />

when most of the other proteins in the mixture<br />

are adsorbed [114].<br />

VIII. APPLICATIONS, FUTURE AND IMPLICATIONS<br />

OF HIC.<br />

The relevance of HIC lies in the fact that it is<br />

perhaps the first technique which takes advantage<br />

of the hydrophobic properties of the biomolecules.<br />

The hydrophobic character of biomolecules<br />

should be a specific property, similar to the ionic<br />

characler, since this is a function of the primary<br />

structure in the case of proteins and nucleic<br />

acids. So, it is not unrealistic to consider that


<strong>Hydrophobic</strong> (<strong>interaction</strong>) <strong>chromatography</strong>. 13<br />

their selective separations are feasible. As an<br />

example, the purification of two interconvertible<br />

forms of glycogen-phosphorylase, which differ in<br />

a unique serine residue, has been successfully<br />

achieved through their ability to adsorb on two<br />

distinct members of the alkyl-agarose series [129].<br />

Unfortunately, the types and nature of the<br />

hydrophobic adsorbents available to date are probably<br />

not well systematized, and the lack of intermediate<br />

hydrophobic matrices limits the applicability<br />

of the method. To provide for a larger<br />

variety in the types of ligands, including the<br />

introduction of aromatic structures, it might be<br />

expedient to attach these ligands to (< arms )~ that<br />

by themselves are not hydTophobic, e.g., carbohydrates<br />

instead of hydrocarbons [96]. At first<br />

sight, hydrophobically coated supports may be<br />

used as concentrating systems as well [52]. In<br />

fact, the protein 'which is going to be purified<br />

does not require a pre-concentration step if the<br />

conditions under 'which the <strong>chromatography</strong> is<br />

performed allow its adsorption.<br />

The application of HIC to the purification of<br />

nucleic acids using matrices of varied hydrophobicities<br />

according to the type of ligand attached<br />

or to its degree of subslitution awaits for further<br />

studies which may greatly improve the application<br />

of this technique. This aspect is especially<br />

interesting when the molecules are not stable or<br />

soluble at either too high or too low salt concentrations.<br />

The separation of particles, like sub-cellular<br />

fractions or entire cells [110, 121] has been reported,<br />

and it seems to be a very promising application<br />

of HIC, since membranes can be considered<br />

as aggregates of molecules of varied hydrophobicities.<br />

A similar application in a quite different<br />

field of biochemistry is preparation of enzymatic<br />

reactors through the immobilization of enzymes<br />

on hydrophobic supports [32, 101, 122, 126]. The<br />

basic idea is the possibility to have a reactor<br />

which can be periodically recycled, when the<br />

adsorbed material loses its activity by elution<br />

and adsorption of freshly active substances<br />

(which may be of the same or of another biological<br />

activity, but similarly adsorbed on the<br />

hydrophobic support). Thus, the hydrophobic<br />

matrix may function as an universal support of<br />

easier handling than those systems in which the<br />

immobilization necessitates a covalent linkage<br />

between the biomolecule and the adsorbent, with<br />

their corresponding limitations in function and<br />

utility,<br />

It should be reminded that some hydrophobic<br />

ligands denature proteins through a > action [96]. This effect can be reduced<br />

by employing hydrophobic ligands of milder<br />

influence, or by introducing additional polar or<br />

ionic groups in the hydrocarbon chain [9]. Finally,<br />

the problem of >, which<br />

constitutes another important drawback in protein<br />

separation by <strong>chromatography</strong> on hydrophobic<br />

columns, must be considered. If inhomogeneity<br />

occurs in the adsorbent, that is if the interacting<br />

sites on a given member of the suhstituted-agarose<br />

series are different in their strength of <strong>interaction</strong><br />

with a particular protein (fig. 2), elution by<br />

decreasing salt concentration or increasing concentration<br />

of > agents, will<br />

never result in a narrow peak but rather in flat<br />

curves, which increase contamination of the sample<br />

by other proteins. Some suggestions to avoid<br />

this phenomenon have already been mentioned<br />

taking advantage of the hydrophobicity of the<br />

matrix.<br />

In conclusion, the correct application of hydrophobic<br />

<strong>chromatography</strong> implies the consideration<br />

of three main factors :<br />

-- the nature and hydrophobicity of the adsorbent<br />

and of the protein,<br />

-- the conditions of the adsorption,<br />

-- the conditions of the elution of the adsorbed<br />

material.<br />

It may be expected that as the mellmd will be<br />

more and more aplied wi'.h various biomolecules,<br />

in different conditions, our understanding will be<br />

enriched so about many biological events in<br />

which a hydrophobic <strong>interaction</strong> is involved.<br />

Acknowledgments.<br />

The author expresses his gratitude to Dr. Jean-Marc<br />

Egly for his encouragement in the preparation of this<br />

article. Special thanks are due to Professor Shmuel<br />

Shaltiel and Professor Stellan Hjertdn for their criticism<br />

and suggestions, and to Professor Barbarin<br />

Arreguin for his continuous interest on my ,work. My<br />

appreciation to Dr. Dana Fawlkes for the linguistic<br />

reoision and to Mrs. Inga Johansson and Mr. GSsta<br />

ForMing for typing the manuscrip and drawing the<br />

figures, respectioely.<br />

To the National University of Mexico and the National<br />

Council of Sciences and Technology of Mexico<br />

many thanks for their financial support.<br />

REFERENCES.<br />

1. Hirs, C. W. W. (1955) Methods in Enzymol., Colowick<br />

S. P. and Kaplan, eds, Academic Press,<br />

New York° 1, p. 113.<br />

2. Colowick, S. P., ibid, p. 90.<br />

3. Porath, J. (1968) Nature (London), 218, 834.


14 J.-L. Ochoa.<br />

4. Green, A. A. a Hughes, W. L. (1955) Methods in<br />

Enzymol., 5, 67.<br />

5. Cuatreeasas, P. ~ Anfinsen, C. B. (1971) Ann. Reo.<br />

Biochem., 40, 259.<br />

6. Cuatreeasas, P., Wilehek, M. a Anfinsen, C. B.<br />

(196~) Proc. Nat. Acad. Sci. U. S., 61, 636.<br />

7. O'Carra, P..a Griffin, T. (1974) Methods in Enzytool.,<br />

34, 108.<br />

8. Steers, Jr., E.. Cuatreeasas, P. ~ Pollard, H. B.<br />

(1971) J. Biol. Chem., 2~6, 196.<br />

9. Yon, R. J. (1972) Biochem. J., 126, 765.<br />

10. Er-el, Z., Zaindenzaig, Y. & Shaltiel, S. (1972)<br />

Biochem. Bioohys. Res. Commun., 49, 383.<br />

11. Shaltiel, S. (1974) Methods in Enzymol., 34, 126.<br />

12. !Klotz, I. M. (1970) Arch. Biochem. Biophys., 138,<br />

704.<br />

13. Kauzmann, W. (1959) Adv. Prol. Chem., 14, 1.<br />

14. Vol'kenshtein, M. V. (1970) Molecules and Life,<br />

Plenum Press. N. Y., pp. 189 and 357.<br />

15. Didkerson, R. E. ,~ Gets, I. (1969) The structure<br />

and action of proteins, Haper ~ Roxv Publ.,<br />

N. Y., D. 92.<br />

16. Perrin, J. (1905) J. Chem. Phtls., 3~ 50.<br />

17. Langmuir, I. (1917) J. Am. Chem. Soc., 39, 1848.<br />

18. Dandliker, W. B. a de Saussure, V. A. (1971) The<br />

Chemistry of Biosnrfaces, Vol. I (Hair, M. L.,<br />

Ed.), Marcel D~kker, Inc., N. Y.<br />

19. Tanford, C. (1972) J. Mol. Biol., 67. 59.<br />

20. Frank, H. S. ~ Evans, M. W. (1945) J. Chem. Phys.,<br />

13, 5O7.<br />

21. Tanford, C. (19.62) J. Am. Chem. Soc., 84, 4240.<br />

22. Noz~ki, Y. ~ Tanford, C. (19.71) J. Biol. Chem.,<br />

'246, 2211.<br />

23. Bull, H. B. ~ Bresse, K. (1974) Arch. Bioehem., 161,<br />

665.<br />

24. Bigelo,w, C. C. (1967) J. Theor. Biol., 16, 187.<br />

25. Waugh, D. F. (1954) Adv. Prof. Chem., 9. 325.<br />

26. Fisher, H. F. (1964) Proc. Nat. Aead. Sci. U. S., 51,<br />

1285.<br />

27. Dandliker. W. B.. Alonso, R., de Saussnre. V. A.,<br />

Kierszenb.~um. F.. Levison. S. A. ~ Schapiro, H.<br />

C. (1967) P;ocbemistry, 6, 1460.<br />

28. Davey, M. W., Huan~', J. W.. Sulkowski. E. &<br />

Carter. 'W. A. (1974) .L Biol. Chem., '2A9, 635~.<br />

29. Boldt. D. H., Spee~rart. S. F- Richards. R. 1,.<br />

Alving, C. R. (1977) Biochem. Biophys. Res.<br />

Comrnun., 74, 208.<br />

30. Barry, A. ~ Briningar, W. S. (1975) J. Biochem.,<br />

250, 8856.<br />

31. Tanford, C. (1973) The h udrophobic effect : Formation<br />

of micelles and biological membranes,<br />

Wiley, N. Y., p. 9.<br />

32. Davis, M. A. F.. Hanser. H.. Leslie. R. B.<br />

Phillil)s, M. C. (1973) Bioehim. Biophys. Aeta,<br />

3]7, 214.<br />

33. Brandts, J. F. (1969) Biol. MacromoL, 2, 213.<br />

34. P~hlman, S., Rosengren, J. ~ Hjert~n, S. (I977)<br />

J. Chromatoqr.. I.-~1. 99.<br />

35. Hofstee, B. H..L (I975) Biochem. Biophys. Bes.<br />

Commnn., ~;~. 618<br />

36. Hofmeister. F. (1888) Arch. Exp. Pathol. Pharmakol.,<br />

24, 247.<br />

37. Tanford, C. (1968) Adv. Prof. Chem., 2.~, 121.<br />

38. Tiselius, A. (1948) Ark. Kemi Min. Geol., B 26.<br />

39. King, T. P. (1972) Biochemistry, ll, 367.<br />

40. Memoly, V. A. ~ Doellgast, G. J. (1975) Biochem.<br />

Biophys. Res. Commun., 66, 1011.<br />

41. Holmes, W. M., Hurd, R. E., Reid. B. R., Rimerman,<br />

R. A. ~ Hattie]d, G. Vv'. (1975) Proc. Nat.<br />

Acad. Sci. U. S., 72, 1068.<br />

42. Doellgast, G. J. &Plaut, A. G. (1976) Immunochem.,<br />

1;3, 135.<br />

43. Von Hippel, P. H. ~ Schlieh, T. (1969) Structure<br />

and stability of macromolecules (Timascheff, S.<br />

N. and Fasman, G. D., Eds), Marcel Dekker, N.<br />

Y., p. 41.<br />

44. Hamaguchi, K. & Geiduschek, E. P. (1962) .L Am.<br />

Chem. Soc., 84, 1329.<br />

45. Lakshmi, T. S. ,& Nandi, P. K. (1976) Y. Chromatoyr.,<br />

116, 177.<br />

BIOCHIMIE, 1978, 60, n ° 1.<br />

46. Hamar, L., P~hlman, S..~ Hjert6n, S. (1975) Biochim.<br />

Biophys. Acta, 403, 554~<br />

47. Mevarech, M., Leicht, W. & Werler, M. M. (1976)<br />

Biochemistry, 15, 2383.<br />

48. Larsen, H. (1967) Advances in Microbiolotlical<br />

Physiology (Rose~ A. H. ~ Wilkinson, J. F., Eds.),<br />

Aead. Press, N. Y., p. 97.<br />

49. Reistad, R. (1970) Arch. Mikrobiol., 71, 353.<br />

50. Shanbhag, V. P. ~ Axelsson, C. C. (1975) Eur. J.<br />

Biochem., 60, 17.<br />

51. Hjert6n, S. (1976) Meth. Prof. Sep., 2, 233.<br />

52. Hjert~n, S., Rosengren, J. ~ Phhlman, S. (1974) J.<br />

Chromatoar., 101. 281.<br />

53. Jennissen, H. P. (1976) Biochemistry, 15, 5683.<br />

54. Rimerman, R. A. ~ Hatfield, G. W. (1973) Science,<br />

182, 1268.<br />

55. Hjert~n, S. (1973) J. Chromatoar., 87, 325.<br />

56. Rosengren, J., Phhlman, S., Glad, M. & Hjert6n,<br />

S. (1975) Biochim. Biophys. Acta. 412, 51.<br />

57. O'Carra, P.. Barry, S..& Griffin, T. (1974) FEBS<br />

Letters, 43, 169.<br />

58. Visser, J. • Strating, M. (1975) Biochim. Biophys.<br />

Acta, 384, 69.<br />

59. Scheraga. H. A., Nemethy, G. ~ Steinberg, I. Z.<br />

(1962) J. Biol. Chem., ~7, 2506.<br />

60. Hertz, H. G. • Zeidler, M. D. (1964) Z. Electrochem.,<br />

68, 821.<br />

61. Wishnia, A..& Pinder, Jr., T. W. (1964) Biochemistry,<br />

3, 1377.<br />

62. Wishnia, A. &Pinder, Jr., T. W. (1966) Biochemistry,<br />

5, 1534.<br />

63. Raibaud, O., H6gberg-Raibaud, A. ~ Goldberg, M.<br />

(1975) FEBS Letters, ;50, 130.<br />

64. Visser, J. ~ Strating, M. (1975) FEBS Letters, 57,<br />

183.<br />

65. Dziegiele~vski, T. & Jakubo~wski, H. (1975) J. Chromatogr.,<br />

10g, 364.<br />

66. Shaltied, S. (1975) in Enzymes (Desnuelle, P., Ed.),<br />

North Holland, Amsterdam, pp. 117-127.<br />

67. Jennissen, H. P. (197'5) Biochemistry, 14, 754.<br />

68. Busey, H., Rimerman, R. A. ,~ Hatfield. G. W.<br />

(1975) Anal. Biochem., 64, 380.<br />

69. Oakenfull, D. G. (1974) Proc. Aust. Bioehem. Soc,<br />

7, 20.<br />

70. Weber, M. M. (1976) Anal. Biochem., 76, 177.<br />

71. Sch6p,. W., Meinert, S., Thyfronitov, J. ~ Anrich,<br />

M. (1975) J. Chromatogr.. 10d, 99.<br />

72. Hatefi. Y. • Hanstein, V~'. H. (1969) Proc. Nat.<br />

Acad. Sci. U. S., 62, 1129.<br />

73. Neurath. A. R.. Lerman. S.. Chen, M. ~ Prince, A.<br />

M. (1975) J. Gen. Virol., 28. 251.<br />

74. Hofstee, B. H. J. (1974) in > (Dunlop, R. B.,<br />

Ed.), Plenum Pub]. Corr., N. Y., op. 43-59.<br />

75. Hofstee, B. H. J. (1975) Prep. Biochem., 5, 7.<br />

76. Weiss. H. a Bueher, T. (I970) Eur. J. Biochem.,<br />

17, 561.<br />

77. Bizelis, R..a Umbarger, H. E. (1975) J. Biol.<br />

Chem., ~5~. 4315.<br />

78. Jennissen, H. P. ,~ Heilmeyer, Jr., L. M. G. (1975)<br />

Biochemistry, 14, 754.<br />

79. Peters. Jr., T~, Taniuehi, H. a Anfinsen, Jr., C. B.<br />

(1973) J. Biol. Chem., ~4'8. 9447.<br />

80. Lee, B. a Richards, F. M. (1971) J. Mol. Biol., 55,<br />

379.<br />

81. Chothia, C. H. (1974) Nature (London), 248, 338.<br />

82. Klotz, I. M. ,¢ Franzen, D. S. (1945) J. Am. Chem.<br />

Soc., 67, 1003.<br />

83. Chothia, C. ~ Janin, J. L. (1975) Nature (London),<br />

256, 705.<br />

84. Jennissen, H. P. (1976) Hoppe-Seyler's Z. Physiol.<br />

Chem., ,357, 1201.<br />

85. Arnott, S., Fulmer, A. Scott, E. E., Dea, I. C.,<br />

Moorhouse, R. ~ Reis, D. A. (1970) J. Mol. Biol.,<br />

90, 269.<br />

86. Jennissen, H. P. (1976) in Prof. Biol. Fluids, Proc.<br />

Colloq., 1975 (Peeters, L. H., Ed.), Vol. 23, Pergamon<br />

Press, Oxford ~ N.Y., pp. 675-679.<br />

87. Cuatrecasas, P. (1970) J. Biol. Chem., 245, 3059.<br />

88. Barry, S. ~ O'Carra, P. (1973) Biochem. J., 135,<br />

595.


<strong>Hydrophobic</strong> (<strong>interaction</strong>) <strong>chromatography</strong>. 15<br />

89. Hofstee, B. H. J. (1974) Adv. Exp. Med. Biol., 42,<br />

43.<br />

90. L/~/ls, T. (1975) J. Chromatogr., 111, 373.<br />

91. Ellingboe, J., Nystr6m, E. ,¢ Sj(ivall, J. (1970) J.<br />

Lipzd Res., 11, 266.<br />

92. Clifford, J., Pethica. B. A. ~ Senior, W. A. (1965)<br />

Ann. N. Y. Acad. Sei., 125, 458.<br />

93. Porath, J., Sundberg, L., Fornstedt, N. ,~ Olson, I.<br />

(1973) Nature (London), 245, 465.<br />

94. Yon, R. J. ~ Simmonds, R. J. (1975) Bioehem. J.,<br />

161, 281.<br />

95. Jost, R., Miron, T. ,¢ Wilchek, M. (1974) Biochim.<br />

Biophys. Acta, 362, 75.<br />

96. Hofstee, B. H. J. (1973) Biochem. Biophys. Res.<br />

Commun., 50, 751.<br />

97. Geren, C. R., Magee, S. C. ,~ Ebner, K. E. (1976)<br />

Arch. Biochem. Biophys., 172, 149.<br />

98. Ho~ksc, H. (1974) Carbollyd. Res., 37, 390.<br />

99. Shaltiel, S., Adler, S. P., Purich, D., Caban, C.,<br />

Senior, P. ~ Stadtman, E. R. (1975) Proe. Nat.<br />

Acad. Sci. U. S., 72, 3397.<br />

100. Palade, G. E. (1959) in

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

Saved successfully!

Ooh no, something went wrong!