+ All Categories
Home > Documents > Model studies in bio-organic processes : sodium transport ... · Sodiun1 ion transport across...

Model studies in bio-organic processes : sodium transport ... · Sodiun1 ion transport across...

Date post: 18-May-2018
Category:
Upload: doanxuyen
View: 213 times
Download: 0 times
Share this document with a friend
98
Model studies in bio-organic processes : sodium transport across biological membranes : an experimental study : quantum chemical calculations on the stereochemistry of coenzyme B12 dependent carbon- skeleton rearrangements Merkelbach, I.I. DOI: 10.6100/IR179116 Published: 01/01/1985 Document Version Publisher’s PDF, also known as Version of Record (includes final page, issue and volume numbers) Please check the document version of this publication: • A submitted manuscript is the author's version of the article upon submission and before peer-review. There can be important differences between the submitted version and the official published version of record. People interested in the research are advised to contact the author for the final version of the publication, or visit the DOI to the publisher's website. • The final author version and the galley proof are versions of the publication after peer review. • The final published version features the final layout of the paper including the volume, issue and page numbers. Link to publication Citation for published version (APA): Merkelbach, I. I. (1985). Model studies in bio-organic processes : sodium transport across biological membranes : an experimental study : quantum chemical calculations on the stereochemistry of coenzyme B12 dependent carbon-skeleton rearrangements Eindhoven: Technische Hogeschool Eindhoven DOI: 10.6100/IR179116 General rights Copyright and moral rights for the publications made accessible in the public portal are retained by the authors and/or other copyright owners and it is a condition of accessing publications that users recognise and abide by the legal requirements associated with these rights. • Users may download and print one copy of any publication from the public portal for the purpose of private study or research. • You may not further distribute the material or use it for any profit-making activity or commercial gain • You may freely distribute the URL identifying the publication in the public portal ? Take down policy If you believe that this document breaches copyright please contact us providing details, and we will remove access to the work immediately and investigate your claim. Download date: 25. Jun. 2018
Transcript

Model studies in bio-organic processes : sodiumtransport across biological membranes : an experimentalstudy : quantum chemical calculations on thestereochemistry of coenzyme B12 dependent carbon-skeleton rearrangementsMerkelbach, I.I.

DOI:10.6100/IR179116

Published: 01/01/1985

Document VersionPublisher’s PDF, also known as Version of Record (includes final page, issue and volume numbers)

Please check the document version of this publication:

• A submitted manuscript is the author's version of the article upon submission and before peer-review. There can be important differencesbetween the submitted version and the official published version of record. People interested in the research are advised to contact theauthor for the final version of the publication, or visit the DOI to the publisher's website.• The final author version and the galley proof are versions of the publication after peer review.• The final published version features the final layout of the paper including the volume, issue and page numbers.

Link to publication

Citation for published version (APA):Merkelbach, I. I. (1985). Model studies in bio-organic processes : sodium transport across biological membranes: an experimental study : quantum chemical calculations on the stereochemistry of coenzyme B12 dependentcarbon-skeleton rearrangements Eindhoven: Technische Hogeschool Eindhoven DOI: 10.6100/IR179116

General rightsCopyright and moral rights for the publications made accessible in the public portal are retained by the authors and/or other copyright ownersand it is a condition of accessing publications that users recognise and abide by the legal requirements associated with these rights.

• Users may download and print one copy of any publication from the public portal for the purpose of private study or research. • You may not further distribute the material or use it for any profit-making activity or commercial gain • You may freely distribute the URL identifying the publication in the public portal ?

Take down policyIf you believe that this document breaches copyright please contact us providing details, and we will remove access to the work immediatelyand investigate your claim.

Download date: 25. Jun. 2018

MODEL STUDIES IN BIO-ORGANIC PROCESSES.

Sodiun1 ion transport across biologica! membranes. An experimental study.

Quanturn chemical calculations on the stereochemistry of coenzyn1e B12 dependent carbon-skeleton rearrangements.

lnge Merkelbach

MODEL STlTDIES IN BIO-ORGANIC PROCESSES.

Sodium ion transport across biological membranes. An experimental study.

Quanturn chemical calculations on the stereochemistry of coenzyme B12 dependent carbon-skeleton rearrangements.

PROEFSCHRIFT

TER VERKRIJGING VAN DE GRAAD VAN DOCTOR IN DE

TECHNISCHE WETENSCHAPPEN AAN DE TECHNISCHE

HOGESCHOOL EINDHOVEN, OP GEZAG VAN DE RECTOR

MAGNIFICUS, PROF. DR. S.T.M. ACKERMANS, VOOR

EEN COMMISSIE AJ,NGEWEZEN DOOR HET COLLEGE VAN

DEKANEN IN HET OPENBAAR TE VERDEDIGEN OP

VRIJDAG 10 MEI 1985 TE 16.00 UUR

DOOR

INGRID IRENE MERKELSACH

GEBOREN TE MOZAMBIQUE

DIT PROEFSCHRIFT IS GOEDGEKEURD

DOOR DE PROMOTOREN

PROF. DR. H.M. BOCK

EN

PROF. DR. K. VAN DAM

COlft'ENTS

Sodium ion transport across biologica! membranes.

An experimental study.

I. Introduction.

I. 1 •

I. 2.

I. 3.

I. 4.

Membrane properties.

Membrane function. Scope of the first part

of this thesis.

Developments in organophosphorus chemistry.

Transfer of conformational changes in a

phosphate group to the hydrophobic part of

organic molecules.

References and notes.

II. Sodium ion transport across biologica!

membranes.

9

11

1 4

18

22

!1.1. Theory. 25

!1.1.1. Cluster formation in biological membranes 25

induced via a phospholipid P(V) TBP

intermediate.

!!.1.2. Activatien of membrane proteins

in phospholipid clusters.

II.2. Experiments.

II.2.1. Introduction.

II.2.2. Results.

11.2.3. Discussion.

II.2.4. Conclusion.

II.2.5. Experimental.

References and notes.

by uptake 33

37

37

40

43

50

50

53

Quantum chemical calculations on the stereochemistry of

coenzyme 812-dependent carbon-skeleton rearrangements.

III. ~e formation of a carbanionic intermediate

in the carbon-skeleton rearrangement step.

III.l Structure and function of coenzyme B12• III.2. Mechanism of action of the carbon-skeleton

rearrangements. III.3. The stereochemistry of the carbon-skeleton

rearrangements as 'test' for the carbanionic

machanism. Scope of the second part of this

thesis. III.4. The nature of the hydrogen transferred

temporarily to coenzyme B12 during the carbon-skeleton rearrangements.

References and notes.

IV. Quantum chemical calculations.

IV. 1.

IV. 2.

IV. 3.

IV. 4.

IV. 5.

Introduction.

The choice of the calculational method. Results. Discussion.

Conclusion.

References and notes 1

SUIIIlllary

Samenvatting Curriculum Vitae Dankwoord

59

62

65

67

69

73

74

76

85

87

88

89.

91

93

94

Voor het slagen van het kwaad is niets anders

nodig dan dat de goede mensen niets doen.

Maarten Luther King.

Sodium ion transport across biologica! membranes.

An experimental study.

I. Introduction.

I.l. Membrane properties.

Biologica! membranes play a crucial role in almost all

cellular phenomena, yet our understanding of the molecular

organization of membranes still can be called far from

exhaustive. While the composition of membranes varies with

their source, they generally contain approximately 40% of

their dry weight as lipid and 60% as protein1. Usually

carbohydrate is present to the extent of 1-10% of the total

dry weight. In addition to these components, membranes

contain some 20% of their total weight as water, which is

tightly bound and essential to the maintenance of their

I.l. The fluid mosaic mode of Singerand Niaolson. Integral proteins

(crossing the lipid bilayer), pheripheral proteins (bound to the exterior

the bilayer) and proteins embedded in the matrix are bound to

a functional complex or dissolved individually in the membrane bilayer.

9

structure. These components are organized according to the

fluid mosaic model of Singer and Nicolson2 (see Figure I.1.). The lipids span a discontinuous bilayer, with their

hydrophobic tails pointing towards the interior of the membranes and their hydrophilic head groups in contact with

the water phase outside the membrane. In this bilayer integral proteins are embedded, occasionally crossing the

total lipid bilayer matrix, while pheripheral proteins are

bound exterior to the bilayer. Dependent on e.g. temperature

and water content3, the more or less extended hydracarbon ebains of the lipide tilt away from the perpendicular to the

plane of the membrane, thus changing the ratio of the cross

sectional area~ of head group and chain region4. In this way a modification in density of the membrane can be reached, comparable to the melting phenomena of classica! chemica! compounds, e.g. from a fluid liquid-crystalline to a solid gel-like phase. This melting can, even for pure lipids, not be described as a thermodynamic first-order phase transition, since the transition is certainly not discontinuous, as shown, for example, by measurements of volume changes5. One can imagine a gel-like phase in which an appreciable lateral

diffusion of the lipid molecules exists, while on increasing temperature this lateral diffusion will be accelerated

throughout a phase-transition region to the fluid liquid­

crystalline phase. The width of this phase-transition region will be increased when mixing different lipids with each other, with proteins or with other membrane constituents5. In

natura! membranes local gel-like domains6 exist over a large

temperature range in liquid-crystalline matrices and vice­versa. These domains, often called clusters, develop in a continuing process of ordering and successive relaxation to a

disordered state. They consist of mainly lipide, mainly proteins or a mixture of both, but always can be described as a region of different density compared to the surrounding matrix. Diffusion over longer distances sametimes will be

opposed in biologica! membranes by a cytoskeletal system locallzing essential proteins in a well-defined region and by

10

a number of multivalent ligands that can induce aggregation

of proteins into clusters, patches or caps? in an alternative

way as mentioned above. Here proteins reside in a defined

region near to each other, but are not necessarily located in

one and the same cluster in the sense of same degree of order

or same density. When the word cluster is used throughout this thesis, a small region of one aggregational state in a

matrix of another will be ment, the first description of

clusters given above.

I. 2. Membrane function. Scope of the first part of this

thesis.

There is a lot of controversy concerning the role of lipids

in a membrane. Some authors state that lipids are the

insulating constituents of the membrane, separating different

cellular compartments8. They form the structural support of

membrane proteins, thus maintaining a constant spatial

relationship between them. In this view ion transport across

~tf!!il\ ~~

plasma membrane Golgi membrane rough endoplasmie reticulum

nuclear membt'lli'le outer mitochondria! membrane inner mitochondria! membrane

I.2. Lipid aomposition by weight of different subcellu~ membranes

rat liver; PL phospholipid; C cholesterol; G glycolipid; unlabel-

led= mono-> and triacylglycerols, free fatty acids and stearyl esters.

11

plasma membtane Golgi membtane rough endoplasmie reticulum

w PS

nuclear membtane outer mitochondria! membrane inner mitochondria! membrane

Figure I.3. Phospholipid aomposition by weight of different subaelluLar

membranes of rat Ziver; PC = phosphatidyZahoZine; PE = phosphatidyZethanoZ­

amine; PI = phosphatidylinositoZ; PS = phosphatidylserine; S sphingo-

myeZin.

membranes is a property of integral proteins, that are influenced by external factors like metal-ions, protons, potential field etc. If one conaiders however, that lipid

composition of animal membranes vary both with their tissue souree and intracellular location9 (see Figure I.2 and I.3), the question is raised whether the lipids could play a role

themselves. So it has been suggested that the level of free fatty acids, that varies with the functional state of the

membrane, may be involved in changes in membrane permeabil­ity10. Until now, little attention has been paid to the

varying amounts of phospholipids in the membranes, in

particular to the·transition of the four co-ordinated to the

five co-ordinated state of the phosphate group, bound on its crucial position between the polar headgroup and hydrophobic

hydracarbon region, in relation to the phenomenon of ion

transport. Recent studies on a number of model compounds for biologica! reactives11,12 suggest a general principle in

passing biologica! information from ionic polar regions to

hydrocarbon zones of natura! molecules via a five co-ordi­nated (P(V)) trigonal bipyramidal (TBP} intermediate (see

12

Cbapter !.3. and !.4.). A classica! four co-ordinate (P(IV))

tetragonal intermediate, possesses four ligands tbat are

arranged spberically around tbe pbospborus atom, i.e. tbe P-L

bond lengtbs and tbe L-P-L bond angles are equal. Wben a five

co-ordinated TBP intermediate is formed{ a structural

inequivalence between two types of ligands (tbe equatorial

and the axial ones) is introduced, since five ligands can not

be spherically arranged around one atom (see Figure 1.4.).

T

TBP

I. 4. The four co-or'dir.ated tetY>aeder (T J and the co-ordinated

(TBP) of phosphorus.

The axial ligands {an incoming group e.g. water and the group

through wbich tbe pbosphate is bound to tbe main chain of the

biomolecule) are more electron withdrawing groups (see

Chapter !.3.) tban the equatorial ones, thus inducing an

electron flux into tbe axis of the TBP. 1f the group in the

axis consists of tbe 0-C-C-0-sequences often encountered in

biomolecules, this extra negative charge on the axial oxygen

(031 in Figure !.5. in the case of lipids), will result in a

repulsion of the other oxygen (021 in Figure 1.5.). In this

way a conformational change in the phospholipid headgroup

will be transferred in a re-orientation of the hydrophobic

region of tbe lipid molecule. As will be explained in Chapter

11, this re-orientation can induce cluster formation, which

in turn may influence integral membrane proteins as ion

channels, thus triggering them to open or to close. To get

experimental support for this model, a number of vesicles bas

been synthesized with different lipid composition. The

influence of this lipid composition on the ion transport over

13

0 11

..._". P· - N / v,,,,, o-/~o 'o

0 13 ro 2 R 1

0

)=o R ---

Figure I. 5. The ext2•a negative aharge on the apieal paaition of the five

eo-ordinated phospholipid intermediate results in repulsion of the oxygen

bound via an P-D-G-C-0 sequenee.

the vesicle cell wall, using one and the same ionophore, is

described in Chapter II to investigate several aspects of the theory.

1.3. Developments in organophosphorus chemistry.

In the past few decades, research in organophosphorus chemistry bas developed enormously13. Study of the

reactivity of model compounds has greatly enhanced the comprehension of the properties of five co-ordinated phosphorus compounds. So Westheimers studies14-17 on the

hydrolysis of five-membered cyclic phosphonates have increased the understanding of the mechanistic aspects of phosphorylation reactions. It was found that the hydrolysis of five-membered ring phosphates as 1 in Figure 1.6.,

proceeds millions of times faster in comparison to acyclic

phosphates as 4. This will apply for both ringopening (a) and

exocyclic cleavage (b) of 1. In contrast, the cyclic phosphonate 5 gave only very fast ring opening, no exocyclic hydrolysis. Westheiroer explained these observations on the

14

assumption, that the hydrolysis proceeds via a penta

co-ordinated intermediate in a trigonal bipyramidal

configuration. Before these phenomena can be understood, some

properties of five co-ordinated phosphorus compounds should

be noted. As already described in chapter I.2. an important

o"o) •H2o-(

11

0~ /OJ CH:lO -;p'o

~p HO 2

CH 0/ \0 3

0 0 ~/J

HO/ p\0 + CH30H

EtO"-. /OEt 3 p

,~'"-. o o-

4

0 0 0) ~<J + H20

11

CH30 -;p CH3 0

5 HO 6

Figure I.6. Cyclic and acyclic phosphates use« in the studies of West-

heimer.

aspect of five co-ordination is, that the distribution of the

ligands can not be spherically around the central atom, i.e.

the ligands are not equivalent18. Two possible structures are

favoured, as shown by X-ray analysis19-21: the trigonal

bipyramid (TBP) and the square pyramid (SP), shown in Figure

I.7. In the TBP there are three equivalent equatorial and two

axial honds, in the SP one axial and four basal honds.

Theoretical considerations based on MO and electrastatic

calculations have predicted that the TBP is slightly more

stable for acyclic penta co-ordinated phosphorus

derivatives18, but the difference is not too large. In

15

a e ', I

'p-e

e#" I a

TBP SP

a: axial ligand

b: basal ligand e: equatorial ligand

Figure I.?. The trigonal bipyramidal (TBP) versus the square pyramidal(SP)

configuration.

general an ideal TBP is seldom encountered, mostly a TBP is slightly distorted towards an SP geometry. In the TBP configuration the axial bonds are longer and usually weaker

than the equatorial bonds, a picture that can be ascribed by a pd-hybridization for the axial22-24 and a sp2-hybrid­

ization for the equatorial bonds. However, the exact role of d-orbitals is still a subject of controversy25-27. Recent

publications suggest a remarkable degree of s-character in the axial bonds of some radicals28-30. So the observed

differences between ax,ial and equatorial sites in the TBP

structure can, in a more differentiated picture, better be

described by a substantially higher s-character for equator­ia! than for axial bonds. In addition, axial sites are preferred by electron withdrawing ligands, whereas electron donating ligands tend to occupy equatorial positions31. This

polarity rule has been derived from many experimental data32,33, and is supported by semi-empirica! calcula­tions34,35. Furthermore, it has been found that small rings

usually span an axial and an equatorial position in the TBP configuration, due to the 90° angle between these two bonds16. This is known as the strain rule. In fact, the presence of rings stahilizes this contiguration to such an

extent, that most of the known stable phosphoranes contain

one or more rings. One of the consequences of the differences in bond strength in a TBP is that leaving groups depart from the axial position16,34. Due to the microscopie reversibil­

ity16, incoming nucleophiles also enter in the axis of the

Tsp34. An aspect of five co-ordination which hampers

16

differentiation of axial and equatorial bonds in the TBP

configuration, is the existence of pseudorotation.

Pseudorotatien for phosphorus compounds involves that the

positions of the ligands are interconverted fast on NMR

time-scale36-38. Several types of these 'permutational

isomerizations' are known, e.g. the Berry pseudorotation, in which two equatorial and both axial ligands change places39

via an intermediate SP contiguration (see Figure 1.8.).

/1 / I /

/ / :::::::t'P}--..::-. ~V /.,.l. ~· r<--- I(

/ / / V

TBP SP TBP

ii'igure I.B. The Berry tien process.

The energy barrier for pseudorotation may be very low,

especially if all ligands are identical18. However, if

pseudorotations bring electron withdrawing ligands into

equatorial positions34,40, or force small rings to span two

sites of the same kind34, pseudorotatien will be severely

hindered. Using the properties described above, the

experiments of Westheiroer (vide supra) are now readily

explained (see Figure 1.9.). Initial attack of water

on 1 yields intermediate 7. Subsequent proton transfer

towards the endocyclic axial oxygen atom leads to formation of 8, resulting in 2 after ring opening (the axial P-0 bond

is broken}. However, if the P(V) intermediate 1 undergoes ligand reorganization, 9 is formed. Upon leaving of the

axial protonated methoxy group, 3 is generated. The very fast

rate of both processes is explained by the fact that cyclic,

four co-ordinated phosphorus compounds are more strained than their acyclic analogues, whereas cyclic phosphorane

intermediates such as 7 or 9 are stabilized with respect to

acyclic phophoranes. These factors lower the activation

17

7

Figure I.9. Pseudo-rotation in the experiments of Weetheimer.

enthalphy for hydralysis of cyclic compounds substantially41. Phosphonates as 5 can (see Figure 1.6.), after attack of

water and prQton transfer to the axial oxygen, isomerize to either an intermediate with the ring carbon in an axial

position (normally not occupied by electron donating groups), or to an intermediate with a di-equatorial five-membered

ring, increasing the ring strain. These processes are

energetically unfavourable and can not compete with ring

opening, so no exocyclic hydralysis is found with the phosphonate .•

More recently, many other stereochemical and kinetic data in phosphorylation reactions and in group transfer reactions

have been rationalized by invoking phosphorane intermediates, including the group transfer reactions in tricyclic 'caged' phosphatranes by Van Aken et al~2,43.

18

1.4. Transfer of conformational changes in a phosphate

group to the hydrophobic part of organic molecules.

In the phosphorylation reactions of Westheimer, substitution

was achieved at phosphorus, whereby both incoming and leaving

groups entered and departed via the axis of an P(V) TBP. In

biochemistry a lot of phosphate groups are temporarily

activated without breaking the honds with the rest of the

biomolecule. In this case incoming and leaving group are the

same molecule, e.g. water. Dependent on the life time of the P(V) TBP intermediate the activated system will be able to

relax to a lower energy state. Another configuration will he

occupied, that is better able to accommodate the new charge

distribution over the ligands around phosphorus. Usually this

will happen by turning away one electronegative part of the

molecule from the other. Theoretical verification of charge

0 11

. P, +LH Ho~;! 'o- ....

Hs';.;.<'Os·,.... Hs"~jo

H~,·

net atomie P(IV)

charge ( e. u.)

0 ( 1 I) -0.267

0(5') -0.289 p +0.370

L

...

P(V)H P(V)

r.=HNMe L=OH L=HNMe L=OH

-0.274 -0.274 -0.291 -0.296

-0.316 -0.318 -0.348 -0.355

+0.340 +0.348 +0.409 +0. 451

-0.031 -0.133 -0.320 -0.436

1.10. Charge distribution on the ligands of a tetrahydrofurfuryl

model system ~hen going from a four to a five co-ordinated intermediate;

CND0-2 optimized results.

19

enhancement on apical ligands in a P(V) TBP model compound for DNA was recently published by van Lier et alJ1. The net atomie charge on various atoms in the tetrahydrofurfuryl model system and its P(V) TBP counterpart are given in Figure !.10. Experimental evidence for the rotation, resulting from

this charge enhancement, was very recently given by Koole et alJ2. They synthesized a number of four and five co-ordinated

mutually resembling phosphorus model compounds and found a significantly greater population of the gauche(-) conforma­

tion for axially situated tetrahydrofurfuryls around the C4'-Cs' bond in the 5' P(V) TBP tetrahydrofurfuryls with respecttotheir related P(IV) compounds. In Figure !.11.,

some of the model compounds used are given12.

L:: Ph,OEt Y:O,CH2

L:Ph,OEt

Y = O,CH2

Figure I.ll. Four and five ao-ordinated model compounds with different Zi­

gands substituted on the phosphate group and in the tetrahydPofUrfuryZ

ring.

The Newman projectionsof the rotamers around the C4'-Cs' bond are given in Figure I.12. The rotamer populations x(g+), x(gt) and x(g-) could be determined from the time-averaged

coupling-constants Ja4'as' and Ja4'as"12. In the P(IV) compounds Os' and Y are oriented cis to each other (the gauche(+) or the gauche(t) rotamer) in the case Y = o, due to the gauche effect44. This effect is defined45 as the "tendency to adopt that structure which has the maximum

20

Os· H s" Hs·

Y*C, Y*C,• Y*C,· Hs. Hs" Os• H5• H5 0 5 •

H •. H •• H ••

g• gt g-

I.l2. Newman around the P-0-C-C-0 sequence.

number of gauche interactions between the adjacent electron

pairs and/or polar bands" and originates from bond-antibond

interactions45. The only compound that differs substantially

possesses a CH2 group on the Y position (see Figure !.11.)

and the C4'-Cs' rotaroer distribution is consequently not

dominated by the gauche effect. In the P(V) compounds Os' and

01' are orientated more transtoeach other, i.e. the g

population is enhanced, due to the repulsion of two more

negatively charged oxygens. Only the compounds with Y = CH2

show no difference in population with the four co-ordinated

intermediate, which is clearly the consequence of the

P-0-C-C-C sequence present in the molecule, instead of the

P-0-C-C-O sequence, that is responsible for repulsion.

21

References and notes.

1. R. Harrison and G.G. Lunt, 'Biological Membranes, Their

Structure and Function', Blackie, Glasgow, 1980; p 62. 2. S.J. Singer and G.L. Nicolson, Science 1972, 175,

720-731. 3. A. Tardieu, V. Luzzati and F.C. Reman, J. Mol. Biol.

1973, 75, 711-733. 4. F.T. Presti, R.J. Pace and S.I. Chan, Biochemietry 1982,

21, 3831-3835. 5. A.G. Lee, Biochim. Biophys. Acta 1977, 472, 237-281.

6. A. Blume, R.J. Wittebort, S.K. Das Gupta and R.G.

Griffin, Biochemietry 1982, 21, 6243-6253. 7. In reference 1, p 119.

8. In ref~ence 1, p 12. 9. In refere~e l, p 87.

10. In reference 1, p 86. 11. J.J.C. van Lier, L.B. Koole and H.M. Buck, Reel. Trav.

Chim. Pays-Bas 1983, 102, 148-154.

12. L.B. Koole, E.J. Lanters and H.M. Buck, J. Am. Chem. Soc. 1984, 106, 5451-5457.

13. For up-to-date reviews on the subject, see the series 'Organophosphorus Chemistry' (Specialist Periodical

Reports), S. Trippett, ed., The Chemical Society,

London. 14. F. Covitz and F.H. Westheimer, J. Am. Chem. Soc. 1963,

85, 1773-1777.

15. E.A. Dennis and F.H. Westheimer, J. Am. Chem. Soc. 1966, 88, 3431-3433.

16. F.H. Westheimer, Acc. Chem. Res. 1968, 1, 70-78. 17. R. Kluger, F. Covitz, E. Dennis, L.D. Williams and F.H.

Westheimer, J. Am. Chem. Soc. 1969, 91, 6066-6072.

18. R. Luckenbach, 'Dynamic Stereochemietry of Pentaco-ordinated Phosphorus and Related Elements', G. Thieme, Stuttgart, 1973.

19. B.L. Muetterties and R.A. Schunn, Quart. Rev. Chem. Soc. 1966, 20, 245-299.

22

20. T.E. Clark, R.O. Day and R.R. Holmes, Inorg. Chem. 1979,

18, 1653-1659.

21. T.E. Clark, R.O. Day and R.R. Holmes, Inorg. Chem. 1979,

18, 1668-1674.

22. R.F. Hudson and M. Green, Angew. Chem. 1963, 75, 47-56.

23. A.J. Kirby and S.G. Warren, 'The Organic Chemistry of

Phosphorus', Elsevier Publ. Co., Amsterdam, 1967.

24. F. Ramirez, S. Pfohl, E.A. Tsolis, J.F. Pilot, C.P.S.

Smith, I. Ogi, D. Marquarding, P. Gillespie and P.

Hoffman, Phosphorus, 1971, 1, 1-16.

25. O.A. Bochvar, N.P. Gambaryan and L.M. Epshtein, Russ.

Chem. Rev. 1976, 45, 660-670.

26. T.A. Halgren, L.O. Brown, D.A. Kleier and W.N. Lipscomb,

J. Am. Chem. Soc. 1977, 99, 6793-6806.

27. M.A. Ratner and J.R. Sabin, J. Am. Chem. Soc. 1977, 99,

3954-3960.

28. J.H.H. Hamerlinck, Ph. D. Thesis, Eindhoven Oniversity

of Technology, 1982.

29. J.H.H. Hamerlinck, P. Schipper and H.M. Buck, J. Org.

Chem. 1983, 48, 306-308.

30. J.H.H. Hamerlinck, P. Schipper and H.M. Buck, J. Am.

Chem. Soc. 1983, 105, 385-395.

31. P. Gillespie, P. Hoffmann, H. Klusacek, D. Marquarding,

s. Pfohl, F. Ramirez, E.A. Tsolis and I. Ugi, Angew.

Chem. 1971, 83, 691-721.

32. E.L. Muetterties, W. Mahler and R. Schmutzler, Inorg.

Chem., 1963, 2, 613-618.

33. E.L. Muetterties, K.J. Packer and R. Schmutzler, Inorg.

Chem. 1964, 3,1298-1303.

34. D. Marquarding, F. Ramirez, I. Ogi and P. Gillespie,

Angew. Chem. 1964, 85, 99-127.

35. F. Keiland w. Kutzelnigg, J. Am. Chem. Soc. 1975, 97,

3623-3632.

36. E.L. Muetterties, J. Am. Chem. Soc. 1969, 91, 1636-1643.

37. E.L. Muetterties, J. Am. Chem. Soc. 1969, 91 , 4115-4122.

38. J. I. Musher, J. Chem. Educ. 1974, 51 , 94-97.

39. R.S. Berry, J. Chem. Phys. 1960, 32, 933-938.

23

40. I. Ugi and F. Ramirez, Chem. Br. 1972, 8, 198-210.

41. J.A. Gerlt, F.H. Westheiroer and J.M. Sturtevant, J. Biol. Chem. 1975, 250, 5059-5067.

42. D. van Aken, l.I. Merkelbach, A.S. Koster and H.M. Buck, J. Chem. Soc., Chem. Comm., 1980, 1045-1046.

43. D. van Aken, l.I. Merkelbach, J.H.H. Hamerlinck, P. Schipper and H.M. Buck, A.C.S. Symp. Ser., 1981, 171,

439-442. 44. s. Wolfe, Acc. Chem. Res. 1972, 5, 102-111.

45. T.K. Brunck and F. Weinhold, J. Am. Chem. Soc. 1979, 101, 1700-1709.

24

II. Sodium ion transport across biologica! membranes.

I I. 1. Theory.

II.l.l. Cluster formation in biologica! membranes induced

via a phospholipid P{V) TBP intermediate.

In Chapter I the principle of induction of an electron flux

into the axis of a five co-ordinated phosphorus (P(V))

trigonal bipyramidal (TBP) intermediate was discussed. In

those model compounds, the extra negative charge on the

axial oxygen (Os') in the P(V) TBP intermediate resulted in

repulsion of another oxygen (01 '), bound via an o-e-c-o sequence to the phosphate group, and located in a tetrahydro­

furfuryl ring (see Figure II.l(a)).

-(a)

I. .1'

I

' I bl ' .1•

' Figure II.l. Repulsion between the two oxygens in a P-0-c-c-o seque~e as a aonsequenoe of thè transition from a four ao-ordinated to a five

ao-ordinated intermediate in (a) a tetrahydrofurfurylphosphate and (b)

a phospholipid.

25

In this Chapter, an attempt will be made to show that the same process can occur in phospholipids, and that this process could be the 'trigger' to activate proteins, such as ion channels, emeedded in a lipid bilayer matrix via uptakè

in clusters 1. In Figure II •. l (b) one can distinguish the same o-e-c-o sequence, bound at the axial position of a phosphate group, as discussed for the model compounds in Chapter I.

Repulsion between the two oxygens 021 and 0312 of the P-0-C-C-0 sequence will cause a shift of the sn-2 chain in the direction perpendicular to the bilayer surface (see Figure II.1.(b)). However, the model compound& as (a) in Figure II.1. are monomers, dissolved in organic solvents, and thus able to re-orientate freely in solution. The phospho­lipids, on the contrary, are built in in the lipid bilayer, with their long hydracarbon chains interacting via 'van der Waals' interactions with the neighbouring chains. So the shift of the hydracarbon chains along each other will normally take a high energy barrier to overcome. A plausible adaptation of the bilayer by which this process can be aided,

is accompanied by a change in the angle of tilt of the hydra­carbon chains to the bilayer normal. Phase diagrams of phospholipids show, dependent on temperature and percentage water or different lipid, several one and two phase regions3, in which among others the angle of tilt to the bilayer normal varies. In a special temperature interval, the phase transi­tion region, ranging from the main phase transition tempera­ture down to a temperature around the pretransition4,5,

smal! domains of different density (and thus different angle of tilt) occur next to each other. Such clusters are reported for mixtures of phosphatidylcholines (PC) with phosphatidyl­ethanolamines (PE)3, cholesterol6 or proteins7. The co-oper­

ative change in the angle of tilt of all the lipid molecules in the same cluster, can minimize the energy barrier that has to be overcome. Although there is controversy about the exact nature of the pretransition, a continuous change in the angle of tilt is always included in the description. Some authors conclude that the angle of tilt will change from tilted at

26

the pretranaition temperature to parallel to the bilayer nor­

mal at the main phase transition8. Others believe the angle

of tilt reaches a local minimum at the pretransition, accom­

panied by a transition from a tilted conformation, via a

tilted and rippled two dimensional structure9 to a one dimen­

sional structure with the chains perpendicular to the bilayer

surface. Experimental evidence for a variation in angle of

tilt of the hydracarbon chains accompanying hydracarbon chain

shift was given by Blume3 and Chen10. Comparison of 13c and

2H NMR spectra of phospholipids, labelled respectively with

13c at the sn-2 carbonyl group3 and with 2H at the 4-position

of the same sn-2 chain, suggests that a conformational change

of the carbonylgroup precedes chain melting on increasing

temperature. This could be an indication of constantly

developing P(V) TBP intermediates, meeting below the pretran-

lipid

MMPC

MPPc(a)

MSPc(b)

PMPC

PPPC

PSPC

SMPC

SPPC

SSPC

ma in transition

temperature

23.6

35.1

38.6

27.3

41.1

49.0

29.4

43.9

54.2

pretransition

temper at ure

14.4

22.8

10.8

34.8

39.9

20.0

30.8

50.4

(a) MPPC = a myristoyl chain (M) bound at the sn-1 position

and a palmitoyl chain (P) bound at the sn-2 position of

the phosphatidylcholine (PC) glycerol backbone.

(b) S = a stearoyl chain.

Table II.l. Main transition and pretransition temperature as a function

of the chain Zength of the chain at the sn-1 and the sn-2 position of

phosphatidyZchoZine.

27

sition temperaturel an energy barrier too high to transmit the repulsion between the two oxygens of the o-e-c-o sequence in a shift of the sn-2 chain. The activation around the pre­

transition temperature is high enough, howe.ver, to induce a conformational change at the carbonyl-group next to 021•

At higher temperatures an ever greater part of one and the same sn-2 chain and/or an ever greater part of the total

number of sn-2 ebains will be aetivated, untill at the main phase transition all the ebains are oriented perpendieular to

the membrane surfaee. Moreover, the pretransition behaviour shows strong dependenee on eomposition10. PCs with myristoyl

(C14), palmitoyl (C16) and stearoyl ebains (C1s> at the sn-1 and/or sn-2 position meet a higher energy barrier to melt if the sn-2 chain is longer (see Table 11.1). Another environ­mental constraint is met in the headgroup region (see Figure II.2.).

---

H H ....,_./ 0

-o 1/,,,, •.. ~ -o

-o ""=-k-J-,. 06 -;:N

/ 3

~ =

~ l L1'

I

' Figure II.2. Ringformation in the phosphatiQY~eho~ine headgroup upon

formation of a five ao-oPdinated intermediate.

In the headgroup, accommodation of a fifth ligand to form a five eo-ordinated intermediate will cause re-orientation of

28

the ligands around phosphorus. A lot of phospholipid

molecules contain zwitterionic headgroups (see Figure II.3.),

that are arranged with alternating charge in the bilayer11.

This model of intermolecular interaction between e.g. the

N-methyl protons of one PC molecule and the phosphate of a

neighbouring PC molecule, however, still allows for

considerable freedom of movement about the various honds in

the headgroup11.

phosphatidyl­

choline

II.3.

sphingomyelin phosphatidyl- phosphatidyl-

ethanolamine serine

w·ith a zwittericm:c

Thus formation of a five co-ordinated phosphorus intermediate

must be accompanied by re-organization of the charge in the

total bilayer. This can occur, for example, by intrarnolecular

compensation of the charge12, thus creating a more or less

neutral molecule, or, at a physiological level, by de- and

adsorbtion of mono- and di-valent cations13,14.

Intramolecular compensation of charge can be established by

pseudo-ringformation, in which positively and negatively

charged groups are brought close to each other12 (see Figure

II.4.). Here, another aspect of the pretransition behaviour

is met, the relative cross-sectional areas of headgroup and

chain region. In molecules as PC, the headgroup in

'stretched' P(IV) conformation, will occupy a greater

29

excluded cross-sectional area tban the lipid bydrocarbon

cbains. Tbe ebains adopt a tilted conformation to fill .in a potential void in tbe bydrocarbon cbain region6. Above tbe

pretransition temperature, an increasing number of ebains orient more perpendicular to the membrane surface8,9, so

that the excluded cross-sectional area of tbe .headgroup must have been diminished. Tbis is confirmed by tbe observationlS

tbat PE, N-methyl and N,N-dimethyl PE do not exhibit such

pretransition behaviour, since the cross-sectional areas of

their headgroups are smaller. The effective cross-sectional area of the PC headgroup can be diminished further, after

di-equatorial16 pseudo-ringformation and pass down of the

charge, by pseudo-rotation, through which the pseudo six­

membered ring will be orientated temporarily axial-equator­ial. This pseudorotated structure will, at the same time,

prevent electron back donation to the fifth ligand, e.g.

water, and return of the intermediate to the four co-ordi­nated state. De- and adsorption of mono- and divalent cations can complete the picture sketched above. During the physio­

logical process of the excitation of an axon, for example, momentary desorption of ca2+ ions from the outer monolayer of the membrane is reported13,14 upon activation of the axon, followed by adsorption of monovalent ions as Na+. This decrease in positive surface-bound charge of cations can

dfminish repulsive forces, intended to keep the headgroup 'stretched' in the unactivated axon17.

Finally, again in the example of the excitation of the axon, the potential at rest is negative inside the axon, pulling the positively charged end of the stretched headgroup

-N(CHJ)3+ of the outer monolayer phospholipids into the membranes. During activation of the axon, a positive poten­tial inside17 will push the positive charged end of the head­

group outwards, enabling the headgroup to re-orientate. Thus, a set of environmental physiological conditions is realized, enabling a P(V) intermediate to develop and to pass the

information, stored in its renewed charge-distribution, down to the hydracarbon region, for the case of the motionally

30

restricted phospholipid molecules. On a molecular level this

processcan be summarized as follows (see Figure II.4.).

ca2+ ions desorb, and are replaced by Na+ ions. The internal

potential is changing from negative to positive. As a conse­

quence the headgroup is no langer forced in an extended and

inward pulled conformation and gets the opportunity to

re-orientate. The always existing P(IV) ~ P(V) equilibrium,

under 'resting' conditions laying at the side of the P(IV)

Figure II.4. The in the physioîogicaî conditions accompanying

the transition of a four to a f"ÎVe co-oràinated intermediate.

compounds, will be shifted in the direction of the P(V)

intermediate, since this conformation now is stabilized by

the formation of a di-equatorial16 pseudo-six roerobered ring.

The positively charged nitrogen of the choline headgroup

shields the negative charge of the formerly double bonded

oxygen, thus polarizing the P=O bond, by which the electro­

philicity of the phosphorus atom will be increased18. This

process will be promoted by nucleophilic attack of e.g. a

water molecule19, normally present in excess in the headgroup

31

layer of the membrane, thus generating the P(V) TBP inter­

mediate. The decrease in cross-sectional area of the head­group due to ring formation, will cause decrease in the angle

of tilt of the hydracarbon chains. The increased electron density on the axial oxygen of the phosphate group will

induce repulsion of 021• bound via an o-e-c-o sequence to the phosphate group. This process will aid, or maybe it is the

main cause, of a further decrease in the angle of tilt of the hydracarbon chains, through a 'shift' of the hydracarbon chains along each other. The difference in effective chain length10, 20-22 is diminished. Co-operative change in the

angle of tilt of a number of phospholipid molecules is needed

to maximize the 'van der Waals' interactions between neigh­

bouring acyl chains. This will lead at a macromolecular level to formation of a cluster with an average angle of tilt differing from the surrounding matrix. The relaxation time of

such a cluster is appreciably greater than for other charac­

teristic movements of the molecule23. So the short-living P(V) TBP intermediate initiatea the formation of a cluster

with a much longer life-time, through which a time scale can be reached at which physiological processes can take place23. Although the physiological conditions of the above process are borrowed from the excitation of an axon, one can imagine

the same conditions for other membranes. Over most membranes an ion-gradient is maintained by ion pumps, so a potentlal exists over most membranes. Divalent ions as ca2+ and Mg2+

are bound to most membrane surfaces to an extent dependent on

the physical state of the lipids23. Mostly they are bound more strongly than monovalent ions such as Na+ and K+, that are present in all intra and extra-cellular spaces. The local oircumstances may change, a P(V) TBP intermediate can be

built up under several sets of conditions.

32

11.1.2. Activation of membrane proteins by uptake in

phospholipid clusters

A break in the Arrhenius plot of ATP-ase activity versus the

reciprocal temperature has been reported in dioleoyllecithin

substituted ATP-ase, when reaching the temperature of cluster formation24. A transition in the temperature dependenee çf

ca2+ accumulation25 in Sarcoplasmic reticulum membranes is

attributed to a change in entropy of activatien rather than

to the free energy of activation. Results that are consistent with an order-disorder transition invalving the lipid alkyl

chains25. Computer simulation studies concerning phase

separation in lipid bilayers containing integral proteins7,

show a system that separates into an essentially pure lipid

phase and a protein-rich phase containing melted lipids

between Tk and Tc· Here Tk is the melting point of clusters,

and Tc is the {main) melting point of lipids. Addition of

oleic acid to a lipid deficient membrane26 produces a fluid

membrane structure, which is most likely an essential

requirement for the reconstitution of the calcium dependent

ATP-ase activity. Addition of stearic acid, on the contrary,

has no activating effect on the calcium dependent ATP-ase26

and creates a gel-like lipid structure. From the above­

mentioned and other27-29 articles, it becomes clear that a

certain fluidity is essential for the activatien of membrane

proteins, and that this fluidity is reached in a temperature

range in which cluster formation appears.

An often encountered objection against the relation between

cluster-formation and protein activation is, that gel state

lipids do not appear to be present in most biologica! membranes30. However, this is only partly true. Harrison and

Lunt conclude31 that, although hydracarbon ebains in natural

membranes are believed to be generally in a fluid state at

physiological temperatures, the presence of sterals and

proteins may lead to local variation of the mobility in the

membrane. Moreover, the degree of lateral phase separation is

believed to be influenced by a number of external factors,

33

e.g. water content31, proton and cation concentration32, ionic strength32 and potential fiela19. Under influence o.f the above-mentioned factors, cluster-formation is believed to persist far above the main phase transition temperature of a pure lipid mixture. Also pore-mediated ion transport shows some peculiar characteristics while planar bilayer membranes pass the phase transition region o.n heating33. Planar bilayers consisting of mixed-chain lipids and modified by pore-forming antibiotica as Gramicidin A, do not show any peculiar effect on Tc, the main phase transition temperature (29°C). Bowever, at 22-23°C a pronounced maximum in pore-induced conductance is seen. The effects observed are interpreted in terros of lateral -phase separation into pure lipid and lipid-antibiotic domains33. Consequently, the polypeptide Gramicidin A is an ideal model for the ion-channel forming proteins, obeying the same temperature dependenee of protein activation upon cluster formation. A schematic representation of Gramicidin A34 is given in Figure II.5.

0 N

0 0 • H

...,.. H- bond

Figure II.5. The Gramicidin channel oomprises ~o polypeptide ahains in

B -helix fol'm.

34

On a molecular level, protein activation upon uptake of the

protein in a cluster, can be described as follows. An

integral protein in a fluid environment (the cluster} could

be free to adopt the tertiary structure necessary to function

as an ion-channel. A gel-like matrix can displace some

special group of the channel-forming protein out of its

critical position35 and/or disturb a protein in helix form36.

For the sodium ion-channels of myelinated axons a model is given in which four energy harriers in the pore cernprise the

selectivity filter in the ion-channet35. These four harriers

consist of dehydration and hydration steps, enabling a

partially dehydrated sodium ion37 to pass a narrow gap

besides a strongly co-ordinating carboxylic acid group38.

Completely hydrated, the sodium ion will not be able to pass

the narrow selectivity filter of 3 x 5 Ä. A very small

displacement of the carboxylic acid group can make the ion­

channels impermeable.

2

Figure II.6, The

of the ion channeL.

In

energy harriers comprising the selectivity filter

Although the correlation between a fluid lipid environment

and activation of proteins is suggested by quite a number of

authors (vide supra), this is not necessarily a general

35

principle. Also the reverse process, i.e. protein activation

upon uptake in a gel-like cluster, is a process that should not be neglected. In this way a number of different proteins

could be activated in succession, if e.g. the potentlal is

constantly changing from negative to positive and a whole

scala of states of different rigidity is passed through. Finally, support for the above-mentioned theory is given by

the fact, that 2-amido PC, contrary to PC, is found to be inhibitory for integral proteins39,40 (see Figure II.7.).

The oxygen esterified to c2 of the glycerol backbone of PC is essential for the transfer of conformational change in the

headgroup towards the change of tilt of the hydracarbon

chains. If this oxygen is replaced by the less electro­

negative nitrogen of 2-amido PC, less repulsion and resulting acyl chain shift will be expected. Moreover, the hydrogen bridge found in X-ray analyses of comparable lipids41, will hinder acyl chain shift and headgroup re-orientation (Figure

11.7.).

)(

Figure II.7. Hindered repulsion and aeyZ ahain shift in 2-amido phospha­

tidyZehoZine.

36

II.2. Experiments.

11.2.1. Introduction.

The characteristic feature of the model, described in this

Chapter, is that changes in the lipid environment of a

protein are the 'trigger' for the protein to be activated (to

place some particular functional group just in or just out of

the right position). Other authors propose a mechanism in

which changes in membrane potential, pH, ionic strength etc.,

directly influence the channel-forming protein (in the case

of ion-transport) to open, a mechanism that developed under

the influence of experiments with the ion channel blockers

tetrodotoxin42 and saxitoxin43.

To get a more decisive answer about this question, the

experimental conditions of the investigations described below

are chosen so, that only the lipid composition of the mem­

branes varies, leaving the concentratien of ion channels,

ions, probes and buffers, as well as temperature, as constant

as possible. Vesicles are formed with a diameter of approxi­

mately 1000 Á, their wall existing of one double layer of

lipids. As a reference, vesicles of egg yolk lecithin are

chosen, to which 10 to 50% of synthetic or natural, specific

lipid is added, to vary the total lipid composition. Attempts

to make vesicles of one well-defined synthetic lipid as

reference, failed, since the temperature of formation of the

vesicles had to be above the main phase transition tempera­

ture, Tc. Tc will vary from 24•c for dimyristoyl (C14l, via

41•c for dipalmitoyl (C16l to ss•c for distearoyl (C1al phos­

phatidylcholine44. The last two temperatures where too high

to be constantly maintained throughout the whole procedure of

synthesis. Since Tc of egg yolk lecithin is around o•c, the

choice of this lipid experimentally gave no problems.

As model for the ion-channel protein, the ionophore

Gramicidin A was chosen for a number of reasons, in addition

tothese mentioned in Chapter 11.1.2. Gramicidin A is a pore­

former, specific for sodium ions. 1t is a pore-former and not

37

a carrier, so it builds up a permanent channel comparable to

natura! ion-channels, and does not diffuse through the membrane as carriers do. A pore-former also functions below

the main phase transition temperature so that the temperature range in which it is active is greater, while the conductance

of carriers falls below Tc to the state of bare membrane conductance (not pore mediated)33,45. The spontaneous

current fluctuations observed with unmodified planar bilayers

near the lipid phase transition temperature, containing a few

molecules of Gramicidin A, reminds of the idea of 'cluster­activation' of channela46, i.e. Gramicidin A is activated if

it is taken up in such a cluster. Formation of a cluster

around a Gramicidin A molecule activatea the channel to open.

The channel stays open during the life-time of a cluster, that can change randomly, but the conductance reached is always the same, unless the cluster will decay before the maximum conductance is reached. A Gramicidin A channel is formed by association of two poly­petides at their N-formyl ends. Each chain is folded into a B

helix, which resembles a rolled-up B pleated sheet (see (a) in Figure 11.8.).

Figure II. 8. Four possib Ze confoi'mations of t;he Gramicidin channe L ,- -~

Finally, Gramicidin A incorporates spontaneously in the veeiele wall after addition, since its amino acid sequence is one of the most hydrophobic ones known47.

38

' " (el 1=2410 (I) I 11

1=11.0 00

' " [d) t=13 30

(k)

I 11

(c) 1=458

( j)

(i)

( b) I 11

t = 0 00

( h)

(g)

(a) I 11

t = 0 00 ( f )

Figure II.9. 23Na Nii!R spectra (79.4 MHz) of a dispersion of vesicles of

egg yolk phosphatidylcholine plus 10% sphingomyelin (a) at t=O before

addition of the probe, (b) at t=O after addition of the probe, (c)-(l)

after addition of Gramicidin; T = 25°C.

39

Moreover the exterior of the channel consists again of the

most hydrophobic parts of the peptide48, thus creating an oxygen lined, probably water-filled, central channel.

Gramicidin A is added in last resort to a buffered solution with approximately 1000 A diameter vesicles, around which the

NaCl solution is replaced by LiCl by means of ultrafiltra­tion. Thus a sodium ion gradient is formed of about 102, the

equilibration of which is started at the moment Gramicidin A is added. A shift reagent49, Dy I N(CH2C02)3J23- has been used

to distinguish between 23Na+ inside and outside the vesicles.

II.2.2. Results.

A homogeneous suspension of 1000 Ä diameter, single bilayer

phospholipid vesicles is prepared by the procedure of Enoch

and Strittmatter50. The solubility and nature of the surfac­tant51 deoxycholate, together with its concentration50 (a phospholipid to deoxycholate molar ratio of 2:1 is used),

determine the diameter of the resulting vesicles. After formation of the vesicles, the detergent is removed to make

the vesicles impermeable to sodium ions, next external NaCl is replaced by LiCl by means of succeeding ultrafiltration

steps. After this stage there is 100 mM NaCl inside the

vesicles and the aqueous space outside the vesicles is 50 mM in LiCl and < 1 mM in NaCl. The 23Na NMR spectrum (79,4 MHz, Bruker CXP-300) of a dispersion of vesicles consisting of egg

yolk lecithin +10% sphingomyelin is given in Figure II.9.a,

the single sharp resonance at 1802 Hz ± 2 Bz repreaenting both the Na+ inside and outside the vesicles. Fig. II.9.b shows the spectrum after the outside aqueous space is made

3 mM in triethanolamine dysprosium nitrilotriacetate

[HN(CH2CH20H)3)3Dy[N(CH2C02)3]2, according to the metbod of Pike et al49. The single resonance of Figure II.9.a is split into two peaks, the smaller one49 65 ± 10 Hz upfield from the

peak position of Figure II.9.a repreaenting Na+ outside the

vesicle and the large one 4 ± 1 Hz downfield repreaenting Na+

40

1.0

PC trom

0.9 egg yol k

0.8 6 + 50 •;. OMPC 7 • so·~ OPPC

0.7

R + , oo;. OPPC

1

5 . 10% OSPC 0.6

3 • 10% OMPC 0.5

0.4

2 + 10% sphingo-mye I in

0.3 9 •10% P inOSitOI 10 + 1 0°/o OPPE

PC pure

fro m egg yol k

0.2

0.1

8 + 10o/o P serine 6

3 5

-tfhrl

Figure II.10. The time of the re~ative integra~s of the two

peaks of Figure II.9. ten different differing in ~ipid com-

position of the vesicle ceZZ-waZl.

41

inside, the absolute magnitude of the shift being dependent

on the exact concentratien of the probe49. The fraction of

the total integral, due to the inside resonance varied from

0.77 to 0.92 for the various samples. The value for the

fraction of the total aqueous volume inside the vesicles is

calculated, assuming an internal volume of 2960 ml per mole of lipid52 and a final lipid concentratien of 10 mM, to be

0.05. There could be some error in this ratio because of inaccuracy in the knowledge of the final lipid concentration.

Bowever, other studies using the method of Enoch and Stritt­matterSO, where phosphate analyses was conducted on the final

solutions, were in good agreement with the above-mentioned

fraction of lipid left after synthesis of the vesicles.

Combination of these numbers with 100 mM Na+in• yields a Na+out concentratien of 0.27 mM in the case the fraction of the total integral due to inside resonance at t = 0 is 0.92

or 0.92 mM in the case this fraction is 0.77. Since any

1eakiness of the vesicles would affect the observed ratio, all starting spectra were obtained between 0.5 and 3 hrs

after the last ultrafiltratien step, and samples with a

fraction due to inside resonance smaller than 0.77 were not used further. E.g. for the sample of egg yolk lecithin with 50% DSPC it was not possible, even after several attempts, to

obtain a sample with a fraction due to internal resonance greater than 0.30. Immediately after the spectrum of sample + probe was obtained, the salution was made 0.04 pM in the ionophore Gramicidin53-55. This amounts to ca. 3 Gramicidin

channels per vesicle49 and induces a rapid efflux of Na+ down

its ccncentration gradient, as can be seen in the Figures

II.9.c-l. They depiet some of the spectra obtained and show the time evolution of the spectrum measured in minutes from

the time of addition of Gramicidin, for a sample displaying

an intermediate time course in the total series measured (egg

yolk + 10% spingomyelin). The spectra given are power spectra in which the integral is proportional to the square of the number of the sodium nuclei56. The power spectra are taken to

cancel the influence of the phase-correction, that will vary

42

with time during the period the automated measurements are

recorded. A plot of the logarithm of ratio R, of the frac­

tional integral (inner/total) at t = t to the fractional

integral at t = 0, against time is shown in Figure II.10. The

time during which the spectra were recorded after addition of

Gramicidin varied from 2 to 13 hrs, dependent on how fast the

ion transport took place. For 50% DPPC not the total time­

course is given, since between 5 and 13 hrs after addition of

Gramicidin the slope of the plot was identical to that of the

part shown between 0 and 5 hrs.

II.2.3. Discussion.

The efflux of Na+ ions shows at least two stages49. Directly

after addition of Gramicidin a passive one-for-one Na+ for

Li+ exchange out of and into the vesicles takes place, both

ions moving down their concentratien gradients, although the

cl- transport may play a role at this stage (vide infra}.

This stage will end when the Li+ gradient is dissipated

(47.5 mM both inside and outside). Since the Na+ gradient

still exists at this point (52.5 mM inside, : 3 mM outside,

R = 0.52), during the second stage the Na+ gradient will be

further dissipated at the expense of creating a new Li+

gradient, still through a one-for-one exchange. This stage

will end when the ratio of inside concentratien to outside concentratien has the same value for both Na+ and Li+, so

that the same diffusion potential for both ions is reached.

Here Na+ in is 10 mM, Na+ out is 5.3 mM, Li+ in is 90 mM

and Li+ out is 45.3 mM, the fraction of total Na+ inside is

0.09. True equilibrium will only be obtained after a third

stage, namely passive nonfacilitated cl- transport out of the

vesicles. This stage will end when the fraction of total Na+

inside is equal to the analogous volume fraction, 0.05. In

these rough calculations osmotic swelling, a possible pH

gradient due to permeation of the counter ion of the probe

HN(CH2CH20H)3+ and the Donnan effect caused by the impermeant

43

Dy[N(CH2C02)3l23- ar~ ignorea50. The transitions of the

curves, shown in Figure II.10, are located between R = 0.5 and R = 0.3, the transitions being at lower R if ion trans­

port is faster. This could be an indication of leakage before

Gramicidin is added, although measurements of ten different

blancos (after addition of the probe) of the same sample in a time period of half an hour showed no significant decrease of

the ratio of Na+in to Na+out• The fast process between R = 1 and R = 0.5, corresponding to

the one-for-one exchange of Na+ and Li+ down to their gradient, is believed to be limited by the Gramicidin induced Li+ transport, which is ca. 1/6th as fast as that of Na+ 57.

Thus the slow process, below ca. R = 0.5, would correspond to the essentially simultaneous occurrence of the second and third stages (vide supra) implying that they have very simi­lar permeability constants. This is supported by measurements

of permeability coefficients of pass!ve nonfacilitated trans~ port of cl- 52. So both stages befare and after R = 0.5, are involved in sodium ion transport.

As can beseen in Figure II.10, the relative sequence in velocity of ion transport for the various samples is not

interchanged when passing R = 0.5. The reference sample, phosphatidylcholine from egg yolk

exists of predominantly C16 (34.3%) and C1a chains (59.8%

C18:0• C18:1 and c18:2)58, as can beseen in Table II.2. Ion transport over a vesicle cell wall consisting of egg yolk phosphatidylcholine is relatively fast, compared to

most of the other samples (2-10 in Figure II.10.). Clearly, the relative quantities of different chain lengtbs and

saturation is such, that the rigidity of the matrix is ideal to accommodate for the chain shift that results from the

formation of a five co-ordinated phosphorus intermediate. Increasing the percentage of synthetic saturated ebains

(samples 3-6), the matrix will adopt a more gel-like charac­ter, in which the packing of the optimally ordered ebains is

tighter, so the chain shift is more difficult. When increas­ing the chain lengtbs of the saturated chains, ion transport

44

becomes slower, although the difference between ~he sample

with 10% DSPC and with 10% DPPC is too smal!, compared to the

deviation in slope due to measuring faults, to be called

significant. Addition of more (50%)'of the same saturated

chains again gives slower ion transport, in which the slopes

of the plots of 50% DMPC and 50% DPPC are indistinguisable.

Addition of 10% of sphingomyelin from egg yolk, containing

primarily saturated palmitic acid chains at the sn-2 position

results in ion transport that is slower than that of the

reference sample of phosphatidylcholine from egg yolk (1),

but faster than that of the samples 3-6. The slower ion

transport of sample 2 compared to sample 1 could stem from

the restricted repulsion of the sn-2 nitrogen (see Figure

II.11.) through which the sn-1 chain is bound to the glycerol

backbone. The significantly faster ion transport compared to

the samples 3-6 can be explained by the fact that the sn-1

chain is significantly shorter (two atoms less than in DPPC,

viz. -CH=CH-(CH2l12-CH3 directly bound to the glycerol back­

bene, and the appearance of a double bond), thus creating a

molecule comparable to MPPC (see Table II.1.).

Moreover, addition of a clearly different lipid will enhance

the heterogenity of the matrix, making the matrix more fluid,

and enabling lipids other than sphingomyelin to translate a

five co-ordinated phosphorus intermediate more easily in a

hydrocarbon chain shift.

Samples 8, 9 and 10 contain additions with primarily C16 or

C18:0 to C18:2 (see Table II.2.) as lipid alkyl chains,

thus changing the overall fluidity of the matrix not too

much, compared to 1, on additions of 10%. Egg yolk phosphati­

dylcholine with 10% phosphatidylserine shows considerably

faster ion transport compared to the reference sample. The

availibility of two ligands within the serine, one as fifth

ligand to build up a five co-ordinated intermediate and one

to polarize the original P = 0 bond, could cause an appreci­

ably increased life-time of the five co-ordinated inter­

mediate, which explains the fast ion transport observed (see

Figure I I. 12.).

45

chain length phosphatidyl- sphingomyelin phosphatidyl- phosphatidyl- phosphatidyl-

and choline S-0756 ethanolamine inositol serine P-8518

saturation P-6013 P-4513 P-5766

C14:0 0.1% 0.6% - - -

c16:0 34.3% 86.2% 23.6% 32.8% -

c1s:o 12.0% 6.3% 3. 3% 6.6% 38.5%

c,s: 1 31.4% - 8.0% 6.4% 26.9%

c18:2 16.6% - 54.0% 48.0% -

c18:3 - - - 5.6% -

c22:6 - - - - 11.6%

Table II.2. Chain length and chain saturation in natuPal lipid extraets jrom Sigma.

H H 0 '\".+/

0 11 -o,,,, I

'-..+ P· '··.P-O -:;::;N~ / ...:.::;:~~ 0 - -a~ 0 0

'-..+ 0

0 ~3 /N '~ ro 0

2

,

,' 3

aft ~~o !a)

ra ~ ~

~ \

~ --

~ H H

0 '\".+/ 0

11 -o,,,, I ......... + p .,,,, ····P-O ,.....N / "''o- -oV ~o. o

~N o, H 3 / ' rN 2 H :f

OH 0 ,,

~ rN OH ( b) )( ...

~ ~ ~ -

~ Figure II,ll. Compared to phosphatidyLcholine (a), in sphingomyelin (b)

chain following from the formation of a five co-ordinated

intermediate, is blocked.

47

FiguPe II.l2. VePy fast build-up of a five ao-ordinated intermediate in

the phosphatidyZsePine headgPoup, due to twofold stabilization.

Phosphatidylinositol possesses a large ligand in the phospho­

lipid headgroup, the diameter of which can hardly be changed by the transition from the four to the five co-ordinated intermediate. A possibly formed five co-ordinated intermedi­ate will not be stabilized by intramolecular ring formation

so its life-time will not be long enough to induce cluster

formation. Phosphatidylethanolamine already possesses a very

small headgroup in stretched conformation (see II.1.1.), so intramolecular ring-formation will not be able to influence

the angle of tilt of the hydrocarbon chains. Both samples 9

and 10 display a more or less expected rate of ion transport, somewhat slower than that of phosphatidylcholine (sample 1). Addition of these lipids hardly influences the overall lipid fluidity (the ebains display about the same degree of

saturation) and will not be able to induce cluster formation

48

7

(a)

( b)

Figure II.13. (a) Formation a five co-oPdinated intermediate will not

be able to influence hydracaPbon chain tilt in phosphatidylethanolamine.

(b). A five co-ordinated phosphatidylinositol intermediate will not be

stabilized by intramolecular ringformation.

49

via a P(V) TBP intermediate. So the number of phospholipids

that can induce cluster fo·rmation (the phosphatidylcholines} is simply reduced with ± 10%.

11.2.4. Conclusion.

The measurements of ion-fluxes through Gramicidin channels

over the walls of phospholipid vesicles, enabled us to

establish ion transport of different rates, if all the experimental conditions were left constant except the lipid

composition. This can be considered as a support for the

model, indicating lipid mediated influence of membrane

proteins and not just direct influence of proteins by exter­nal factors. To test several aspects of the model, in which

the transition of a four to a five co-ordinated phospholipid

intermediate is the trigger to cluster formation and protein

activation, phospholipids with various headgroups and hydro­car.bon ebains were added to a reference lipid. The results

can be considered as a support of the proposed model.

II.2.5. Bxperimental.

- Preparation of vesicles.

The vesicles are-prepared according to metbod II of Enoch and

Strittmatte.r50, using the same concentrations noted there. SUbsequently, nine· ultrafiltration steps (Millipore, Immer­

sibie CX-30) with a concentration of 10 to 1 ml in every

step, are necessary to reach a fractional integral of inter­

na! resonance relative to internal + outside resonance of

approx. 0.90. This is an indication that Na+ is strongly

bound to the external phospholipidmonolayer and that a high

concentration ratio between the external phospholipid mono­

layer and the external solution is needed for the Na+ ions to desorb. After the last concentrating step to 1 ml buffered

50

vesicle solution 1 ml 020 is added to get the internal

reference for the NMR measurements •

.- Preparatien of the probe.

The probe is synthesized according to the double heterogen­eous reaction59-63:

of 0.746 g Dy203 (dysprosiumoxide, Sigma No. D-0381) and

1.530 g H3NTA (nitrilotriacetate, Sigma No. N-9877) in 25 ml

water with 1.60 ml TEA (triethanolamine, Brocades, s.d. 1.12

1.13) to produce 0.16 Mof shift reagent59. The pH must

never be allowed to rise above 6 or 7 (even transiently)

during early stages, lest oy3+ hydrolysis and precipitation

becomes a problem. So at first 1.0 ml TEA is added toa

stirred suspension of Dy203 and H3NTA at a temperature of 40•c over a time-period of two hours. After heating the

solution becomes clear at 73"C, the pH is 3.5 at that moment.

After cooling to room temper.ature the solution is clowdy

again after standing over one night. Heating to 1o•c and

addition of the last 0.6 rol TEA will give a clear solution

again of pH = 4.5, that stays clear after cooling. Part of

this stock solution is brought to pH = 7 with LiOH just

before addition of the probe to the vesicle suspension, to

keep the pH of that solution as constant as possible and

diluted 1 : 1 with DzO to obtain a final concentratien of

0.065 M of the probe. From this solution 90 ~1 is added to

2 rol of vesicle suspension during the measurements •

.- Gramicidin solution.

2.3 rog Gramicidin no. G-5002 (Sigma) per 100 ml methanol

gives a stock solution of 1.23 x 1o-5 M. 6 ~lof this

solution is added to 2 ml vesicles suspension in 50% 020 I 50% H20. The concentratien of the stock solution is made such

51

that only very little methanol is added to the sample, since

methanol is an anesthetic that perturbs the bilayers strongly64.

~ NMR measurements.

The 23Na NMR {79.4 MHz) spectra are measured on a 300 MHz

Bruker CXP-300 spectrometer. For each measurement 256

free-induction decays (FID) are accumulated in 128 s. For each series an average of 8 blanks at t = 0 (after addition

of the probe) is taken as reference Na+in/Na+total,t=O• After addition of Gramicidin A, the spectra are taken automatically with a computer program, taking successively 30 measurements every 2 minutes, 10 every 4 minutes, 10 every 6

minutes, 10 every 16 minutes and 15 every 31 minutes. If the

ratio Na+in, t=t/Na+in, t=O falls below ± 0.3 the recording is stopped because of too small signal/noise ratio

for Na+in, t=O· Most series are stopped too if t ~ 3 hours. As internal loek 020 is used, as described by Pike et a150.

However, measurements of 2 mM NaCl in H20 or H20/D20, with

acetone or 020 as an external loek, show that the Dys-reagent is a 0-shift reagent as well. Since the shift with 020 as

loek is appreciably greater than without, the choice for 020 is maintained.

52

References and notes.

1. I.I. Merkelbach and H.M. Buck, Reel. Trav. Chim.

Pays-Bas, 1983, 102, 283-284.

2. M. Sundaralingam, Ann. N.Y. Acad. Sci., 1972, 195,

324-355.

3. (a) A. Tardieu, V. Luzzati and F.C. Reman, J. Mol.

Biol. 1973, 75, 711-733; (b) A. Blume, R.J. Wittebort,

s.K. Das Gupta and R.G. Griffin, Biochemistry, 1982, 21,

6243-6253.

4. S.H. Wu and H.M. McConnell, Biochemistry, 1975, 14,

847-854.

5. E.J. Shimshick and H.M. McConnell, Biochemistry, 1973,

12, 2351-2360.

6. F.T. Presti, R.J. Pace and S.I. Chan, Biochemistry,

1982, 21, 3831-3835.

7. T. Lookman, D.A. Pink, E.W. Grundke, M.J. Zückermann and

F. de Verteuil, Biochemistry, 1982, 21, 5593-5601.

8. R.P. Rand, D. Chapman and K. Larsson, Biophys. J., 1975,

15, 1117-1124.

9. M.J. Janiak, O.M. Small and G.G. Shipley, Biochemistry,

1976, 15, 4575-4580.

10. s.c. Chen and J.M. Sturtevant, Biochemistry, 1981, 20,

713-718.

11. P.L. Yeagle, w.c. Hutton, c. Huang and R.B. Martin,

Biochemistry, 1976, 15, 2121-2124.

12. D. Lichtenberg, S. Amselem and I. Tamir, Biochemistry,

1979, 18, 4169-4172.

13. B. Hille, Progr. Biophys. Mol. Biol., 1970, 21, 1-32.

14. B. Hille, Ann. Rev. Physiol., 1978, 38, 139-152.

15. D.J. Vaughan and K.M. Keough, FEBS Lett., 1974, 47,

158-161.

16. S.A. Bone, s. Trippettand P.J. Whittle, J. Chem. Soc.,

Perkin I, 1977, 80-84.

17. H.C. Lüttgau and H.G. Glitsch, Fortschr. Zool., 1976,

24, 1-131, p. 31,33.

53

18. A.M.C.F. Castelijns, Ph. D. Thesis, 'Reaction Mechanisms

in Organophosphorus Chemistry', Eindhoven University of

Technology, 1979.

19. F. Jähnig, K. Harlos, H. Vogel and H. Eibl, Biochemistry, 1979, 18, 1459-1468.

20. J. Seelig, Biochem. Soc. Trans., 1978, 6, 40-42. 21. A. Seelig and J. Seelig, Biochem. Biophys. Acta, 1975,

406, 1-5.

22. G. BÜldt and J. Seelig, Biochemistry, 1980, 19,

6170-6175. 23. H. Träuble, Naturwissensch., 1971, 58, 277-284.

24. A.G. Lee, N.J.N. Birdsall, J.C. Metcalfe, P.A. Toon and

G.B. Warren, Biochemistry, 1974, 13, 3699-3705.

25. G. Inesi, M. Millman and S. Eletr, J. Mol. Biol., 1973, 81, 483-504.

26. J. Seelig and w. Hasselbach, Europ. J. Biochem., 1971, 21, 17-21.

27. P. Woolley and H. Eibl, FEBS Lett., 1977, 74, 14-16. 28. P.R. Cullis, B. De Kruyff. A.E. McGrath, C.G. Morgan and

G.K. Radda, Nobel Symposium, 1977, 34, 389-407. 29. G. Boheim, W. Hanke, S. Oberschär and H. Eibl, Transp.

Biomembr. Model Syst. Reconstr., 1982, 135-143. 30. P.R. Cullis, B. De Kruyff, Biochim. Biophys. Acta, 1979,

559, 399-420.

31. R. Barrison and G. Lunt, Biologica! Membranes. Their

Structure and Function, Blackie, Glasgow, 1980, p. 116, 118.

32. s. Tokutomi, K. Ohki and S.I. Ohnishi, Biochim.

Biophys. Acta, 1980, 596, 192-200.

33. G. Boheim, w. Hanke and B. Eibl, Proc. Natl. Acad. Sci. USA, 1980, 77, 3403-3407.

34. Yu.A. Ovchinnikov, Biomembranes, Sth FEBS meeting, 1972, 279-306~

35. B. Hille, J. Gen. Physiol., 1975, 66, 535-560.

36. W. Veatch, s. Weinstein, B.A. Wallace and E.R. Blout, Pept. Struct. Biol. Funct., Proc. Am. Pept. Symp. 6th, 1979, 635-638.

54

37. B. Hille, Proc. Natl. Acad. Sci., 1971, 68, 280-282.

38. B. Hille, Fed. Proc., Fed. Am. Soc. Exp. Biol., 1975,

34, 1318-1321.

39. N.S. Chandrakumar, V.L. Boyd and J. Hajdu, Biochim.

Biophys. Acta, 1982, 711, 357-360.

40. N.S. Chandrakumar and J. Hajdu, J. Org. Chem., 1982, 47,

2144-2147.

41. I. Pascher and s. Sundell, Chem. Phys. Lipids, 1977, 20,

175-191.

42. D.A. Metzler, 'Biochemistry, The Chemical Reactions of

Living Cells, Acad. Press, New York, 1977, p. 276.

43. Ibid, p. 1015, 1016.

44. R. Harrison and G.G. Lunt, 'Biological Membranes, Their

Structure and Function', Blackie, Glasgow, 1980, p. 268.

45. S. Krasne, G. Eisenman, G. Szabo, Science, 1971, 174,

412-415.

46. L. Stryer, 'Biochemistry', W.H. Freeman and Company, San

Fransisco, 1981, p. 877.

47. Y.A. Ovchinnikov, Eur. J. Biochem. 1979, 94, 321-336.

48. R. Harrison and G.G. Lunt, 'Biological Membranes, Their

Structure and Function', Blackie, Glasgow, 1980, p. 172.

49. M.M. Pike, S.R. Simon, J.A. Balschi and C.S. Springer,

Proc. Natl. Acad. Sci. USA, 1982, 79, 810-814.

50. H.G. Enoch and P. Strittmatter, Proc. Natl. Acad. Sci.

USA, 1979, 76, 145-149.

51. Y. Almog, S. Reich and M. Levy, Br. Polymer. J., 1982,

131-136.

52. L.T. Mimms, G. Zampighi, Y. Nozaki, C. Tanford and J.A.

Reynolds, Biochemistty, 1981, 20, 833-840.

53. P. Läuger, J. Membr. Biol., 1980, 57, 163-178.

54. D.W. Urry, C.M. Venkatachalam, A. Spisni, P. Läuger and

M.D.A. Khaled, Proc. Natl. Acad. Sci. USA, 1980, 77,

2028-2032.

55. A. Finkelstein and O.A. Andersen, J. Membr. Biol., 1981,

59, 155-171.

56. I. Ando and G.A. Webb, 'Th~ory of NMR parameters',

Acadamic Press, 1983.

55

57. J.A. Dani and D.G. Levitt, Biophys. J., 1981, 35,

501-508. 58. Information provided by Sigma.

59. M.M. Pike and c.s. Springer, J. Magn. Reson., 1982, 46,

348-353. 60. C.C. Bryden, C.N. Reilley and J.F. Desreux, Anal. Chem.,

1981, 53, 1418-1425.

61. M.M. Pike, O.M. Yarmush, J.A. Balschi, R.E. Lenkinski and c.s. Springer, Inorg. Chem., 1983, 22, 2388-2392.

62. C.S. Chu, M.M. Pike, E.T. Fossel, T.W. Smith, J.A. Balsebi and c.s. Springer, J. Magn. Reson., 1984, 56,

33-47. 63. J.A. Balschi, V.P. Cirillo, W.J. le Noble, M.M. Pike,

E.C. Schreiber, S.R. Simon and c.s. Springer, Earths Mod. Sci. Technol., 1982, 3, 15-20.

64. L.S. Koehler, E.F. Fossel, K.A. Koehler, Progr. Aenestesiol., 1980, 2, 447-455.

65. E.J.A. Lee, Int. J. Biolog. Macromolecules, 1979, 1, 185-187.

56

Quantum chemical calculations on the stereochemistry of

coenzyme B12-dependent carbon-skeleton rearrangements.

57

58

III. The formation of a carbanionic intermediate in the

carbon-skeleton rearrangement step.

III.l Structure and function of coenayme B12-

Vitamin 812 is an essential nutritional element for the liver

to cure a special form of anaemial, a disease which is

characterized by a disturbed ripening and accelerated

degradation of the erythrocytes. Vitamin 812 itself is cyano­

cobalamin, whose structure was determined by Hodgkin et alf,

using X-ray crystallography (see Figure III.1.). This struc­

ture is composed of two principal parts, the highly substitu­

ted, reduced, porphyrin-like corrin ring and the nucleotide,

which, unlike those obtained from nucleic acids, contains an

a-glycosidic linkage. The corrin ring contains tervalent

cobalt chelated to the four nitrogen atoms of this ring, to a

nitrogen atom of the 5,6-dimethylbenzimidazole ring and to a

cyanide ion, which is an artifact of the isolation procedure.

The entire structure, except for the cyanide, is termed

cobalamin. In addition to cyanide, hydroxide, water and

nitrous acid can be bound to the 6'-position. Vitamin 812

itself is not active as a coenzyme for any known enzymatic

reaction, but there exist two coenzymatically active

derivatives, 5'-deoxyadenosylcobalamin and aquocobalamin.

R" H ,,1 R'-C -C-R"'

1'--1 H H

Figure III.2. \1,2]- Shift of a hydragen atom and a group R.

The last one is involved in three biochemica! processes, the

synthesis of methionine, methane formation and acetate

synthesis. Aquocobalamin falls out of the scope of this

59

Figure III.l. Structure of vitamin

X-ray diffraction studies. as dete~ined by

thesis. Of the known enzymatic reactions which require

5'-deoxyadenosylcobalamin as cofactor, all but one (ribo­

nucleotide reductase} involve an intramolecular [1,2] -shift

of hydrogen coupled with a [1,2]-shift of some other group,

as shown in Figure III.2. The eleven rearrangement reactions

known until now are divided in three groups, the carbon­

skeleton rearrangements, the hydroxyl and the amine

migrations, according to the bond that is broken during the

reaction: a c-c, c-o or C-N bond3,4.

In this study special attention is given to the carbon­

skeleton rearrangements, i.e. the isomerization of L-methyl­

malonyl-Coenzyme A to succinyl-Coenzyme A, threo-B-methylene­

glutarate to L-glutamate and B-methylitaconate to a-methyl­

eneglutarate, where hydrogens for sake of clarity deuterons

with R=

o=( SCoA

methyl­

malonyl-Co A

isomerization

----retention

NH~><' - )=o HO

methyl­

aspartate

isomerization

or

inversion

HO

methyl­

itaconate

isomerization

Figure III.3. The three carbon-skeleton rearrangements dependent on

coenzyme B12

.

61

are used in Figure III.3.) and a carbon-centered group R

migrate in an intramolecular (1,2]-shift. Onder enzymatic

conditions the hydragen (deuteron in Figure III.3.) is trans­ferred via the 5'-methylene group of coenzyme B12S-8 and migrates in methylmalonyl-Co A with retention of configur­

ation9-11 (i.e. the incoming hydragen and the leaving group

R occupy the same position). The migration in methylaspartate

occurs with inversion of configuration12,13, while the

stereochemistry of the methylitaconate isomerization is still

unknown.

111.2. Mechanism of action of the carbon-skeleton rearrange­

ments.

In gene~al for the coenzyme B12-dependent rearrangements a

radiaal mechanism is proposed, as was recently summarized by Rétey14, based upon data arising from isotape labeling15,16

electron paramagnetic resonance17-19 and UV/VIS spectro­

scopie measurements20. As Rétey describes in this

HRe H ~COSCoA

Hs · ''1 ""-, l'cH 3 cooHI _

HS. ( ,H 1 /,, ,_• ----1\:-....r-;;:.-' COS Co A

c~ 'cooH / _

enzyme enzyme

COSCoA H

Hs i~~--~{:::-·' H Re

I CH3 COOH -enzyme enzyme

Figur>e III.4. 3etent·ion of aonfigur>ation in t7w r>adiaal meahanism for> the

methylmalonyl-Co A rearrangement as pr>oposed by ;?étey.

62

article14, this radical mechanism is put forward independent­

ly of the stereochemical results published, while in particu­

lar the steric course of a reaction is extremely useful to

draw conclusions as to the mechanism. In order to integrate

the stereochemical results known and the radical mechanism,

he assumes14,21 a very intimate and unambiguous interaction

between enzyme and substrate, that prevents rotatien around

the C1-C3 bond (see Figure III.4.).

While the Retey model emphasizes the role of the enzyme, it

does not take into account the intrinsic properties of the

substrate, i.e. an enzyme can only lead a substrate through a

reaction-path that is pre-set in (allowed by} the substrate.

Moreover, no explanation is given of the modifications of the

enzyme needed to achieve inversion in the methyl aspartate

isomerization22.

From enzymatic data it has been shown by Pratt23, that the

three groups of rearrangements demonstrate some remarkably

distinct features. The enzymes of the carbon-skeleton

rearrangements require no other cofactors23, while those of

amine migrations all apparently require pyridoxal phosphate

and sametimes other cofactors such as K+, Mg+ and ATP. The

enzymes of the hydroxyl migrations all require simple ions

such as K+. Although the role of some of those factors, e.g.

pyriQOXal phosphate, is uncertain, the question is raised

whether there is a common denominator to the mechanism of

reaction of the different groups of substrates.

Another indication that the three groups of coenzyme 812-de­

pendent reactions differ originates from ESR spectra24. The

enzymes that catalyze the isomerization of diols, glycerol

ethanolamine and the reduction of ribonucleotides give very

unusual and characteristic ESR spectra in the presence of

substrates (i.e. when frezen during catalysis) or substrate

analogues and even, in the case of ehanolamine ammonia lyase,

in the absense of any substrate or substrate analogue25,26.

In all cases, the ESR spectrum consists of two components,

namely a braad resonance at g - 2.3 due to the Co(II) ion

and a narrow doublet centered about g 2 due to an organic

63

radical, the splitting being explained by interaction with

the Co(II) ion. The fact that these paramagnetic species may

account for up to 65% of the coenzyme present depending on the enzyme and the substrate and, where detectable, are formed at a rate comparable to, or greater than, the turnover number of the enzyme strongly suggests that they repcesent

intermediates in the catalytic cycle25. No such signal could, however, be observed with methylmalonyl-Co A mutase27.

Though a radical mechanism might be appropriate for hydroxyl and amine migrations, the probability of such a mechanism

with respect to the carbon-skeleton rearrangements becomes

less. The only artiele publisbed to date indicating a free radical

rearrangement invalving the [1,2]-migration of a thioester

group as a model for the methylmalonyl Co A mutase reaction, is recently publisbed by Halpern et al28. This article,

however, is not very convincing. The kinetic measurements are presentea in such a way that radical and anionic processes

are not directly comparable. As far as one can see from the data presented, the uncatalyzed rearrangement rate of the radical is 235 sec-1 at 60.5"C and 2.5 sec-1 at 30"C, while

kcat for the. enzymatic methylmalonyl-Co A reaction has been estimated to be 102 sec-1. Rearrangement·of the anion gave

after 2 min. more thioester rearranged product (42% versus 1-9% in the radical rearrangement) at a lower temperature (-78"C versus 30"C) which comes closer to the enzymatic

process than the radical rearrangement. Moreover, no reason

is given why Halpern considers the formation of such a

substrate carbanion to be highly unfavorable and much less likely than the alternative free radical rearrangement,

although contributions from pathways invalving carbanion

intermediates cannot be definitively excluded!

Not only there is lot of literature suggesting the three

groups of carbon-skeleton rearrangements have d~fferent

mechanisms of action, there is also accuroulating evidence29-38 that in methylmal9nyl-Co A mutase reactions

the Co-C bond assists in the formation of a substrate carban-

64

ion in the rearrangement step. In the next Chapter we will

try to make plausible, why this could be the mechanism of

action for all the carbon-skeleton rearrangements.

III.3 The stereochemistry of the carbon-skeleton

rearrangements as 'test' for the carbanionic

mechanism. Scope of the second part of this thesis.

The few attempts made to date to include the stereochemical

results known for the carbon-skeleton rearrangements in the

back-side

ethylene glycol

0 H 11 I c---c L CoAS/ èf)::L

front-side

front-side

0 11 H c I

CoAS/ '-c L

\~ / Co

L'\1 "L CH

I COOH

COOH I H

/CH I NH2 'c L

\" / Co L~I"-L

CH

I COOH

Figure III. 5. 'l'he stereoc:hemistry of the

aspartate rearrangement~ according to the mechanism proposed by Corey et al.

65

mechanisme proposed, were incomplete (Rétey14, see Chapter

11.2.) or simply incorrect (Corey et alf9). The keystep Corey

proposes for the inversion of configuration in the conversion

of ethylene glycol to acetaldehyde29 is the migration of a

hydroxylgroup (with electron pair) to an electrophilic adjacent carbon and simultaneous bonding of cabalt (through

its vacant co-ordination side) to the methylene carbon in a back-side displacement of OH (see Figure !!!.5.). This

back-side displacement will lead to inversion of configuration. 1f we suppose that this mechanism is also

applicable to the carbon-skeleton rearrangements, then the -cosco A migration in the methylmalonyl-Co A isomerization to

a nucleophilic adjacent carbon and simultaneous bonding of cabalt to the methylene carbon in a front-side displacement

of the -cosco A group explains the retention of contiguration found for this rearrangement. However, the -CHNH2COOH group in the methyl-aspartate isomerization migrates to the same

nucleophilic carbon as found for the methylmalonyl Co A

rearrangement, which is in contradiction to the inversion of

contiguration known for the methylaspartate isomerization.

As indicated in Chapter 11.2. model studies suggest29-38

that in methylmalonyl-Co A mutase reactions the Co-C bond

assists in the formation of a substrate carbanion in the rearrangement step. 1f this model description (carbanion-formation) could predict the stereochemistry of

the carbon-skeleton rearrangements in general, this could be considered as an extra indication, in addition to the literature known to date29-38, towards a rearrangement of a

anionic intermediate. So the scope of this part of the

thesis is to find a correlation between the evolution of the

coefficients of the atomie orbitals in the Highest Occupied Molecular Orbital (HOMO) of anionic intermediates, and the

stereochemistry known for these reactions.

66

III.4 The nature of the hydrogen transferred temporarily to

coenzyme B12 during the carbon-skeleton rearrangements.

The way in which the anions can be generated is subject to a

lot of speculation. Three possible ways are summarized below.

All three should obey the observation of Miller et al~o,

showing the hydrogen which migrates during the isomerization

of methylmalonyl-Co A to succinyl-Co A becomes one of three

equivalent hydrogens on Cs' of coenzyme B12• before a hydrogen is returned to the substrate. The first two,

suggested by those who hold to initia! radical generation for

all the coenzyme B12-dependent rearrangements consist of radical generation in the substrate, foliowed by either elec­

tron transfer from cobalt to the substrate:

R· ;+- Co(II)·:;;::!':Co(III)+ + R-, or by charge transfer from

protein basic and acidic sites to the substrate radical, a suggestion put forward by Finke41,42. A third possibility

is proton loss from the substrate to Cs' of the coenzyme (see

Figure III.6.), whereby one of the three hydrogens of the

S1.1bstrate &.Ad

+

+

I N

Substrate-

TH.6. Tr>ansfer of a proton :rom the substrate to C/ of coenzyme

of a Co-H-C bPidge.

substrate methylgroup originated becomes covalently bonded to

cobalt via an agostic M(H)C interaction, as proposed by Broekhart et al43,44. They state44 that carbon-hydrogen

honds, especially those of saturated (sp3) carbon centres,

are normally regarded as being chemically inert. Generally,

the C-H group is not thought of as a potential ligand which

can have a structural role or play an energetically

67

significant part in ground statea or in reaction intermedi~ ates. They review recent observations which show that the~

are in fact many circumstances in which a carbon-hydrogen group will interact with a transition metal centre with

formation of a two-electron three-centre bond,and that the \

effect of the interaction is such as to have ~'marked effect

on the molecular and electrooie structure and hence reactivity of the molecule. Once the anionic intermediate is formed, substrate-enzyme interaction can accommodate for the charge build-up in the substrate at the various stages of the rearrangements.

68

References and notes.

1. H.R. Mahler and E.H. Cordes, 'Biological Chemistry',

Harper and Row Publishers, New York, 1971, p. 424.

2. D.C. Hodgkin, J. Kamper, J. Lindsey, M. MacKay, J.

Pickworth, J.H. Robertson, C.B. Shoemaker, J.G. White,

R.J. Prosen and K.N. Trueblood, Proc. Roy. Soc. (London)

A, 1957, 242, 228-263.

3. D. Dolphin, 'B12'• J. Wiley and Sons, New York, 1982,

volurne 1, p. 328.

4. B. Zagalak and W. Friedrich, 'Vitarnin B12', W. de

Gruyter, Berlin, 1979, p. 557.

5. J. Rétey and D. Arigoni, Experientia 1966, 22, 783-784.

6. G.J. Cardinale and R.H. Abeles, Biochirn. Biophys. Acta,

1967, 132, 517-518.

7. R.L. Switzer, B.G. Baltirnare and H.A. Barker, J. Biol.

Chern., 1969, 244, 5263-5268.

8. H.F. Kung and L. Tsai, J. Biol. Chern., 1971, 246,

6436-6443.

9. M. Sprecher, M.J. Clark and D.B. Sprinson, Biochern.

Biophys. Res. Comrn., 1964, 15, 581-587.

10. M. Sprecher, M.J. Clark and D.B. Sprinson, J. Biol.

Chern., 1966, 241, 872-877.

11. J. Rétey and B. Zagalak, Angew. Chern., 1973, 85,

721-722.

12. M. Sprecher and D.B. Sprinson, Ann. N.Y. Acad. Sci.,

1964, 112, 665-660.

13. M. Sprecher,·R.L. Switzer and D.B. Sprinson, J. Biol.

Chern., 1966, 241, 864-867.

14. J. Rétey, Recent Adv. Phytochern., 1979, 13, 1-27.

15. P.A. Frey and R.H. Abeles, J. Biol. Chern., 1966, 241,

2732-2733.

16. J. Rétey, A. Urnani Ronchi, J. Seibl and D. Arigoni,

Experientia, 1966, 22, 502-503.

17. B.M. Babior, T.H. Moss, W.H. Orrna-Johnson and H.

Beinert, J. Biol. Chern., 1974, 249, 4537-4544.

69

18. S.A. Cockle, H.A.O. Hill, R.J.P. Williams, S.P. Davies

and M.A. Foster, J. Am. Chem. Soc., 1972, 94, 275-277.

19. T.H. Finlay, J. Valinsky, A.S. Mildvan and R.H. Abeles,

J. Biol. Chem., 1973, 248, 1285-1290.

20. K.N. Joblin, A.W. Johnson, M.F. Lappert, M.R. Hollaway and H.A. White, FEBS Lett., 1975, 53, 193-198.

21. J. Rétey, Stud. Phys. Theor. Chem., 1982, 18, 367-389.

22. Reference 14, p. 19.

23. D. Dolphin, 'B12'• J. Wiley and Sons, New York, 1982, Volume 1; J.M. Pratt, p. 331.

24. D. Dolphin, 'B12'• J. Wiley· and Sons, New York, 1982, Volume 1; J.M. Pratt, p. 381 •.

25. B. Babior, Acc. Chem. Res., 1975, 8, 376-384.

26. J.F. Boas, P.R. Hicks, J.R. Pillbrow and T.D. Smith, J.

Chem. Soc., Faraday Trans. II, 1978, 417-431. 27. B. Zagalak and w. Friedrich, 'Vitamin B12'• w. de

Gruyter, Berlin, 1979; J. Rétey, p. 440. 28. s. Wollowitz and J. Halpern, J. Am. Chem. Soc., 1984,

106, 8319-8321.

29. P. Dowd and M. Shapiro, J. Am. Chem. Soc., 1976, 98, 3724-3725.

30. B. Zagalak and w. Friedlich, 'Vitamin B12'• W. de

Gruyter, Berlin, 1979; P. Dowd, p. 557. 31. A.I. Scott, J. Kang, D. Dalton and S.K. Chung, J. Am.

Chem. Soc., 1978, 100, 3603-3604.

32. A.I. Scott and K. Kang, J. Am. Chem. Soc., 1977, 99,

1997-1999. 33. G. Bidlingmaier, H. Flohr, O.M. Kempe, T. Krebs and

J. Rétey, Angew. Chem., 1975, 87, 877-8.

34. H. Flohr, W. Pannhorst and J. Rétey, Angew. Chem., 1976,

88, 613-614.

35. H. Flohr, u.w. Kempe, W. Pannhorst and J. Rétey, Angew. Chem., 1976, 88, 443-444.

36. M. Fountoulakis and J. Rétey, Chem. Ber., 1980, 113,

650-668.

37. B. Zagalak, w. Friedlich, 'Vitamin B12'• W. de Gruyter, Berlin, 1979; J. Rétey, p. 439.

70

38. J.H. Grate, J.W. Grate and G.N. Schrauzer, J. Am. Chem.

Soc., 1982, 104, 1588-1594.

39. E.J. Corey, N.J. Cooper and M.L.H. Green, Proc. Natl.

Acad. Sci. USA, 1977, 74, 811-815.

40. w.w. Millerand J.H. Richards, J. Am. Chem. Soc., 1969,

91, 1498-1507.

41. Personal communication of R.G. Finke.

42. R.G. Finke, O.A. Schiraldi and B.J. Mayer, Coord. Chem.

Rev., 1984, 54, 1-22.

43. M. Brookhart, M.L.H. Green and R.B.A. Pardy, J. Chem.

Soc., Chem. Commun., 1983, 691-693.

44. M. Brookhart and M.L.H. Green, J. Organomet, Chem.,

1983, 250, 395-408.

71

72

IV. Quantum chemica! calculations.

IV.l Introduction.

In order to test the model description, suggesting the

formation of a substrate carbanion in the rearrangement step

(see Chapter III.), for the stereochemistry of the

carbon-skeleton rearrangements in general, we selected the

anionic cyclopropane intermediates as giv~n in Figure IV.l.

for the quanturn chemica! calculations.

with R=

o=( SCoA

methyl­

malonyl-Co A isomerization

0

OH H

H R

NH~><' ~ )=o HO

methyl­

aspartate isomerization

methyl­

itaconate isomerization

Figure IV.l. The anionia ayalopropane intermediates which account for the

stereoahemistry of the aarbon-skeleton rearrangements.

For sake of simplicity the calculations were confined to the

substrate system without introducing enzyme or coenzyme

specific sites. The key intermediate as illustrated in Figure

IV.1. is formed by proton abstraction from the methyl group

73

(C3) followed by an approach of group Rand the now negative­ly charged methylene group. The intermediate ring closure

during the isomerization of methylmalonyl-Co A and methyl­

itaconate is facilitated by polarization of the C=O and C=C

bond in group R respectively. In the isomerization of methyl­aspartate a keto-enol tautomerization can provide an

analogous structure, able to accommodate negative charge (see

Figure IV. 2. )

OH

HO

OH

0=

•' ,\H

CH 3

/OH HO

Figure IV.2. The keto-enol tautomerization in methylaspartate resulting

in a structure which accommodates the negative charge after ring closure,

Instead of one enzyme essential for the isomerization of

methylmalonyl-Co A and methylitaconate, the enzyme complex for the isomerization of methylasparate consists of two proteins. The fully optimized MNDO structure of the enol form

is only 2 kcal/mole higher in energy than the keto form, an

energy difference which is smaller than the unpredictable error of the MNDO method1.

The final rearrangement product is formed by proton addition

at the acid substituted c, and rupture of the c,-R bond.

IV.2. The choice of the calculational method.

As indicated by Meiver and Komornicki2, a detailed under­standing of the dynamics and stereochemistry of organic

74

reactions requires, above all, a knowledge of the many

dimensional potential energy surface. The very dimensionality

of this surface, however, precludes its evaluation for all

but the simplest systems. To reduce this problem to one of a

tractable size, two general approaches have been employed.

The first type of approach seeks to reduce the dimensionality

of the surface by eliminating eertaio degrees of freedom, for

example the length of a carbon-hydrogen bond. A related

technique involves choosing one or two degrees of freedom as

independent variables of the potential energy and to allow

the system to relax by optimizing the remaining degrees of

freedom for each value of the independent variables.

The second type of general approach involves consideration of

all the degrees of freedom of a system, but seeks only to

locate eertaio chemically interesting points on the potential

energy surface. For a one-step reaction, these points would

be the local minima corresponding to the equilibrium

geometries of reaetauts and products and a col or saddle

point which separates the local minima.

For a number of reasons this last method, which is preferred

by the author of the cited article, cannot be followed in

this case. The main reasen is, that our approach to the

problem is a dynamic one. Just localizing stationary points

on the curve would not elucidate effects such as fast electron

flow and orbital inversion (vide infra). In the 'product'

minimum these effects will be damped out, while comparison of

the points on the reaction path somewhere in between the

transition state (whose exact position is not important here)

and the 'product' minimum (called 'End' further on in the

article) will give the desired information. An extra handicap

is the definition of a stationary point at the 'Start' (vide

infra) of the ring closure. 'Start' is defined as the linear

molecule with tetrahedral carbon angles just after proton

abstraction. This is not a stationary point, but a rather

unstable intermediate. Reasons why is chosen for the first

type of approach indicated above.

75

IV.3. Results

The formation of the cyclopropane internediatea of Figure

IV.J. bas been studled with MNDO calculationsl.

OH OH OH H

H

OH

Figure IV.3. The three inteTmediates of uhiah the ring alosure has been

studied.

Of course MNDO results cannot give the final proof of the

assumed reaction mechanism, but a detailed anderstanding of the dynamica and stereochemietry of organic reactions requires, above all, a knowledge of the potentlal energy sur­face2. The energy profile of all three ring closure reactions is calculated by optimization of all distances, angles and

torsion angles to minimal heat of formation at a number of fixed values of r1 and r2 (see Figure IV.J.), ranging between 1.35 A and 1.80 A for r1 and between 1.40 A and 2.50 A for

r2. In this way the angle CJ-C1-C2 changes from tetrabedral (in the linear molecule direct after proton abstraction) to triangular (at the end of the ring closure). As initial values for distances and bond angles, those optimized by Dewar et al. for the MNDO program are used. The reaction path is drawn along the line of minimal energy. The results of these calculations are given in the Figures IV.4., IV.S. and IV.6. Following the reaction path, the heat of formation as a

function of the reaction co-ordinate r2 is given for all three ring closure reactions in Figure IV.S. The graphs are closely related, except that the transition state of the

76

2.30 • • I

I

2.20 I • • I I • I • 2.1 0 I • • • • • • • I • I

• I

2.00 • •

1.9 0

1.80

1.70

1.60 ' -130 ' ' •,

' 1.50 ' • •

1.40 1.50 1.60 1.70 1.80 .... r

1 ( Ä)

Fi~~re IV.4. Heat of formation in kcaZ/moZe as a function of r1

and r2

for intermediates during ring cZosure for the methyZmaZonyZ-Co A isome-

rization. 77

1.40 1.50 1.60 1.70 1.80 __ ..... .,_ r

1 (A)

Figure IV.5. Heat of formation in kaat/mote as a fUnation of r1

and r2

for intermediates during ring alosure for the methylaspartate isomerization.

78

r2 { Ä l -145

~· -140 2.40 • •

• • • 2.30

135

• 2.20

• • 2.1 0 •• • I - 13 5 • I • • I

2.00 .~ • I

• ~ • • 1.90 I •• \

\ • , . 1.8 0

\

• \ • 135

-140

1.70 -145

- 1 50

1.60

1.50 • •

1.40 1.50 1.60 1.70 1.80

---~- r 1 ( Ä } IV. 6. Heat kcal/mole as a function of and

for intermediates ring closure for• the methylitaconate isomere-

zat ion. 79

~Hf

(kc a 1/ mole}

• 11 0 *140 0 1 50

• 1 1 5 *145 0155

• 1 2 0 'lf.150 o160

• 1 25 * 1 55 0 1 65.

•130 *160 0 1 70

1

Start End

l

o Methylmalonyl- CoA

• Methylaspartate

*Methyl i toconate

0

\o

' 0-(

2.30 220 2.10 2.00 1.90 1.80 1.70 1.60 r2(A)

Figure IV.?. Heat of formation in kaaZ/moZe as a funetion of the reaetion

ao-ordinate r 2 for the three aarbon-skeZeton rearrangements.

80

methylitaconate ring closure is situated later on the

reaction co-ordinate. A minimum in energy is reached for

values of r2 between 1.60 and 1.50 Ä. The calculations are

not extended beyond this value because in our model the

proton addition to Ct takes place before this minimum in

energy is reached. The charge density accumulated on the

various groups of the intermediate structures varies with the

reaction co-ordinate r2 in a quite similar way for the three

rearrangements, with the exception of the charge density on

the group which stahilizes the negative charge by polariz­

ation of a double bond, i.e. C=O, C=C(OH)2 and C=CH2 (C=X in

Figure IV.8.). In the isomerizations of methylaspartate and

methylitaconate, the charge accommodated by this group in the

beginning of the reaction co-ordinate is high in comparison

with the isomerization of methylmalonyl-Co A, as can be seen

in Figure IV.8.

0 Methylmalonyl-Co A . Methylaspartate e

1

0 . Methylitaconate ' * . ' . ' *-.. .,

*-.. . ' 0.2 *-..

* .,

' ., * . '* ' '*

., '*

., . -0.4 O-o

'* ..... -o-o '

., '0,0 *

. "" -o_ ' o, *

o_o '* -o_ '*--*-*'* o-o

-0.6 -o-o-o

2.2 2.0 1.8 1.6 -r2 !.&.l

Figure IV.B. Charge on group =X as a function of r2

for the carbon-skele­

ton rear1•angements.

81

-0.25 -0.16 -0,21

-aot,~~H H I H

-M>( COOH H

A. ~ r3 1.43 3 -0.63 1.46 3 -0,57 1.53 .... ~

CQ H / H H :;:::

/ \1) ":i " /

~ \1)

" r2 / ~ f..; ..

/ /2.15 ti- :0::: .. 2.35 (0 ;c

-M>( Q

/ ti- Start TS _,.,( End ~

CQ

l _",( 0 -\l) (0

-0.39 SH 0 SH ~ SH -0-44

\l) -0.18 -0.58 -0.28 - 0.17

":i' ~ .... Jl ......

C>

Q 2 -0,29 ...... ...... -G.20 -0.17 C> ....

II ( COOH H

-M<( COOH H

_", ( COOH H (0 ~ I :;::: ;::)

~ <i- - o.oa H 1.47 3 -0.52 1.52 .... 3 o.oo C> +0.23 H / H ;:-! H H H C> / <::: ..

/ i \1) " r2 I ":i

" 2.35 /2.12 (65 1.50 ti- / ;:.• Start / TS End \l)

.!! (Ho-c - -M<( - -· .. ( ~ -o.23 1 NH2 HO-C NH2 Ho--C NH2 .g OH I -o.ov I -0.09 ":i -0.23 -0.05 OH .g -0.47 OH

~ - 0,79

-0.30 -0,22 .... III -0.21

-· .. ( ;:-!

( COOH H -MO(

COOH H COOH H <i-\1) -o.o1 H ~ 1.48 3 -0,49 1,&2 3 +0.01 +0.24 \1) H / H H H A. / .... ~ "' / <i- ",. "' r2 / \1)

, "'2.35 / 2.00 1,73 1.&0 Co

~ ... ( " Start / End - / TS -CH2 2 /0,04 ( ""' -0.12 -~( CH2 COOH COOH -o.:sa -0.14

-0.51 CH2 COOH -0.11 -0.54 -0.16

The charge distribution over the intermediate structures at

different stages of the ring closure is given in Figure IV.9.

The definition of the intermediates Start, TS (transition

state) and End is given in Figure IV.7.

The evolution of the coefficients of the atomie orbitals in

the HOMO of the cyclopropane intermediates along the reaction

co-ordinate is followed. The two atomie orbitals with the

highest coefficient in the HOMO are given in Figure IV.lO. as

a function of r1 and r2.

In all three reactions the overall picture is the same. In

the beginning of the ring closure the contribution of the

atomie orbitals on C3, the carbon from which the proton is

abstracted, is dominant. After the transition state the

atomie orbital on the formerly double bonded oxygen

I ',~ I

x

2.1

1.9 ~ x I

\

1.7

1.5

\

\~ .... 2

' x

1.4 1.6 1.8

Figure IV.lO. The two atomie orbitaZs with the

HOUO for all three carbon-skeleton rearrangements.

83

respectively carbon, in the direction of C3 becomes important. Near the end of the ring formation the orbital on

c, in the direction of c2 will have a large coefficient in the HOMO, as can beseen in Figure IV.10. A very important difference between the three reactions becomes clear, if also atomie orbitals with smaller coefficients in the HOMO are taken into account. If the intermediate is situated in the y-z plane, with the c,-c2 bond on the y-axis, the contribution of the atomie orbitals of the three carbons constituting the ring is given in Table IV.1. for rt=1.7 A and r2=l. 6 A.

methyl- methyl- methyl-malonyl-Co A aspartate itaconate

c, s - 0.15 - 0.12 - 0.14 x + 0.08 + 0.03 + 0.04 y + 0.50 + 0.31 + 0.38 z - 0.30 - 0.20 - 0.25

c2 s + 0.03 + 0.04 + 0.04 x + 0.01 + 0.03 + 0.03 y - 0.09 + 0.07 + o. 11

z - 0.21 + 0.03 + 0.03

c3 s + 0.11 + 0.06 + 0.06 x + 0.00 + o.oo + o.oo y - 0.31 - 0.18 - 0.22 z + 0.22 + 0.08 + 0.09

TabZe IV.l. The aoeffieients of the atomie orbitaZs of the ring aarbons

at t/he moment of proton addition to the ayaZopropane intermediates.

84

The coefficients of the atomie orbitals of Ct and c2 are of opposite sign in the y-direction and of equal sign in the

z-direction in the methylmalonyl-Co A intermediate. Both

indicate an overlap between the atomie orbitals on Ct and c2, as can beseen in Figure IV.11., i.e. the HOMO has a bonding

character between Ct and c2, the electron density between Ct and C2 is high. In the methylaspartate and methylitaconate

intermediates, the contributions of the atomie orbitals of Ct and C2 to the HOMO are of equal sign in the y-direction and of opposite sign in the z-direction, i.e. the HOMO has an

anti-bonding character between Ct and c2. There is a node in

electron density between Ct and c2. The two orbitals just below the HOMO in energy do not play an important role in the

picture between Ct and c2.

methyl­

malonyl-Co A

methyl­

aspartate

methyl­

itaconate

IV.ll. mn'uJ.?,ruJ and anti-bonding character of the c1-c2 bond in

h0,10.

IV.4. Discussion.

Further examination of the pattern given by the way charge

delocalizes in the model anionic intermediate {intermediate Start in Figure IV.9.), shows that the negative charge is

best stabilized on the formerly double bonded oxygen in

85

methylmalonyl-Co A, in comparison with the way the -C(OB)2

group accommodates the negative charge in methylaspartate.

The methylene group of methylitaconate behaves intermediate in accommodation of the negative charge. The relative small

initia! accommodation of negative charge by the diol group in methylaspartate, can be considered as a prerequisite for a

larger charge migration during ring closure. Such a large

charge migration from C3 over C2 to C1 might force the elec­tron density to pass momentarily beyond c,. Subsequent

orbital inversion at C2 will prevent the electron density to

flow back via the c,-c2 bond and to delocalize over the cyclopropane ring. Proton addition at that moment on the reaction co-ordinate will lead to inversion of configuration on c,, i.e. the proton comes in at the opposite side of the

leaving glycyl group. The relative small charge migration during ring closure of methylmalonyl-Co A will not be strong enough to force the electron density to pass c,, resulting in

retention of contiguration at c,. This rather new concept of

orbital inversion to prevent electron back-donation to c2 is

strongly supported by the picture originating from the

development of the coefficients of the atomie orbitals in the HOMO. Figure IV.10. shows charge migration via C2 and not

directly from c3 to c,. This direction of migration is necessary for the electron density to pass c, in line of the

c,-c2 bond, which will lead to inversion of configuration on c,. Figure IV.11. indicates an anti-bonding character for the

c,-c2 bond of the methylaspartate intermediate, which

prevents the electron density to flow back to C2 and the proton to add to the c,-c2 bond. The bonding character of the

c,-c2 bond in the methylmalonyl-Co A intermediate will cause proton addition witn retention of configuration at c,, due to

the high electron density between c, and C2. The anti-bonding character of the c1-c2 bond in the HOMO of the methyl­itaconate intermediate suggests inversion of configuration in

this rearrangement. Finally it may be of interest to note that in the case of methylmalonyl-Co A mutase using ethyl­

malonyl-Co A as substrate instead of methylmalonyl-Co A only

86

partial inversion is observed4. Besides the role of the

enzyme in the enzyme-substrate binding (ethylmalonyl reacts

at only one thousandth the rate of the natural substrate),

the loss of stereospecificity in the case of ethylmalonyl­

Co A may be also electronic in nature.

IV.5. Conclusion.

Energy profiles of the carbon-skeleton rearrangements with

cationic or radical intermediates in the rearrangement step

are required to discuss a possible preferenee for the anionic

pathway. However, the fact that an anionic intermediate can

explain the known stereochemistry, can be considered as an

extra indicator in addition to the chemica! evidence sugges­

ting a carbanion in the rearrangement step of the methyl­

malonyl-Co A isomerization.

87

References eand notes

1. J.S. Dewar and W. Thiel, J. Am. Chem. Soc., 1977, 99,

4899-4907.

2. J.W. Meiver and A. Komornicki, J. Am. Chem. Soc., 1972, 94, 2625-2633.

3. J.S. De war and w. Thiel, J. Am. Chem. Soc., 1977, 99,

4907-4917.

4. J. Rétey and B. Zagalak, Angew. Chem., 1973, 85, 721-2.

5. l.I. M.erkelbach, H.G.M.. Becht and H.M.. Buck, accepted for

publication in J. Am. Chem. Soc., 1985, 107.

88

Summary.

The first part of this thesis describes investigations con­

cerning the role of penta co-ordinated phospholipid inter­

mediates ~n the activatien of membrane embedded proteins,

viz. ion channels. On the basis of extensive literature

search a model is drawn in which a conformational change in

the phospholipid headgroup from a four co-ordinated tetra­

gonal (P(IV)) toa five co-ordinated trigonal bipyramidal

(P(V) TBP) structure results in a change of conformation of

the lipid hydracarbon region. The conformational change in

the phospholipid headgroup is induced via external factors as

cation concentration, potentlal field and binding of e.g. a

water molecule. Enhanced electronegativity on the axial phos­

phate oxygen of the five co-ordinated intermediate, leads to

repulsion of the sn-2 oxygen bound via an 0-C-C-0 sequence to

the phosphate group. The resulting adaptation in the hydra­

carbon chain region will lead to cluster formation in the

sense of a region of different density compared to the

surrounding matrix. Cluster formation around a membrane

protein (viz. ion channel) will activate that ion channel to

open. To test several aspects of this model description,

phospholipid vesicles are synthetized of varying lipid

composition. Over their wall an ion gradient is created,

whose equilibration is followed with 23Na NMR after addition

of Gramicidin A. Leaving all the experimental conditlans as

constant as possible, except the lipid composition of the

vesicle wall, passive ion transport was measured of different

rates. The relative rate of ion transport as predicted on the

basis of the described model, was established by the measure­

ments.

89

The second part of this thesis uses the stereochemistry of

the carbon-skeleton rearrangements dependent on coenzyme B12 as a 'test' for the mechanism of action postulated for these

isomerizations. An anionic mechanism is given, based on

extensive literature search, suggesting an anionic mechanism for the methylmalonyl-Co A rearrangement. With the help of

MNDO calculations, energy profiles are constructed for the

three ring closure reactions towards anionic enzyme stabil­ized cyclopropane intermediates. Following the reaction path,

charge distribution and migration in the substrates are moni­tored, as well as the evolution of the coefficients of the

atomie orbitals in the highest occupied molecular orbital (HOMO) of the cyclopropane intermediates. Large charge migra­

tion as a consequence of small charge stabilization by a

certain group in the anionic intermediate is found to be a prerequisite for the inversion of configuration known for the methylaspartate isomerization. Conversely small charge migration as a consequence of good

charge stabilization in the methylmalonyl-Co A rearrangement

will lead to retention of configuration. For the methylita­conate isomerization inversion of configuration is suggested.

90

Samenvatting.

Het eerste gedeelte van dit proefschrift beschrijft het

onderzoek naar de rol van vijf-gecoördineerde fosfolipide

intermediairen bij de activering van in het membraan opgeno­

men eiwitten, zoals ion-kanalen. Op basis van een uitgebreid

literatuur onderzoek is een model opgesteld, waarin een con­

formatie verandering in de fosfolipide hoofdgroep van een

vier-gecoördineerde tetragonale (P(IV)) naar een vijf-gecoör­

dineerde trigonale bipyramidale (P(V) TBP) structuur, resul­

teert in een conformatie verandering in de koolwaterstof

ketens. De conformatie verandering in de fosfolipide hoofd­

groep wordt geïnduceerd door externe factoren als kation-con­

centratie, potentiaal veld en de binding van bijvoorbeeld een

water molecule. Een verhoogde electronegativiteit op de axi­

ale fosfaat zuurstof van het vijf-gecoördineerde intermedi­

air, leidt tot afstoting van de sn-2 zuurstof, die via een

o-e-c-o keten gebonden is aan de fosfaat groep. De resulte­

rende aanpassing in de koolwaterstof ketens induceert cluster

vorming. De term 'cluster vorming' wordt hier gebruikt voor

een gebied met een dichtheid, verschillend van die van de

omgeving. Cluster vorming rond een membraan eiwit (het ion­

kanaal) zal het ion-kanaal activeren, waardoor het ion-trans­

port start. Om verschillende aspecten van het model te tes­

ten, zijn fosfolipide 'vesicles' gesynthetiseerd met varie­

erende lipide samenstelling. Over de wand van deze vesicles

is een ion-gradient gecreëerd. Het verdwijnen van deze

gradient na toevoeging van Gramicidine A is gevolgd met

behulp van 23Na NMR. Tijdens deze metingen zijn alle experi­

mentele omstandigheden, behalve de lipide samenstelling van

de vesicles, zo constant mogelijk gehouden. Zo kan ion-trans­

port gemeten worden van verschillende snelheid. De met behulp

van het beschreven model voorspelbare relatieve snelheid van

ion-transport is bevestigd door de metingen.

91

Het tweede gedeelte van dit proefschrift gebruikt de stereo­chemie van de coenzyme 8·12 afhankelijke koolstofskelet isome­

risaties als een 'test' voor het werkingamechanisme dat gepostuleerd is voo• deze omleggingen. Een anionisch mecha­nisme wordt voorgesteld, gebaseerd op een groeiende hoeveel­heid literatuur, die een anionisch mechanisme voor de methyl­malonyl-Co A omlegging suggereert. Met behulp van MNDO bere­keningen, zijn energieprofielen opgesteld voor de drie ring­sluitingareacties naar anionische enzyme-gestabiliseerde cyclopropaan intermediairen. Ladingsverdeling en migratie zijn gevolgd gedurende de reactie, alsook de verandering van

de coëfficiënten van de atomaire orbitalen in de hoogst bezette moleculaire orbital (= HOMO) van de cyclopropaan

intermediairen. Sterke ladingsmigratie, als gevolg van een geringe ladinga­stabilisatie door een bepaalde groep in het anionisch inter­mediair, blijkt een voorwaarde te zijn voor de inversie van configuratie, gevonden voor de methylaspartaat isomerisatie. In tegenstelling hiermee leidt een geringe ladingsmigratie, als gevolg van een goede ladingsstabilisatie in de methyl­malonyl-Co A omlegging, tot retentie van configuratie. Voor de methylitaconaat isomerisatie wordt op grond van de bereke­ningen een inversie van configuratie verondersteld.

92

Curriculum vitae.

De schrijfster van dit proefschrift werd op 3 augustus 1954

geboren te Mozambique. Na het behalen van het diploma HBS-B

aan het Lorentz Lyceum te Eindhoven in 1971, werd hetzelfde

jaar begonnen met de opleiding tot Chemisch Analist A aan het

Instituut voor Hoger Beroeps Onderwijs te Eindhoven. In 1973

werd dit diploma behaald, waarna begonnen werd met de studie

op de afdeling der Scheikundige Technologie aan de Technische

Hogeschool te Eindhoven. Het afstudeerwerk werd verricht bij

de vakgroep Organische Chemie, onder leiding van Prof. Dr.

H.M. Buck en Dr. Ir. D. van Aken. In novémber 1980 werd het

ingenieursexamen met lof afgelegd.

Van 1 december 1980 tot 1 mei 1985 was zij werkzaam bij de

vakgroep Organische Chemie van de Technische Hogeschool

Eindhoven als wetenschappelijk assistente. Tijdens deze

periode werd het onderzoek, beschreven in dit proefschrift,

uitgevoerd onder leiding van Prof. Dr. H.M. Buck.

93

Dankwoord.

Gedurende het onderzoek, dat geleid heeft tot dit proef­schrift, heb ik van velen steun ondervonden. Alhoewel ik niet allen bij naam kan noemen, wil ik toch enkele van hen zeer

specifiek bedanken. Als eerste wil ik bedanken het technisch en administratief personeel, zowel binnen als buiten de vakgroep. Zij hebben

mij geholpen de in tijden van bezuiniging verouderde appa­

ratuur op te knappen. Ten tweede wil ik bedanken degenen die mij moreel ondersteund hebben, waarbij ik mijn echtgenoot, familie, vrienden en de koffie-club uit de rookvrije kantine graag expliciet wil

noemen. Bij het tot stand komen van de 23Na NMR heb ik steun onder­vonden van Jan de Haan en Leo van de Ven.

Bij het rekenwerk heb ik veel steun ondervonden van Maarten Donkersloot. Hanneke Becht wil ik bedanken voor het werk dat zij tijdens haar afstuderen op het gebied van Vitamine B12 heeft verricht.

Tenslotte wil ik Ria Hoozemans hartelijk bedanken voor het typewerk, dat leidde tot de uiteindelijke vormgeving van dit proefschrift. De mogelijkheid die zij creëerde, dit proef­schrift met behulp van de tekstverwerker tot stand te laten komen, gaf mij de gelegenheid deze laatste fase van de promo­

tie overeenkomstig de eisen van deze tijd te volbrengen. De bijzonder fraaie tekeningen in dit proefschrift zijn gemaakt door Henk Eding en Cas Bijdevier.

Financial support for this investigation was provided by

Unilever Research, Vlaardingen, The Netherlands.

94

I. De bewering van Roberts en medewerkers dat de 3p én de 3p orbital x y

deel uitmaken van de enkel bezette MO in TBP fosforanylradicalen, moet

op grond van symmetrie overwegingen verworpen worden.

J.W. Cooper, M.J. Parrotten B.P. Roberts,

J. Chem. Soc., Perkin II, 1977, 730-741.

2. Modelverbindingen voor koolstofskelet isomerisaties onder invloed van

Vitamine B12

, waarin het substraat gebonden is aan het centrale cobalt-

atoom, kunnen niet worden gebruikt om conclusies te trekken met betrek-

king tot waterstof abstractie van niet geactiveerde koolstofatomen.

P. Dowd en B.K. Trivedi, J. Org. Chem. 1985, 50, 206-217.

3. Het is de vraag of de ontwikkeling van nat naar droog ontwikkelbare

lakken in de lithografie ten behoeve van de I.C. technologie, analoog

aan de ontwikkeling van nat naar plasma etsen, zal leiden tot een be-

tere patroondefinitie.

4. Het inschatten van de mogelijkheid korter te gaan werken en een deel

van de eigen functie over te dragen aan anderen, is leiding geven-

de functionarissen mede afhankelijk van het vermogen de eigen complexe

functie te analyseren en van het vermogen tot delegeren,

Mr. H. Luik, Intermediair 35, augustus 1984, 1-5.

5. De conclusie van Bock en medewerkers, dat bij de isomerisatie van D-

fructose naar D-glucose onder invloed van het enzym glucose-isomerase,

het a-anomeer het begin van de omzetting met een hogere prefe-

rentie gevormd wordt dan uit de evenwichtsteestand blijkt, is een

overschatting van de mogelijkheden van de NMR techniek.

K. Bock, M. Meldal, B. Meyer en L. Wiebe,

Acta Chem. Scand. B, 1983, 37, 101-108.

6. De constatering dat sommige proteinen goed functioneren in een milieu

bestaande uit enkel detergentia, behoeft niet in strijd te zijn met

de bewering dat membraanproteinen een specifieke lipide-omgeving nodig

hebben om geactiveerd te worden.

P.R. Cullis en B. de Kruyff, Biochim. Biophys •. Acta, 1979,

559, 399-420.

7. De 'closed-shell' 1A1 overgangstoestand met c2v symmetrie van de ther­

mische [1,5] - H shift in pentadiëen, zoals berekend doorHessen Schaad,

is een aangeslagen toestand en niet een grondtoestand.

B.A. Hess en L.J. Schaad, J. Am Chem. Soc. 1983, 105, 7185-7186.

8. Het wekt op zijn minst enige verwondering, dat het vrijkomen van for­

maldehydgas uit spaanplaat bekend is en tot klachten leidt bij een aan­

zienlijk gedeelte van de Nederlandse bevolking, terwijl de aanwezigheid

van ditzelfde gas in sigaretterook systematisch genegeerd wordt.

9. Bij de productie van (Al, Ga) As - halfgeleiderlasers, geschikt voor

toepassing in 'compact-disk' spelers, met behulp van organometaalgas­

fase epitaxie, verdient het aanbeveling meer aandacht te besteden aan

de epitaxiale groei van (Al, Ga) As - lagen waarbij de groeisnelheid

kinetisch bepaald is.

International Conference on HOVPE II, april 1984,

J. Cryst. Growth, 1984, 68.

JO. Bij de bestudering van conformatieveranderingen in DNA structuren,

die optreden ten gevolge van complexering met metaalionen, wordt een

onjuist beeld van de fosfaat-metaal complexatie verkregen, indien ge­

bruik wordt gemaakt van 5'-nucleotiden die enkelvoudig zijn veresterd

op de fosfaatgroep.

S.K. Hiller, D.G. VanDerveer en L.G. Marzilli, J. Am. Chem.

Soc., 1985, 107, 1048-1055.

ll. Bedrijven en overheidsinstellingen zullen via de wet gedwongen moeten

worden om een bepaald percentage vrouwen in dienst te nemen.

Mr. F.H.A.M. Kruse, directeur-generaal Arbeidsvoorzieningen,

i'finis ter ie van Sociale Zaken, januari 1985, op het congres

'Wilma wil werk'.

Ir. I. I. l,!erkelbach,

10 mei 1985.


Recommended