7
Exp. Eye Res. (1996) 62, 47–53 Lipid–Protein Interactions in Human and Bovine Lens Membranes by Fourier Transform Raman and Infrared Spectroscopies HIDETOSHI SATO a,b , DOUGLAS BORCHMAN a *, YUKIHIRO OZAKI b , OM P. LAMBA a , W.CRAIG BYRDWELL c , M. C. YAPPERT c C.A.PATERSON a a Department of Ophthalmology and Visual Sciences, and c Department of Chemistry, University of Louisville, Louisville, KY 40292, U.S.A. and b Department of Chemistry, Kwansei Gakuin University 1-1-155, Uegahara, Nishinomiya 662, Japan (Received Rochester 15 June 1995 and accepted in revised form 23 August 1995) In other systems, proteins have been shown to alter the molecular structures of lipids in the cell membrane bilayer. We wished to determine if proteins altered the structure of lens lipids. The structure of lipid hydrocarbon chains in urea purified human lens membrane vesicles containing intrinsic, hydrophobically bound proteins was compared to the structure of lipids in vesicles without protein. Fourier transform Raman spectroscopy was used to characterize lipid and protein structure. To study lipid interactions with extrinsic, surface bound proteins, the lipid structure was compared in bovine lipid vesicles with and without α-crystallin bound to the surface of the membrane. Lipid structure was studied using Fourier transform infrared spectroscopy. No change in lipid structure was detected even at protein}lipid weight ratios of two to one. Human lens intrinsic proteins contained a high amount of a helical structure (60 %), but did not alter hydrocarbon chain interactions. # 1996 Academic Press Limited Key words : lens ; spectroscopy ; protein ; lipid structure. 1. Introduction Mateu et al. (1978) have shown that proteins alter lipid structure. Extrinsic proteins, proteins that bind to the surface of membranes, were found to disorder the hydrocarbon chains of lipids. Intrinsic proteins, proteins that span the bilayer or are deeply imbedded in the bilayer, ordered the hydrophobic layer of lipid membranes. Changes in lipid–protein interactions in cataractous lenses (Takemoto and Rintoul, 1983) may affect the structure of protein bound lipids and thus could impact upon lens membrane passive perme- ability properties and}or cation pump activity. About 50 % of the weight of lens membranes is attributable to intrinsic proteins (Takemoto and Rintoul, 1983 ; Roy et al., 1982 ; Broekhuyse and Kuhlmann, 1978 ; Broekhuyse and Kuhlmann, 1979). These proteins could contact as much as 40 % of the lipid hydrocarbon chains and could affect the structure of the membrane. α-Crystallin has been shown to bind to lens membranes (lfeani and Takemoto, 1989 ; Mulders et al., 1985, 1989 ; Liang and Li, 1992 ; Ifeani and Takemoto, 1990, 1991a, 1991b ; Zhang and Augusteyn, 1994), and to immobilize the lipids (Liang and Li, 1992 ; Puskin and Wiese, 1982), evidence that structural alterations could exist. We have determined that hydrocarbon chain structural order increases in human lenses with age (Borchman et al., 1994b) and cataract (Borchman et al., 1993) in the cortex and * For correspondence at : Department of Ophthalmology and Visual Sciences, 301 E Muhammad Ali Blvd., Louisville, KY 40292, U.S.A. nucleus. Dynamic order has also been shown to increase (Takemoto and Rintoul, 1983 ; Puskin and Wiese, 1982 ; Liang et al., 1989). The increase in lipid order with age has been correlated with an increase in sphingomyelin and decrease in phosphatidylcholine (Borchman et al., 1994a, 1994b). Our previous lipid structural studies were conducted without protein being present. In this study we examined protein–lipid inter- actions by extrinsic and intrinsic proteins in bovine and human lens membranes, respectively. 2. Materials and Methods All aspects of this work adhered to the recom- mendations from the Declaration of Helsinki, Finland. Extrinsic Protein–Lipid Interactions Using Bovine Lipids and Proteins Bovine lens lipid extraction. Bovine eyes were obtained fresh from a slaughterhouse. Lenses were removed and cortical and nuclear tissues were separated and lipid was extracted from the cortical tissue. A monophasic methanolic extraction (Borchman et al., 1994) followed by a hexane– isopropanol purification was used to extract lipid from 136 g of pooled cortical material. After the methanol extraction (Borchman et al., 1994), the methanol was evaporated in an atmosphere of argon in a rotary evaporator, RE121 (Buchi, Switzerland) at 55C. The lipid was redissolved in hexane–isopropanol (1 : 1, v}v) and centrifuged. The purpose of this purification was 0014-4835}96}01004707 $12.00}0 # 1996 Academic Press Limited

Lipid –Protein Interactions in Human and Bovine Lens Membranes by Fourier Transform Raman and Infrared Spectroscopies

Embed Size (px)

Citation preview

Exp. Eye Res. (1996) 62, 47–53

Lipid–Protein Interactions in Human and Bovine Lens Membranes

by Fourier Transform Raman and Infrared Spectroscopies

HIDETOSHI SATOa,b, DOUGLAS BORCHMANa*, YUKIHIRO OZAKIb, OM P. LAMBAa,

W. CRAIG BYRDWELLc, M. C. YAPPERTc C. A. PATERSONa

aDepartment of Ophthalmology and Visual Sciences, and cDepartment of Chemistry, University of

Louisville, Louisville, KY 40292, U.S.A. and bDepartment of Chemistry,

Kwansei Gakuin University 1-1-155, Uegahara, Nishinomiya 662, Japan

(Received Rochester 15 June 1995 and accepted in revised form 23 August 1995)

In other systems, proteins have been shown to alter the molecular structures of lipids in the cellmembrane bilayer. We wished to determine if proteins altered the structure of lens lipids. The structureof lipid hydrocarbon chains in urea purified human lens membrane vesicles containing intrinsic,hydrophobically bound proteins was compared to the structure of lipids in vesicles without protein.Fourier transform Raman spectroscopy was used to characterize lipid and protein structure. To study lipidinteractions with extrinsic, surface bound proteins, the lipid structure was compared in bovine lipidvesicles with and without α-crystallin bound to the surface of the membrane. Lipid structure was studiedusing Fourier transform infrared spectroscopy. No change in lipid structure was detected even atprotein}lipid weight ratios of two to one. Human lens intrinsic proteins contained a high amount ofa helical structure (60%), but did not alter hydrocarbon chain interactions.

# 1996 Academic Press LimitedKey words : lens ; spectroscopy; protein ; lipid structure.

1. Introduction

Mateu et al. (1978) have shown that proteins alter

lipid structure. Extrinsic proteins, proteins that bind

to the surface of membranes, were found to disorder

the hydrocarbon chains of lipids. Intrinsic proteins,

proteins that span the bilayer or are deeply imbedded

in the bilayer, ordered the hydrophobic layer of lipid

membranes. Changes in lipid–protein interactions in

cataractous lenses (Takemoto and Rintoul, 1983) may

affect the structure of protein bound lipids and thus

could impact upon lens membrane passive perme-

ability properties and}or cation pump activity. About

50% of the weight of lens membranes is attributable to

intrinsic proteins (Takemoto and Rintoul, 1983; Roy

et al., 1982; Broekhuyse and Kuhlmann, 1978;

Broekhuyse and Kuhlmann, 1979). These proteins

could contact as much as 40% of the lipid hydrocarbon

chains and could affect the structure of the membrane.

α-Crystallin has been shown to bind to lens membranes

(lfeani and Takemoto, 1989; Mulders et al., 1985,

1989; Liang and Li, 1992; Ifeani and Takemoto,

1990, 1991a, 1991b; Zhang and Augusteyn, 1994),

and to immobilize the lipids (Liang and Li, 1992;

Puskin and Wiese, 1982), evidence that structural

alterations could exist. We have determined that

hydrocarbon chain structural order increases in

human lenses with age (Borchman et al., 1994b) and

cataract (Borchman et al., 1993) in the cortex and

* For correspondence at : Department of Ophthalmology andVisual Sciences, 301 E Muhammad Ali Blvd., Louisville, KY 40292,U.S.A.

nucleus. Dynamic order has also been shown to

increase (Takemoto and Rintoul, 1983; Puskin and

Wiese, 1982; Liang et al., 1989). The increase in lipid

order with age has been correlated with an increase in

sphingomyelin and decrease in phosphatidylcholine

(Borchman et al., 1994a, 1994b). Our previous lipid

structural studies were conducted without protein

being present.

In this study we examined protein–lipid inter-

actions by extrinsic and intrinsic proteins in bovine

and human lens membranes, respectively.

2. Materials and Methods

All aspects of this work adhered to the recom-

mendations from the Declaration of Helsinki, Finland.

Extrinsic Protein–Lipid Interactions Using Bovine Lipids

and Proteins

Bovine lens lipid extraction. Bovine eyes were

obtained fresh from a slaughterhouse. Lenses were

removed and cortical and nuclear tissues were

separated and lipid was extracted from the cortical

tissue. A monophasic methanolic extraction

(Borchman et al., 1994) followed by a hexane–

isopropanol purification was used to extract lipid from

136 g of pooled cortical material. After the methanol

extraction (Borchman et al., 1994), the methanol was

evaporated in an atmosphere of argon in a rotary

evaporator, RE121 (Buchi, Switzerland) at 55°C. The

lipid was redissolved in hexane–isopropanol (1:1, v}v)

and centrifuged. The purpose of this purification was

0014-4835}96}010047­07 $12.00}0 # 1996 Academic Press Limited

48 H. SATO ET AL

to eliminate a non-protein impurity (or impurities)

(which is not soluble in hexane–isopropanol) that

absorbs at 270 nm. The elimination of the impurity

from our extracted lipid was confirmed by the absence

of any absorbing band at 270 nm. We found that this

monophasic extraction}purification procedure was

the simplest, quickest and most efficient method for

the extraction of lens lipids and is much better than

the standard Folch procedure (Folch et al., 1957). To

be certain that all of the lipid was extracted from the

lens tissue, the pellet from the first methanol extraction

step, which may contain trace amounts of lipid, was

extracted by the standard Folch (1957) procedure and

then by diethyl ether extraction. The lipids from all of

the extractions were pooled and stored in an at-

mosphere of nitrogen at ®70°C. The amounts

obtained after the Folch (1957) extraction wash and

diethyl ether extractions were negligible which indi-

cates that all the lipid was extracted in the methanol

step. The yield of lipid was 0±4 g per 136 g of lens

tissue (wet weight), measured gravimetrically. The

$"P-nuclear magnetic resonance (NMR) spectrum, (see

Borchman et al., 1994, for methods) Fig. 1, shows

qualitatively no phospholipid compositional abnor-

malities. Because a fast pulsing sequence was used to

acquire the $"-NMR spectrum in Fig. 1, only a

qualitative estimation of phospholipid composition

can be made as the relaxation of the phospholipids was

not complete.

α-Crystallin–lipid binding protocol. Bovine lipid was

dispersed by sonication in Hepes buffer (pH 7±4, 5 m)

at 4 mg lipid ml−". α-crystallin (Sigma Chemical Co.,

St. Louis, MO, U.S.A., 12% other crystallins) was

dissolved at a concentration of 1±2 mg ml−" in Hepes

buffer (pH 7±4, 5 m). The concentration was verified

by a modified Lowry assay (Peterson, 1977) designed

for membrane protein samples. The secondary struc-

ture of the α-crystallin was determined by infrared

spectroscopy (Lamba et al., 1993). Six solutions

containing 4 mg lipid each were incubated overnight

at 36°C with increasing amounts of protein (0–1±2mg) in a total volume of 2 ml. Each sample was then

spun at 170000 g for 1 hr. The supernatant was

decanted and assayed for protein. Infrared spectra of

some of the pellets containing the highest levels of

protein were measured to assess lipid structural order.

Spectroscopic instrumentation. Infrared spectra

were acquired with a Mattson model 5000 FTIR

spectrometer equipped with a TGS detector. Approxi-

mately 250 interferograms were recorded, coadded,

and apodized with a Happ–Genzel function prior to

Fourier transform, yielding an effective resolution of

1±0 cm−". Temperature was monitored to ³0±4°C by a

copper-constantan type T thermocouple and LFE

Instruments (Chesterland, OH, U.S.A.), model 3000

controller. Temperature was maintained at 36 °Cusing a Specac Limited (Fairfield, CT, U.S.A.) variable

temperature cell (model 21500). All procedures of

spectral routines have been described in our recent

publications (Borchman et al., 1993; Lamba et al.,

1993). Signal averaging, data smoothing using the

Savitsky–Golay procedure, baseline correction, buffer

subtraction, and related spectral routines were per-

formed with the Grams}386 software (Salem, NH,

U.S.A.).

Intrinsic Protein–Lipid Interactions Using Human Lens

Membranes

Human lens tissue. Clear human lenses were

obtained within 8 hr of death through the Kentucky

Lions Eye Bank. The epithelium, cortex and nucleus

were dissected. Three regions were pooled according

to age as follows: pool I, 0 to 15 years old (n¯16);

pool II, 16 to 30 (n¯34); pool IV, 46 to 60 (n¯42);

pool V, 61 to 75 (n¯91); and pool VI, 76 and older

(n¯44).

Urea treatment of membrane. The protocol of

Broekhuyse and Kuhlman (1978) was followed to

prepare urea-insoluble membranes enriched in in-

trinsic proteins. All reagents were purchased from the

Sigma Chemical Company. Lens tissue was homo-

genized, using a Teflon douncer in 0±1 Tris

(hydroxymethyl)aminomethane (Tris–HCL) at pH¯7±4 and 5 m dithiothreitol (DDT). One microliter of

buffer per lens cortical tissue was used. The homo-

genate was spun at 10000 g for 30 min at 10°C. The

pellet was resuspended in 7 urea, 0±1 M Tris–HCl,

pH¯7±4 with 5 m DTT using the same volume as

for the initial homogenization. The homogenate was

diluted seven-fold with 0±1 Tris–HCl, pH¯7±4,

5 m DTT and spun at 80000 g for 30 min. The

supernatant was decanted and the pellet resuspended

in 0±1 Tris–HCl, pH¯7.4, 5 m DTT by homo-

genization with a Teflon douncer and spun at 80000

g for 30 min. The pellet was resuspended and spun

three more times. The final pellet was placed in a

capillary tube for the FT-Raman structural study.

FT-Raman spectroscopy of urea-treated membrane. The

FT-Raman (FTR) system was a Jeol JRS-FT 6500N

spectrometer equipped with an InGaAs detector. The

excitation wavelength at 1064 nm was provided by a

cw Nd:YAG laser (Spectron SL301), and the laser

power at the sample position was typically 500 mW.

Raman spectra were obtained with a spectral res-

olution of 4 cm−", and 500 scans were accumulated to

obtain an acceptable signal-to-noise ratio.

All numerical manipulations and data treatment

were performed with Grams}386 software (Salem, NH,

U.S.A.). Spectra were 11-point smoothed using the

Savitsky–Golay procedure.

CH#

stretching band and lipid order. Lens material

for this study came from the same pool as that used in

LENS LIPID–PROTEIN INTERACTIONS 49

F. 1. $"P-NMR spectrum of bovine lipid extracted from the lens cortical region. U¯unidentified; PEplas¯phosphatidylethanolamine plasmalogen; PE¯ phosphatidylethanolamine; PC¯ phosphatidylcholine; PS¯ phosphatidyl-serine ; SM¯ sphingomyelin ; PG¯ phosphatidylglycerol.

Raman shift (cm–1)

2900 2850 2800

γ

F. 2. FT-Raman spectra of the CH#

stretching regionof (------) dipalmitoylphosphatidylcholine solid, which iscompletely ordered, (——) lens cortical lipid completelydisordered in chloroform.

published lipid structural and compositional studies

(Borchman et al., 1994a, 1994b, 1995). In this study

the lipid hydrocarbon order of membranes with

intrinsic protein is compared to that of lipid mem-

branes without protein to determine if the membrane

proteins alter lipid structure. The Raman spectral

features used to characterize lipid order include the CH

stretching band near 2850 cm−", which is the largest

band in the Raman spectra of all lipids, both native

and synthetic. Figure 2 depicts the CH stretching

region for a typical lens lipid disordered in chloroform

and for dipalmitoylphospatidylcholine that is com-

pletely ordered. The abscissa unit is wavenumbers

and, in general terms, is proportional to the frequency

of a vibration, in this instance the C–H stretching

vibration. The ordinate shows the relative number of

molecules vibrating at a given frequency. Note that

when lipids become more disordered, the band

frequency expressed in wavenumbers increases. The

frequency changes in this band have been used in our

FTIR and FTR studies to estimate lipid hydrocarbon

chain order (Borchman et al., 1993, 1994, 1995).

In addition to a shift to higher wavenumbers when

lipid acyl chains become more disordered, the intensity

of the band at 2880 cm−" decreases. The 2880 cm−"

Raman band is a Fermi resonant band, sensitive to

intra- as well as interchain interactions (Lamba et al.,

1993). The intensity of this band decreases as intra-

and intermolecular chain disorder increases. The peak

height intensity ratio I#))!

}I#)&!

, has been used as an

order parameter to determine the relative hydrocarbon

chain ‘fluidity ’ of a membrane (Gaber and Peticolas,

1977).

3. Results

Extrinsic Protein–Lipid Interactions Using Bovine Lipids

and Proteins

α-Crystallin–lipid binding model. As a model for the

study of structural changes caused by extrinsic

proteins binding to lens lipids, α-crystallin was bound

to lens lipids. Figure 3 shows the maximum weight

ratio of α-crystallin to lipid in this study was 0±3. This

corresponds to 164 g protein per mole of lipid

(phospholipid}cholesterol, 1 :1 mole ratio). In a pre-

vious study, Ifeanyi et al. (1991) demonstrated that

phosphatidylcholine vesicles bound a maximum of

31±3 g of protein per mole of phospholipid. Bovine

lipids in this study were capable of binding five times

more α-crystallin, perhaps because of differences in

technique or the differences in the composition of the

vesicles. After centrifugation, all of the α-crystallin

bound to the vesicles leaving no α-crystallin in the

supernatant [Fig. 3(b)]. Centrifugation alone could not

bring the α-crystallin out of solution (Fig. 3). Fifteen

hours were required to complete the binding process.

50 H. SATO ET AL

0.4

–0.10.0

(B) Protein/lipid (w/w)

O.D

. (75

0 n

m)

0.3

0.2

0.1

0.0

0.1 0.2 0.3 0.4

0 0.4 0.8 1.2 0.4

(A) [Protein] (mg ml–1)

(a)

(b)

F. 3. Optical density of supernatant after centrifugationof sample ; (A) aqueous bovine α-crystallin ; (B) aqueousα-crystallin at the same concentration as (A), with 4 mgbovine lens cortical lipid.

Raman shift (cm–1)

3100 2900 28003000

2924

2852

(B)

(A)

F. 4. Fourier transform infrared spectra of (A) pellet fromcentrifugation of α-crystallin plus bovine lens cortical lipidsolution. Pellet contained 0±3 protein}lipid (w}w); (B) pelletfrom centrifugation of bovine lens cortical lipid.

To determine whether lipid structure was changed

due to protein binding, infrared spectra of the vesicle

pellets containing protein were acquired. The bands in

the CH stretching region, 3100 to 2800 cm−", are very

sensitive to lipid order and have been used to probe the

lipid order of human, and rabbit lens lipids (see

Methods for details). Hydrocarbon chain lipid order is

defined structurally in terms of the freedom of rotation

about the carbon–carbon hydrocarbon bonds. Figure

4 shows the infrared spectra of vesicle pellets with [Fig.

4(A)] and without [Fig. 4(B)] protein bound to the

CH2 stretch

(i)

(ii)

(iii)

2800 2700 2600 2500

Wavenumber (cm–1)

(i)

(ii)

(iii)1700 1600 1400 1200

Am

ide

1

Tyr

osin

e

Lipidamide

Tryptophan Random

Hel

ix Tyr

osin

e

(iii)

Sheet

Helix

1700 1650

(A)

(B)

(C)

F. 5. Fourier transform infrared spectra of : i, urea-purified membrane; ii, aqueous dispersion of lipid extractedfrom urea purified membrane; iii, spectrum i®spectrum ii.(A) CH stretching region, (B) fingerprint region, (C) amide Iband.

pellet. Careful analysis of the spectra revealed no

differences in the CH#

symmetric stretching band

frequency, which indicates there was no change in

lipid order, even at high lipid to protein ratios of 164

g α-crystallin per mole lipid (cholesterol}phospholipid,

1:1 mole ratio). α-Crystallin has no affect on the

conformation of the lipid chains.

Intrinsic Protein–Lipid Interactions Using Human Lens

Membranes

The CH#

symmetric stretching band was used to

compare the lipid order of lens lipid membranes with-

out protein, with urea-purified membranes enriched

with intrinsic proteins (see Materials and Methods).

Lipid contributes about 80% the CH stretching

band intensity in the Raman spectra of membranes,

Fig. 5(A). Fig. 5(A)i shows the spectral region

LENS LIPID–PROTEIN INTERACTIONS 51

100

2851

28490

Age (years)

Pea

k ce

nte

r of

gra

vity

(w

ave

nu

mbe

rs)

20 40 60 80

2850

Moreordered

Morefluid

F. 6. The center of gravity was measured for the CH#

symmetric stretching band and is proportional to theconformational order of the lipid hydrocarbon chains. At2854±5 cm−" the lipid is completely disordered. At 2849cm−" the lipid is completely ordered. (D) human lens corticallipid ; (E) urea-purified human lens membranes.

corresponding to the CH#stretching Raman bands for

urea purified cortical membranes in aqueous buffer.

The frequency corresponding to the center of gravity

of the CH#

symmetric stretching band, measured in

wave numbers, was shown to decrease with increasing

age for the cortical lipids (no protein) in aqueous

buffer, Fig. 6. There was no statistical difference

between lipid order measured with or without intrinsic

membrane protein present (Fig. 6). Four of the six

urea-treated membrane samples exhibited lipid orders

within the dotted 95% confidence limits (Fig. 6) and

the order of these membranes is therefore indis-

tinguishable from those measured for lipids without

protein. In terms of lipid order, completely ordered

lipids have a CH#symmetric band frequency of 2847±2

cm−". Lens lipids completely disordered in CHCl$have

a CH#symmetric band frequency of 2854 cm−". Using

these numbers to estimate the percentage lipid order,

the maximum difference between lens membranes

with and without protein was, at most, 12%. The

averaged difference was 4±0%, comparable to the

standard deviation of the percentage order estimation

for the six samples measured. Lipid order may range

from 0 to 100%, so a difference of 4% order is small

and because of the experimental deviation, this

difference can not be considered significant.

The infrared spectral region corresponding to the

amide I band is shown in Fig. 5(B). This band arises

from the protein backbone amide moiety. Protein

secondary structure can be estimated by mathematical

treatment of the amide I band (Lamba et al., 1993).

We determined the secondary structure of bovine

crystallins using infrared spectroscopy and found that

the crystallin proteins were predominantly β-sheet

and contained less than 10% α-helix (Lamba et al.,

1993). Using a similar approach, we estimated the

average secondary structure of the intrinsic membrane

proteins in our samples. The estimation of secondary

structure of membrane proteins is more complicated

than that of soluble proteins because the lipid amide

band from sphingolipids interferes with the protein

amide I band. By curve fitting the amide I band of the

spectra with lipid and protein, we estimate that

41%³14 (.., n¯6) of the membrane amide I band

intensity is due to sphingolipids. Comparison of the

spectrum in Fig. 5(B)i with that in Fig. 5(B)ii, one may

see that the spectrum in Fig. 5(B)i contains a

considerable number of protein bands which indicates

a large mass of the sample is presumably due to

membrane intrinsic protein. To remove the con-

tribution of the lipid from the spectrum of lipid and

protein, Fig. 5(B)i, the spectrum of pure lipid, Fig.

5(B)ii was subtracted. The results of this subtraction

are shown in Fig. 5(B)iii.

By curvefitting the amide band as illustrated in Fig.

5(C), we estimate that the proteins are predominately

α-helical (72³15%), 18³11% β-sheet, 6³7% turns

and 4³4% other structure. Bands near 1680, 1655,

1634 and 1601 cm−" were assigned to turn, α-helix,

β-sheet and amino acid residues, respectively. The

standard deviation of the mean was calculated from

six different samples. No age-related correlations were

observed. There was no difference in the secondary

structure determined by analysis of the original amide

I band or of the Fourier self deconvolved amide bands.

The intensity at 1274 cm−" in the amide III region

[Fig. 5(B)iii] confirms the large α-helical content as

calculated from the amide I band (Warren et al.,

1976). The small amount of β-sheet and large α-

helical content indicate that the proteins in the

membrane preparation must be mostly intrinsic, since

all of the crystallins are predominately β-sheet (Lamba

et al., 1993). The 18% β sheet component may come

from crystallins shown to be present in urea mem-

branes (Russel et al., 1981). Since extrinsic proteins do

not influence lipid structure, they do not interfere with

our assessment of the effect of intrinsic proteins on

lipid structure.

4. Discussion

Weight ratios of about 1 protein to 1 phospholipid

have been reported for membrane preparations such

as the one used in this study (Takemoto and Rintoul,

1983; Roy et al., 1982; Broekhuyse and Kuhlmann,

1978, 1979). Assuming that a 100000 molecular

weight protein may contact 30 annular (the lipids in

contact with the protein) lipids (Warren et al., 1976)

(the main intrinsic protein 26 contacts eight lipids) we

calculate that 25% of the lens membrane lipids would

be in contact with the intrinsic protein. From a static

structural perspective, lipid hydrocarbon order was

not perturbed by either extrinsic or intrinsic proteins.

52 H. SATO ET AL

Even at a high 1:1 weight ratio of extrinsic protein to

lipid, no statistical differences in lipid hydrocarbon

chain order were detected. Our results are unlike those

reported by Mateau et al. (1978) which show that

extrinsic proteins disorder membranes and intrinsic

proteins order membranes. We define lipid order in a

structural sense and should not be confused with

other definitions of lipid order based on lipid mobility

or wobble. Liang et al. (1989) have shown that lens

membrane proteins do immobilize lens lipids. This has

been shown for many proteins (Devaux and

Seigneuret, 1983).

We have shown that lipid order increases with age

and cataract. These studies were performed on lipid

vesicles without protein present. It was important to

determine if proteins influence lipid order and if the

changes in lipid order reported with age and cataract

were to compensate for structural perturbations

caused by increases in membrane associated protein.

It is well known that the association of proteins with

the membrane insoluble fraction increases with

increasing age and cataract. We learn from this study

that protein, at concentrations found in human lenses,

has no influence on lipid hydrocarbon chain structure.

The increase in lipid structural order with age may be

associated with phosphatidylcholine and sphingo-

myelin compositional changes (Borchman et al.,

1994a, 1994b). The structure of lipid hydrocarbon

chains in lens lipid vesicles containing no protein is

likely to be similar to the structure in vivo where

proteins are present.

Acknowledgements

Supported by Public Health Service research grantsEY07975 and EY06916 (Bethesda, MD, U.S.A.) and theKentucky Lions Eye Foundation (Louisville, KY, U.S.A.), andan unrestricted grant from Research to Prevent Blindness,Inc. Christopher A. Paterson, Ph.D., D.Sc. is a Research toPrevent Blindness Senior Scientific Investigator.

References

Borchman, D., Byrdwell, W. C. and Yappert, M. C. (1994a).Regional and age-dependent differences in the phospho-lipid composition of human lens membranes. Invest.Ophthalmol. Vis. Sci. 35, 3938–42.

Borchman, D., Yappert, M. C., Byrdwell, W. C. and Lamba,O. P. (1994b). Relationships between lipid compositionand membrane structure in ocular and other tissues.Exp. Eye Res. 59 (suppl. 1), S105.

Borchman, D., Lamba, O. P. and Yappert, M. C. (1993).Structural characterization of human lens membraneclear and cataractous lipid. Exp. Eye Res. 57, 199–208.

Broekhuyse, R. M., Kuhlmann, E. D. and Jap, P. H. K.(1979). Lens membranes. IX. Some characteristics offiber membranes in relation to aging and cataractformation. Ophthalmic Res. 11, 423–8.

Broekhuyse, R. M. and Kuhlmann, E. D. (1978). Lensmembranes. IV. Preparative isolation and charac-terization of membrane and various membrane proteinsfrom calf lens. Exp. Eye Res. 26, 305–20.

Devaux, P. F. and Seigneuret, M. (1983). Specificity oflipid–protein interactions as determined by spectro-scopic techniques. Biochem. Biophys. Acta 822, 63–125.

Folch, J., Lees, M. and Sloane-Stanley, G. H. (1957). Asimple method for the isolation and purification of totallipids from animal tissues. J. Biol. Chem. 226, 497–509.

Gaber, B. P. and Peticolas, W. L. (1977). On the quantitativeinterpretation of biomembrane structure by Ramanspectroscopy. Biochim. Biophys. Acta 465, 260–74.

Gooden,M. M., Mendelson, R., Sunder, S. and Bernstein,H. J. (1976). The effect of sonication on the hydrocarbonchain conformation in model membrane systems: aRaman spectroscopic study. Biochim. Biophys. Acta 419,563–9.

Ifeanyi, F. and Takemoto, L. (1990). Specificity of α-crystallinbinding to the lens membranes. Curr. Eye Res. 9,259–65.

Ifeanyi, F. and Takemoto, L. (1989). Differential binding ofα-crystallins to bovine lens membrane. Exp. Eye Res.49, 143–7.

Ifeanyi, F. and Takemoto, L. (1990). α-Crystallin fromhuman cataractous versus normal lenses : change inbinding to lens membrane. Exp. Eye Res. 50, 113–16.

Ifeanyi, F. and Takemoto, L. (1991a). Interaction of lens α-crystallin with lipid vesicles. Exp. Eye Res. 53, 535–8.

Ifeanyi, F. and Takemoto, L. (1991b). Involvement of the N-terminal region in α-crystallin-lens membrane rec-ognition. Exp. Eye Res. 52, 305–8.

Ifeanyi, F. and Takemoto, L. (1991c). Characterization of themajor cyanogen bromide fragment of α-crystallin. Curr.Eye Res. 10, 529–35.

Lamba, O. P., Borchman, D., Sinha, S. K., Yappert, M. C.,Shah, J. and Renugopalakrishnan, V. (1993). Esti-mation of the secondary structure and conformation ofbovine lens crystallins by infrared spectroscopy: quan-titative analysis and resolution by Fourier self-deconvolution and curve fit. Biochim. Biophys. Acta1163, 113–23.

Liang, J. N., Rossi, M. R. and Andley, U. P. (1989). Flu-orescence studies on the age related changes in bovineand human lens membrane structure. Curr. Eye Res. 3,293–8.

Liang, J. J. N. and Li, X. (1992). Spectroscopic studies on theinteraction of calf lens membranes with α-crystallins.Exp. Eye Res. 54, 719–24.

Mateu, L., Caron, F., Luzzati, V. and Billecocq, A. (1978).The influence of protein-lipid interactions on the order-disorder conformational transitions of the hydrocarbonchain. Biochim. Biophys. Acta 508, 109–21.

Mulders, J. W. M., Stokkermans, J., Leunissen, J. A. M.,Benedetti, E. L., Bloemendal, H. and De Jong, W. W.(1985). Interaction of α-crystallin with lens plasmamembranes. Eur. J. Biochem. 152, 721-8.

Mulders, J. W. M., Wojcik, E., Bloemendal, H. and De Jong,W. W. (1989). Loss of high-affinity binding of bovinenuclear α-crystallin. Exp. Eye Res. 49, 149–52.

Peterson, G. L. (1977). A simplification of the protein assaymethod of Lowry et al. which is more generallyapplicable. Anal Biochem. 83, 346–56.

Puskin, J. S. and Wiese, M. B. (1982). A spin label study ofhuman lens membranes. Exp. Eye Res. 35, 251–8.

Roy, D., Rosenfeld, L. and Spector, A. (1982). Lens plasmamembrane: isolation and biochemical characterization.Exp. Eye Res. 35, 113–29.

Russel, P., Robison, W. G. and Kinoshita, J. H. (1981). Anew method for rapid isolation of the intrinsic mem-brane protein from lens. Exp. Eye Res. 32, 511–16.

Takemoto, L. J., Rintoul, D. A. (1983). Evidence for reducedlipid order in plasma membranes from cataractoushuman lenses. Curr. Eye Res. 2, 367–75.

LENS LIPID–PROTEIN INTERACTIONS 53

Tu, A. T., Prescott, B., Chou, C. H. and Thomas, Jr., G. J.(1976). Structural properties of mojave toxin of crotaluscululatus (mojave rattlesnake) determined by laserRaman spectroscopy. Biochem. Biophys. Res. Comun. 68,1139–45.

Warren, G. B., Toon, P. A., Birdsall, N. J. M., Lee, A. G. and

Metcalfe, J. C. (1976). Reversible lipid titration of theactivity of pure adenosine triphosphatase-lipidcomplexes. Biochemistry 13, 5501–7.

Zhang, W. Z. and Augusteyn, R. C. (1994) On the in-teraction of α-crystallin with membranes. Curr. Eye Res.13, 225–30.