• Aucun résultat trouvé

Calcium Causes a Conformational Change in Lamin A Tail Domain that Promotes Farnesyl-Mediated Membrane Association

N/A
N/A
Protected

Academic year: 2021

Partager "Calcium Causes a Conformational Change in Lamin A Tail Domain that Promotes Farnesyl-Mediated Membrane Association"

Copied!
9
0
0

Texte intégral

(1)

Calcium Causes a Conformational Change in Lamin A Tail Domain

that Promotes Farnesyl-Mediated Membrane Association

The MIT Faculty has made this article openly available.

Please share

how this access benefits you. Your story matters.

Citation

Kalinowski, Agnieszka, Zhao Qin, Kelli Coffey, Ravi Kodali, Markus J.

Buehler, Mathias Losche, and Kris Noel Dahl. “Calcium Causes

a Conformational Change in Lamin A Tail Domain That Promotes

Farnesyl-Mediated Membrane Association.” Biophysical Journal

104, no. 10 (May 2013): 2246–2253. © 2013 Biophysical Society

As Published

http://dx.doi.org/10.1016/j.bpj.2013.04.016

Publisher

Elsevier

Version

Final published version

Citable link

http://hdl.handle.net/1721.1/89214

Terms of Use

Article is made available in accordance with the publisher's

policy and may be subject to US copyright law. Please refer to the

publisher's site for terms of use.

(2)

Calcium Causes a Conformational Change in Lamin A Tail Domain that

Promotes Farnesyl-Mediated Membrane Association

Agnieszka Kalinowski,

†6

Zhao Qin,

‡6

Kelli Coffey,

†§

Ravi Kodali,

{

Markus J. Buehler,

*

Mathias Lo¨sche,

†k

***

and Kris Noel Dahl

†§

*

Biomedical Engineering, Carnegie Mellon University, Pittsburgh, Pennsylvania;Civil and Environmental Engineering, Massachusetts

Institute of Technology, Cambridge, Massachusetts;§Chemical Engineering, Carnegie Mellon University, Pittsburgh, Pennsylvania;

{Department of Structural Biology, University of Pittsburgh, Pittsburgh, Pennsylvania;kDepartment of Physics, Carnegie Mellon University,

Pittsburgh, Pennsylvania; and **Center for Neutron Research, NIST, Gaithersburg, Maryland

ABSTRACT

Lamin proteins contribute to nuclear structure and function, primarily at the inner nuclear membrane. The

post-translational processing pathway of lamin A includes farnesylation of the C-terminus, likely to increase membrane association,

and subsequent proteolytic cleavage of the C-terminus. Hutchinson Gilford progeria syndrome is a premature aging disorder

wherein a mutant version of lamin A, D50 lamin A, retains its farnesylation. We report here that membrane association of

far-nesylated D50 lamin A tail domains requires calcium. Experimental evidence and molecular dynamics simulations collectively

suggest that the farnesyl group is sequestered within a hydrophobic region in the tail domain in the absence of calcium. Calcium

binds to the tail domain with an affinity

K

D

z 250 mM where it alters the structure of the Ig-fold and increases the solvent

acces-sibility of the C-terminus. In 2 mM CaCl

2

, the affinity of the farnesylated protein to a synthetic membrane is

K

D

z 2 mM, as

measured with surface plasmon resonance, but showed a combination of aggregation and binding. Membrane binding in the

absence of calcium could not be detected. We suggest that a conformational change induced in D50 lamin A with divalent

cat-ions plays a regulatory role in the posttranslational processing of lamin A, which may be important in disease pathogenesis.

INTRODUCTION

A-type lamins, including lamin A (LA), contribute to a

variety of functions that include DNA transcription,

hetero-chromatin regulation, cell senescence, DNA damage repair,

and the sequestration of transcription factors (recent reviews

(

1,2

)). These type V intermediate filament proteins are also

the primary structural elements of the nucleoskeleton (

3

).

Lamins are structurally related to other intermediate

fila-ment proteins (

4

), but the C-terminal tail domain (TD) of

LA is much larger than that of other intermediate filaments

and undergoes significant posttranslational processing.

LMNA codes for the precursor prelamin A (preLA), which

is transported into the nucleus where the C-terminal CaaX

domain (specifically, Cys-Ser-Ile-Met) is farnesylated and

carboxymethylated (

5

). The modified preLA is then cleaved

by the enteroprotease ZMPSTE-1 at Y646 to produce

mature LA. Dysfunctions of this process including loss of

ZMPSTE-1 or mutations including the Y646 cleavage site

are responsible for devastating human diseases. In the

pre-mature aging syndrome Hutchinson Gilford progeria

syn-drome (HGPS), a C1824T mutation of LMNA causes the

activation of a cryptic splice site and a loss of exon 11 in

the TD (

6

). The loss of the 50 amino acid segment, which

contains the ZMPSTE-1 cleavage region, causes retention

of the farnesylation, which in turn results in changes in

ther-modynamic stability of the mutant protein,

D50LA or

progerin (

7

). At a much lower frequency than in HGPS,

acti-vation of the cryptic splice site that produces

D50LA occurs

sporadically in normal cells, and this defect accumulates

with aging (

8

).

The biochemistry of LA TD processing has been

deter-mined, but the biological benefits of the complex processing

pathway and the regulation of its individual steps are unclear

(

9

). Although farnesyl transferases have been found in both

the endoplasmic reticulum and the nucleus, farnesylation

of the protein likely occurs in the nucleus (

5

). The farnesyl

modification is suggested to increase the association of

preLA with the inner nuclear membrane (

10,11

). Similarly,

the pathologies associated with the expression of

D50LA in

HGPS result, in part, from the farnesylation causing

overac-cumulation of structural proteins at the inner nuclear

membrane (

12,13

). Indeed, mislocalization of LA and

LA-associated proteins at the inner nuclear membrane results

in altered DNA repair, transcription, and signaling (

14

), as

well as changes in protein-protein interactions (

15

). On the

other hand, a single farnesylation is insufficient to confer

permanent membrane association of proteins in other

sys-tems (

16–18

). Therefore, it is unclear if the presence of the

farnesyl is sufficient for the cellular pathological phenotype

or if additional factors, including altered proteprotein

in-teractions (

15

) or altered protein packing (

7,19

) are involved.

For the full-length lamin protein, the rod domain

func-tions in dimer formation and the head and tail domains

asso-ciate with the rod to form filaments; full-length filaments

form aggregates and paracrystals in vitro (

1–4

). The TD is

Submitted November 29, 2012, and accepted for publication April 8, 2013.

6Agnieszka Kalinowski and Zhao Qin contributed equally to this work.

*Correspondence:mbuehler@mit.eduorquench@cmu.eduorkrisdahl@ cmu.edu

Editor: David Odde.

(3)

responsible for association with other proteins and

mem-brane localization. For these reasons, we limited our study

to the TD to isolate the interaction between the protein

and the membrane (

2,10

). Here, we quantify

protein-mem-brane binding of farnesylated

D50LA (fn-D50LA) by

inves-tigating purified LA TDs in solution and on model lipid

bilayer membranes. We find that the binding is weak

(K

D

~2

mM) and that free Ca

is required for binding.

The conformational response of the protein to Ca

is

tracked by monitoring its intrinsic fluorescence. Molecular

simulations suggest that the Ca

-induced conformational

change opens the Ig-fold, increases the flexibility of the

terminal CaaX motif, and exposes the farnesyl group, which

is sequestered within the TD in the absence of the ion.

MATERIALS AND METHODS

Protein production, farnesylation,

and characterization

Constructs for the expression of LA TDs were created using polymerase chain reaction from plasmids kindly provided by Misteli at NCI/NIH (12). Plasmids were expressed in Escherichia coli, purified using gluta-thione-s-transferase fusion peptides and farnesylated in vitro using farnesyl-transferase, resulting in a population of ~50% farnesylated protein (see Methods in theSupporting Material). Purified proteins were verified for size analysis using liquid chromatography electron spray ionization mass spectroscopy, for purity using 14% SDS-PAGE, for aggregation by dynamic light scattering (DLS) (nanosizer). Bradford assay (Coomaise Plus, Pierce) was used to determine protein concentration (see Methods).

Solution conditions that minimize aggregation

of

D50LA and preLA TDs

Recombinant LA-TD variants aggregate in water, therefore accurate quan-tification of protein-membrane interactions required stabilization of mono-mer protein. We determined the minimal salt (NaCl) concentrations in the buffer necessary to suppress aggregation ofD50LA-TD and preLA-TD by DLS. Although deconvoluted DLS spectra were not able to determine aggregation form, it clearly showed large-scale aggregation under low-salt conditions (Fig. S1in theSupporting Material). Characteristic DLS peaks of d ~10 nm were regularly observed under high-salt conditions ([NaCl]> 50 mM), and we interpreted a 10 nm peak as a protein monomer. At low salt ([NaCl]< 50 mM), proteins aggregated (d > 100 nm). This analysis showed different minimal salt concentrations for the different lamin A variant TDs studied here, including with farnesylation(fn). D50LA-TD required 25 mM NaCl to maintain a monomeric solution of pro-tein, whereas fn-D50LA-TD required 50 mM NaCl (Fig. S1A). In contrast, preLA-TDs required a minimum of 500 mM NaCl to maintain dispersed solutions (Fig. S1B). The addition of 2 mM CaCl2did not alter the stability

ofD50LA-TD and fn-D50LA-TD (stable in 50 mM NaCl and 2 mM CaCl2, Fig. S1C) or preLA-TD and fn-preLA-TD (stable in 500 mM NaCl and 2 mM CaCl2, Fig. S1 D). Aggregation could also be suppressed in

10 mM sodium phosphate buffer but could not be used in these experiments because phosphate chelates calcium. Other salts, such as KCl or NaHCO3,

were not tried as membrane behavior under those conditions.

Tryptophan fluorescence

The tryptophan amino acid residue has inherent fluorescent properties that have long been used as probes to reflect protein conformation (20). LA-TD

has four tryptophan residues all of which are located on or near the Ig-fold. The peak emission intensity is at 342 nm, after excitation at 298 nm to mini-mize excitation of other aromatic residues. A peak emission at 342 nm indicates that the tryptophan residues are neither in a completely hydropho-bic environment, nor are they fully solvent exposed (20,21).

The peak emission intensity at 342 nm (with the background intensity from the buffer subtracted) was used for analysis and normalized by the peak intensity without Ca2þfor each sample. The binding affinity and stoichiometry of Ca2þassociation with the LA TD was determined by fitting the change in fluorescence intensity at 342 nm as a function of Ca2þconcentration to a Hill model. For our system where the ligand is the Ca2þand the conformational change is assumed to be proportional to the change in fluorescence intensity at the emission peak, DI (342 nm), normalized by the peak emission at 342 nm without Ca2þ in the solution, Io, the Hill equation is DI/Io ¼ [Ca2þ]

n

/(KDþ[Ca2þ] n

) gives a binding affinity of the protein to calcium KD, as well as the

stoichiometry of binding, n (i.e., the number of Ca2þthat associate with one protein molecule).

Membrane bilayer preparation and surface

plasmon resonance experiments (SPR)

SPR was used to quantify the membrane association of the LA-TD to teth-ered bilayer lipid membranes (tBLMs), which provide a fluid membrane fixed to a surface for SPR to monitor membrane association without the necessity of using a protein label. The tBLM model membrane system consists of a planar lipid bilayer suspended to the surface by a minimal number of functionalized lipopolymer interspersed with b-mercaptoetha-nol to create a submembrane hydration layer (see Methods). The ~1.5 nm hydration layer allows the membrane to retain its fluidity and reduce the effects of the substrate on the membrane. See Methods for full details. The molecular structure, functionality, and in-plane dynamics of such tBLMs have been established by the Lo¨sche group in extensive studies involving electrochemical impedance spectroscopy, neutron scat-tering, fluorescence microscopy, and fluorescence correlation spectroscopy (22–24).

Replica exchange molecular dynamics (REMD)

simulations

Protein structure predictions were carried out by the REMD method (25) implemented in the CHARMM c35b1 (26) software. Replica management and conformation analysis are done using the MMTSB script package (27). REMD is a simulation method with atomic detail to improve the dynamic properties of the Monte Carlo method, aiming at a global-minimum free energy state. It samples a wide conformation space without getting trapped in local-minimum free energy states because of the thermal stimulation by high-temperature replicas, and we can find stable structures at the replica with the lowest temperature. The REMD inputs include the peptide sequence and an initial guess of conformation (see Methods).

Modeling of farnesyl diphosphate

The structure and interaction scheme of the farnesyl diphosphate is not defined in the standard CHARMM27 force field. We setup the initial conformation of farnesyl diphosphate according to its atomic coordinates as found in the protein data bank (with PDB ID 1KZO). The interactions (covalent bond, angle, dihedral, electrostatic, and vdW interactions) among the atoms in farnesyl diphosphate are defined in the CHARMM General force field 2b6 (28). We updated the CHARMM27 force field to incorporate the section of the definition of farnesyl diphosphate (see Methods for details of the modeling and validation against the atomic structures identified from experiment).

Biophysical Journal 104(10) 2246–2253

(4)

RESULTS

Membrane binding of fn-

D50LA-TD depends on

calcium

To determine the membrane affinity of fn-

D50LA-TD, we

added

purified

monomeric

farnesylated

protein

(see

Methods and

Supporting Material

information for

produc-tion, characterizaproduc-tion, and monomeric stability) to tBLMs

and measured binding using SPR. The tBLMs are

appro-priate for precise SPR binding measurements because

bila-yers are virtually defect free, thus avoiding nonspecific

protein adsorption (

29

) and tightly coupled to gold-coated

solid substrates via hydrated oligo(ethylene oxide) spacers

(

30

). Bilayer membranes were composed of

dioleoyl-phos-phatidylcholine and the charged

dioleoyl-phosphatidylser-ine (DOPC/DOPS

¼ 7:3) with 3 mol% cholesterol to

mimic lipid composition and charge of cellular membranes.

With or without farnesylation,

D50LA-TD associated

minimally with the membrane in NaCl buffer (

Fig. 1

A).

However, the presence of physiological levels of free

Ca

in the buffer facilitated protein-membrane interaction.

fn-

D50LA-TD bound stably to tBLMs with 2 mM CaCl

2

in

the buffer, but

D50LA-TD binding was not detected under

otherwise identical conditions (

Fig. 1

, A and B). Increasing

concentrations of fn-

D50LA-TD in the Ca

-containing

buffer increased membrane binding (

Fig. 1

C) and showed

a binding affinity, K

D

~2

mM (assuming that only the

farne-sylated proteins associated with the membrane, i.e.,

approx-imately half the concentration of the protein added) when fit

to a Langmuir model (

Fig. 1

C). We note that the membrane

association is a combination of binding and aggregation,

which impacts the K

D

: instantaneous high levels of

fn-

D50LA-TD do not show the same levels of binding as

increased amounts of fn-

D50LA-TD (

Fig. 1

B vs.

Fig. 1

C). Furthermore, protein is not easily dissociated from the

interface. These and other aspects of interfacial aggregation

and binding are the topic of future work. Unfortunately,

the high salt required to maintain monomeric stability of

fn-preLA-TD (see Methods) interfered with membrane

stability, so that fn-preLA-TD binding to membranes could

not be quantified.

Calcium-dependent conformational changes

To investigate the mechanism by which Ca

induces

mem-brane association, we considered Ca

-induced

conforma-tional changes of the TD. The TD is intrinsically

disordered except for the s-type Ig-fold (

31,32

), therefore,

direct structural examination via traditional methods (e.g.,

crystallography) are not possible. Conformational changes

of

D50LA-TD and preLA-TD were followed by monitoring

the fluorescence of their four Trp residues located within

and near the Ig-fold (W467, W498, W514, W520; see full

sequence in

Table S1

). The chemical environment of Trp,

including its solvent accessibility, affects the emission

quan-tum yield, and exposure to the solvent has been shown to

quench Trp fluorescence (

20,21

). Previously, we

investi-gated thermal denaturation of LA-TD and

D50LA-TD

ter-tiary structure by monitoring Trp fluorescence, confirming

that protein denaturation reduces Trp fluorescence (

7

).

The observed fluorescence emission band between 310

and 400 nm was broad, presumably because multiple Trp

residues in different microenvironments contribute to the

band shape. We observed that the emission decreased with

CaCl

2

concentration (

Fig. 2

, A and B A).), and saturated for

[CaCl

2

]

z 1 mM at ~30% of the original intensity. Fitting

to a two-parameter Hill model indicated a Ca

affinity of

261

5 49 mM for D50LA-TD and 286 5 27 mM for preLA

TD, both with 1:1 stoichiometries (

510%) (

Fig. 2

C). These

results suggest that Trp residues of the two TDs become more

exposed to buffer upon Ca

binding. High concentrations of

NaCl (

>200 mM) also reduced Trp emission (

Fig. S2

), but

Ca

-induced effects occur at six orders of magnitude lower

salt concentration (100

mM), independent of NaCl.

10−2 10−1 100 101

0 0.5 1

[fn−Δ50LA−TD], μM

Protein adsorbed (saturated fraction)

2 6 10 0 100 200 Time, hr Density (ng/cm 2) 0 3 6 0 100 200 Time, hr Density (ng/cm 2) 0 1 2 0 100 200 Time, hr Density (ng/cm 2)

A

B

C

D

.01F 0 .05F 0.1F 0.5F 1F 2F 0 1 0.1F 1F 2F 0 1.5 4 3.2F +Ca2+ +Ca2+

−Ca2+ +Ca2+ FIGURE 1 Binding of fn-D50LA-TD to a model

membrane. (A) 1 mM D50LA-TD (gray) did not associate with the tBLM in 50 mM NaCl. Further-more, 0.1mM, 1 mM, or 2 mM fn-D50LA-TD in 50 mM NaCl (0.1 F, 1 F, and 2 F in blue) did not associate with the membrane (B) For a similar tBLM and buffer conditions plus 2 mM CaCl2,

1.5mM or 4 mM D50LA-TD (gray) did not asso-ciate with the membrane. However, 3.2 mM fn-D50LA-TD (3.2 F in red) did show association. (C) Stepwise addition to the solution phase of 0.01mM, 0.05 mM, 0.1 mM, 0.5 mM, 1 mM, and 2mM fn-D50LA-TD (red) showed increased mem-brane association with 50 mM NaCl and 2 mM CaCl2. (D) From this association we can produce

a binding curve to determine the KDof the protein

to the membrane. (tBLM composition: 30:70:þ3 DOPS/DOPC/Chol)

(5)

Fluctuations of the Ig-fold of

D50LA-TD

and preLA-TD in the presence of calcium

To obtain a deeper insight into the structural responses of the

LA TDs to Ca

, we performed molecular dynamics (MD)

simulations of the proteins in buffer. To deal with the high

degree of intrinsic disorder, we simulated multiple

low-energy, pseudoequilibrated structures with REMD (

25,33

).

Starting with amino acid chains with extreme initial

struc-tures (straight or fully coiled) except for the structured

Ig-fold region taken from the NMR structure in the protein

database (

31

), we allowed the protein to relax freely in an

implicit solvent environment. To enhance the sampling

speed as well as to ensure that structures reached relevant

local energy minima, we investigated numerous

indepen-dent structures with implicit-solvent REMD to find as

many low-energy states as possible (see Methods). Results

were categorized according to their structural similarity,

and the four most significant conformations were chosen

for further studies of the structural impact of ions (

7

). These

four conformations were further refined in REMD

simula-tions with explicit solvent under various ion condisimula-tions

(see Methods). To quantify the TD’s structural response to

ion exposure, we focused on four peripheral amino acids

(S429, S437, L478, and I497), which form orthogonal

ex-tremes of the two

b-sheets that form the Ig-fold (

Fig. 3

A).

From the positions of these four amino acids, we determined

an orientation angle between the two

b-sheets, q (

Fig. 3

A).

The Ig-fold is slightly more open (

q is systematically larger)

in the presence of Ca

(

Fig. 3

B).

We computationally included a farnesyl group to the

LA-TDs. For an initial conformation, we used the

pseudoe-quilibrium structure of the unfarnesylated LA-TDs and

embedded the added C-terminal farnesyl group within the

hydrophobic core of the Ig-fold (see Methods for force

field). We then allowed the new structure to come to a

new equilibrium with explicit solvent. The Ig-fold of both

D50LA-TD and preLA-TD were more open if farnesylated,

and exposure to Ca

opened the fold further up by a larger

angle than with the unfarnesylated proteins (

Fig. 3

C).

Because the average distance between I497 and L478

10−1 100 0 0.5 1 log [Ca2+], mM Δ I/I0 ( λ = 342 nm) 315 355 395 0.5 1.0 λ (nm) I/I0 (a.u.) 315 355 395 0.5 1.0 λ (nm) Δ50LA−TD preLA−TD preLA−TD R2 = 0.99 increasing [Ca2+] increasing [Ca2+] Δ50LA−TD

A

C

B

FIGURE 2 Calcium caused a conformational change toD50LA-TD and preLA-TD. The emis-sion spectra of tryptophan residues for purified (A)D50LA-TD in 50 mM NaCl and (B) preLA-TD in 500 mM NaCl excited at 295 nm were measured for increasing Ca2þconcentration. For Ca2þ concentration above 1 mM, there was no change in intensity, indicating no more substantial conformational change induced by Ca2þ. (C) From plots of emission intensity at 342 nm versus Ca2þ concentration, we determined Ca2þ dissociation constant (KD~250300 mM) and binding

stoichi-ometry (1:1). The DI/I0 (342 nm) is averaged

for 2–3 samples and in NaCl concentrations 50–700 mM NaCl for a given CaCl2concentration.

The Ca2þdissociation constant did not change with the NaCl concentration of the solution.

FIGURE 3 Altered Ig-fold structures of preLA-TD and D50LA-TD exposed to calcium. (A) We determined the dihedral angle between the twob-sheets of the Ig-fold, q, from the a-carbon of orthogonal amino acids (S429, S437, I497, and L478) within the full TD. (B) We tracked the fluc-tuations of the dihedral angleq for the preLA-TA conformation in the replica with lowest temperature after each exchange event in pseudoequili-brium. The angle was always larger in the presence of Ca2þ. (C) Statistical comparison of the dihedral angle of the Ig-fold for numerous pseudoequili-brated protein structures for preLA-TD, fn-preLA-TD,D50LA-TD, and fn-D50LA-TD showed increases in q with Ca2þ.

Biophysical Journal 104(10) 2246–2253

(6)

is

>1.7 nm, an increase in the orientation angle q indicates

that the space between the two

b-sheets accommodates

more water (with a van der Waals radius of 0.14 nm) in

the presence of the ion than in its absence.

Fluctuations of the C-terminus of

D50LA-TD

and preLA-TD with calcium

REMD simulations also provided insight into the

mecha-nism by which calcium triggers membrane binding of

farne-sylated lamin A tail domains. We observed that Ca

bound

spontaneously to the C-terminal CaaX motif, and the

inter-action persisted through the entire REMD simulation (cyan

Ca

to red CSIM in

Fig. 4

A,

Fig. S3

). Generally, Ca

binding increased the fluctuation and mobility of the

C-terminal region for both

D50LA-TD (beyond 610

0

; the

prime refers to the new amino acid number for the deletion)

and preLA-TD (beyond 660) by 30% and 39%, respectively

(

Fig. 4

B).

C-terminal motility was reduced when the protein was

farnesylated (

Fig. 4

B), and the Ca

-mediated motility

increase was enhanced for farnesylated proteins (

Fig. 4

B)

by 89% for fn-preLA-TD and 120% for fn-

D50LA-TD. In

the absence of Ca

, the farnesyl (

Fig. 5

, red) maintained

association with the hydrophobic core (

Fig. 5

A,

Ca

).

In the presence of Ca

, the farnesyl group escaped

the hydrophobic interior of the protein and was found at

the periphery of the pseudoequilibrated structures (

Fig.

5

A,

þCa

). This was quantified by the increase in solvent

available surface area of the farnesyl group in the presence

of Ca

(

Fig. 5

B) and increased end-to-end length of the

TDs (

Fig. 5

C). We suggest that the farnesyl group

associ-ates with the hydrophobic protein core, and the association

of Ca

to the CaaX causes increased probability of farnesyl

solvent exposure for membrane association.

DISCUSSION

Membrane binding of farnesylated lamin A tail

requires calcium

We find that fn-

D50LA-TD binds to purified membrane

bilayers with a K

D

~2

mM. This binding coefficient is

mark-edly weaker than most biological binding coefficients;

spe-cific protein-protein binding typically occurs on the order

of 10s to 100s of nM. We suggest that this weak binding

affinity allows protein-membrane association while not

causing irreversible association. In the case of lamin A,

we suggest that the farnesyl group weakly associates the

protein to the membrane, and transmembrane lamin-binding

proteins then more permanently connect lamin A to the

inner nuclear membrane. This weak binding is similar to

qualitative observations that farnesyl groups are insufficient

to stably associate proteins with membranes (

16–18

).

Dur-ing normal processDur-ing, lamin TDs, and other CaaX-domain

proteins, are farnesylated, the aaX residues are cleaved by

the ZMPSTE-1 protease, and the terminus is methylated

in the nucleus (

34

). The proteolysis and methylation steps

may play an additional role in stabilizing membrane

associ-ation by increasing hydrophobicity. However, there have

been conflicting reports regarding the additional stability

conferred by methylation (

35–37

).

The association of the fn-

D50LA-TD is dependent on

Ca

. There have been previous studies examining

quali-tative membrane association of fn-preLA (

38

), but we

note that the buffer used contained the divalent cation

magnesium. Similarly, recoverin has a Ca

-induced

A

B

FIGURE 4 Dynamics of the C-terminus of preLA-TD andD50LA-TD exposed to calcium. (A) We examined all pseudoequilibrated conformations of unfarnesylated preLA-TD produced in explicit-solvent REMD simula-tions and acquired snaphopys of the lowest energy conformasimula-tions (overlaid). In the presence of explicit water and sodium, the C-terminus CSIM (red) was located inside the preLA tail structure. With the addition of Ca2þ(cyan), the C-terminus pointed away from the protein. (B) The C-terminal amino acids for each of the four TDs for 10 ns of MD simulation showed altered fluctuations or root mean-square displacement from initial coordinates in the presence of Ca2þ. For each TD and ionic condition we repeated the simulation for the four lowest energy conformations and statis-tically compared the results. For the last 3 to ~6 residues, Ca2þcaused a substantial increase in amino acid fluctuation. The Ca2þ-induced fluctua-tion changes were more substantial for farnesylated proteins, which showed reduced C-terminal fluctuations in the absence of Ca2þ.

(7)

conformational switch to expose the myristoyl group, which

is responsible for membrane association (

39

). Unlike

recov-erin, lamin A tail domains are mostly disordered, therefore

the conformational switch is much less obvious. We

demon-strated through experiments and simulations that

D50LA-TD and pre-LA D50LA-TD undergo a conformational change

induced by Ca

that opens the Ig-fold and increases the

exposure and mobility of the farnesylated C-terminus.

The Ca

-induced conformational change may also

regu-late protein-membrane interactions, protein-DNA, or

pro-tein-protein interactions. The large size, intrinsic disorder,

and Ig-fold binding pocket of the tail domain may all allow

for promiscuous binding and may be at least partially

responsible for the ability of lamin A tail to bind different

partners (

1

). Here, we have shown Ca

-induced

conforma-tional changes, and protein binding to the lamin A of one of

the many lamin A-associated proteins may also induce

conformational changes in the tail domain. We find that,

similar to Ca

, Mg

can facilitate the membrane

associa-tion of fn-

D50LA-TD, and is the subject of future work.

Membrane association of farnesylated lamin A tail

domain includes binding and aggregation

Simulations of the

D50LA-TD structure show the protein

diameter to be 5.2

5 0.2 nm, and a complete monolayer

of protein coverage is reached at ~1

mM of solution protein

concentration. However, above 1

mM we still observe

addi-tional, stable membrane association. This additional

associ-ation is likely due to a combinassoci-ation of aggregassoci-ation coupled

with binding. Because high salt levels are required to

pre-vent aggregation of fn-preLA-TD, and proteins show no

membrane association as monomers, we suggest that the

protein-protein association of fn-preLA-TD overwhelms

the protein-membrane association with excess protein

(

40

). This highlights that fn-

D50LA-TD has an apparent

greater affinity for the membrane and enhanced membrane

association of

D50LA may contribute to the pathological

membrane association (

12,13

).

Lifetime of wild-type lamin A and pathological

D50LA

Most contemporary literature suggests that prelamin A

far-nesylation is responsible for the association of protein to the

inner nuclear membrane for further incorporation into the

lamina (

10,11

). It is unclear why lamins do not embed

them-selves into membrane structures outside the nucleus because

farnesyltransferases have been located in the cytoplasm near

the endoplasmic reticulum and inside the nucleus.

From the results presented here, we suggest changes in

the tail domain of preLA from the time it is produced

from LMNA mRNA until it is processed to form mature

lamin A. We speculate that after transcription, the

C-termi-nal CaaX domain is hidden inside the TD of preLA. The

preLA (probably as a dimer) is transported into the nucleus

via the TD nuclear localization sequence (NLS). In the

nucleus, the higher levels of charge from DNA and counter

ions induce a conformational change that exposes the

CaaX for farnesylation and other modifications leading to

FIGURE 5 Simulation of farnesyl dynamics of preLA andD50LA TD. We preequilibrated the far-nesyl group within the hydrophobic interior of the respective TDs and allowed equilibration in explicit solvent with 50 mM NaCl or 50 mM NaCl þ 2 mM CaCl2. (A) Comparison of the

four most significant pseudoequilibrated structures of fn-preLA-TD. With no Ca2þthe farnesyl group (red) was found within the molecule. In the pres-ence of Ca2þ, the farnesyl group escaped from the TD. (B) The solvent accessible surface area of the farnesyl group, given by the number of water molecules in contacting with the surface of farnesyl group (threshold distance of 0.38 nm), of fn-D50LA-TD and fn-preLA-TD increased in Ca2þ. (C) The end-to-end length of the entire TD of fn-D50LA-TD and fn-preLA-TD also increased, suggesting that the Ca2þ makes the C-terminus more extended.

Biophysical Journal 104(10) 2246–2253

(8)

insertion into the inner nuclear membrane. There are high

levels of divalent cations including calcium, magnesium,

and zinc inside the nucleus (

41

), but exact levels of free

divalent cations in the nucleus are difficult to determine

because free ions are buffered by surrounding chromatin

structure. It is also possible that small amounts of preLA

may be farnesylated by cytoplasmic farnesyltransferases.

We suggest that these farnesyl groups may incorporate

into the Ig-fold until they are exposed to high Ca

levels.

Ig-folds and similar

b-sheet structures have been shown to

act as solubilization agents for lipidated proteins (

42–44

).

In the case of fn-preLA, the farnesylation could associate

with its own Ig-fold or that of a neighbor.

SUPPORTING MATERIAL

Five figures and supplemental methods are available at http://www. biophysj.org/biophysj/supplemental/S0006-3495(13)00445-1.

We gratefully acknowledge the assistance by Matthew Biegler (CMU MSE) and Peter Yaron (CMU ChemE) and we are grateful to Katherine Wilson and Susan Michaelis (Johns Hopkins Department of Cell Biology) for care-ful reading of the manuscript.

Z.Q. and M.J.B. received support from National Science Foundation (NSF) (CMMI-0642545) and Office of Naval Research-Presidential Early Career Award for Scientists and Engineers (ONR-PECASE) (N00014-10-1-0562). K.N.D. and A.K. received support from the Progeria Research Foundation, NSF (CBET- 0954421 CAREER) and National Institutes of Health (NIH) National Institute on Aging (NIA) (NRSA F30-AG030905 to A.K.). M.L received support from NIH (1R01-GM101647).

REFERENCES

1. Simon, D. N., and K. L. Wilson. 2011. The nucleoskeleton as a genome-associated dynamic ‘network of networks’. Nat. Rev. Mol. Cell Biol. 12:695–708.

2. Dittmer, T. A., and T. Misteli. 2011. The lamin protein family. Genome Biol. 12:222–304.

3. Dahl, K. N., and A. Kalinowski. 2011. Nucleoskeleton mechanics at a glance. J. Cell Sci. 124:675–678.

4. Herrmann, H., H. Ba¨r,., U. Aebi. 2007. Intermediate filaments: from cell architecture to nanomechanics. Nat. Rev. Mol. Cell Biol. 8:562–573.

5. Barrowman, J., C. Hamblet,., S. Michaelis. 2008. Analysis of prela-min A biogenesis reveals the nucleus to be a CaaX processing compart-ment. Mol. Biol. Cell. 19:5398–5408.

6. Eriksson, M., W. T. Brown,., F. S. Collins. 2003. Recurrent de novo point mutations in lamin A cause Hutchinson-Gilford progeria syn-drome. Nature. 423:293–298.

7. Qin, Z., A. Kalinowski,., M. J. Buehler. 2011. Structure and stability of the lamin A tail domain and HGPS mutant. J. Struct. Biol. 175:425–433.

8. Scaffidi, P., and T. Misteli. 2006. Lamin A-dependent nuclear defects in human aging. Science. 312:1059–1063.

9. Davies, B. S., C. Coffinier,., L. G. Fong. 2011. Investigating the pur-pose of prelamin A processing. Nucleus. 2:4–9.

10. Hennekes, H., and E. A. Nigg. 1994. The role of isoprenylation in membrane attachment of nuclear lamins. A single point mutation pre-vents proteolytic cleavage of the lamin A precursor and confers mem-brane binding properties. J. Cell Sci. 107:1019–1029.

11. Lutz, R. J., M. A. Trujillo,., M. Sinensky. 1992. Nucleoplasmic local-ization of prelamin A: implications for prenylation-dependent lamin A assembly into the nuclear lamina. Proc. Natl. Acad. Sci. USA. 89:3000– 3004.

12. Dahl, K. N., P. Scaffidi,., T. Misteli. 2006. Distinct structural and mechanical properties of the nuclear lamina in Hutchinson-Gilford progeria syndrome. Proc. Natl. Acad. Sci. USA. 103:10271–10276. 13. Goldman, R. D., D. K. Shumaker,., F. S. Collins. 2004.

Accumula-tion of mutant lamin A causes progressive changes in nuclear architec-ture in Hutchinson-Gilford progeria syndrome. Proc. Natl. Acad. Sci. USA. 101:8963–8968.

14. Reddy, S., and L. Comai. 2012. Lamin A, farnesylation and aging. Exp. Cell Res. 318:1–7.

15. Kubben, N., M. Adriaens,., T. Misteli. 2012. Mapping of lamin A-and progerin-interacting genome regions. Chromosoma. 121:447–464. 16. Silvius, J. R., and F. l’Heureux. 1994. Fluorimetric evaluation of the affinities of isoprenylated peptides for lipid bilayers. Biochemistry. 33:3014–3022.

17. Nadolski, M. J., and M. E. Linder. 2007. Protein lipidation. FEBS J. 274:5202–5210.

18. Philips, M. R. 2012. Ras hitchhikes on PDE6d. Nat. Cell Biol. 14:128–129.

19. Taimen, P., K. Pfleghaar,., R. D. Goldman. 2009. A progeria muta-tion reveals funcmuta-tions for lamin A in nuclear assembly, architecture, and chromosome organization. Proc. Natl. Acad. Sci. USA. 106:20788–20793.

20. Vivian, J. T., and P. R. Callis. 2001. Mechanisms of tryptophan fluores-cence shifts in proteins. Biophys. J. 80:2093–2109.

21. Chen, Y., and M. D. Barkley. 1998. Toward understanding tryptophan fluorescence in proteins. Biochemistry. 37:9976–9982.

22. Shenoy, S., R. Moldovan,., M. Losche. 2010. In-plane homogeneity and lipid dynamics in tethered bilayer lipid membranes (tBLMs). Soft Matter. 2010:1263–1274.

23. Vockenroth, I. K., C. Ohm,., I. Ko¨per. 2008. Stable insulating teth-ered bilayer lipid membranes. Biointerphases. 3:FA68–FA73. 24. Valincius, G., D. J. McGillivray,., M. Lo¨sche. 2006. Enzyme activity

to augment the characterization of tethered bilayer membranes. J. Phys. Chem. B. 110:10213–10216.

25. Sugita, Y., and Y. Okamoto. 1999. Replica-exchange molecular dynamics method for protein folding. Chem. Phys. Lett. 314:141–151. 26. Brooks, B. R., R. E. Bruccoleri,., M. Karplus. 1983. Charmm: a pro-gram for macromolecular energy, minimization, and dynamics calcula-tions. J. Comput. Chem. 4:187–217.

27. Feig, M., J. Karanicolas, and C. L. Brooks, 3rd. 2004. MMTSB Tool Set: enhanced sampling and multiscale modeling methods for applica-tions in structural biology. J. Mol. Graph. Model. 22:377–395. 28. Vanommeslaeghe, K., E. Hatcher, ., A. D. Mackerell, Jr. 2010.

CHARMM general force field: a force field for drug-like molecules compatible with the CHARMM all-atom additive biological force fields. J. Comput. Chem. 31:671–690.

29. Shenoy, S., P. Shekhar,., M. Lo¨sche. 2012. Membrane association of the PTEN tumor suppressor: molecular details of the protein-mem-brane complex from SPR binding studies and neutron reflection. PLoS ONE. 7:e32591.

30. McGillivray, D. J., G. Valincius,., M. Lo¨sche. 2007. Molecular-scale structural and functional characterization of sparsely tethered bilayer lipid membranes. Biointerphases. 2:21–33.

31. Dhe-Paganon, S., E. D. Werner,., S. E. Shoelson. 2002. Structure of the globular tail of nuclear lamin. J. Biol. Chem. 277:17381–17384. 32. Krimm, I., C. Ostlund,., S. Zinn-Justin. 2002. The Ig-like structure of

the C-terminal domain of lamin A/C, mutated in muscular dystrophies, cardiomyopathy, and partial lipodystrophy. Structure. 10:811–823. 33. Swendsen, R. H., and J. S. Wang. 1986. Replica Monte Carlo

(9)

34. Young, S. G., L. G. Fong, and S. Michaelis. 2005. Prelamin A, Zmpste24, misshapen cell nuclei, and progeria—new evidence sug-gesting that protein farnesylation could be important for disease path-ogenesis. J. Lipid Res. 46:2531–2558.

35. Leung, K. F., R. Baron,., M. C. Seabra. 2007. Rab GTPases contain-ing a CAAX motif are processed post-geranylgeranylation by proteol-ysis and methylation. J. Biol. Chem. 282:1487–1497.

36. Parish, C. A., A. V. Smrcka, and R. R. Rando. 1996. The role of G pro-tein methylation in the function of a geranylgeranylated beta gamma isoform. Biochemistry. 35:7499–7505.

37. Hancock, J. F., K. Cadwallader,., C. J. Marshall. 1991. A CAAX or a CAAL motif and a second signal are sufficient for plasma membrane targeting of ras proteins. EMBO J. 10:4033–4039.

38. Polychronidou, M., A. Hellwig, and J. Grosshans. 2010. Farnesylated nuclear proteins Kugelkern and lamin Dm0 affect nuclear morphology by directly interacting with the nuclear membrane. Mol. Biol. Cell. 21:3409–3420.

39. Xu, X., R. Ishima, and J. B. Ames. 2011. Conformational dynamics of recoverin’s Ca2þ-myristoyl switch probed by 15N NMR relaxation dispersion and chemical shift analysis. Proteins. 79:1910–1922. 40. Marshall, C. J. 1993. Protein prenylation: a mediator of protein-protein

interactions. Science. 259:1865–1866.

41. Tvedt, K. E., G. Kopstad,., J. Halgunset. 1987. Subcellular concen-trations of calcium, zinc, and magnesium in benign nodular hyperplasia of the human prostate: x-ray microanalysis of freeze-dried cryosec-tions. Cancer Res. 47:323–328.

42. Rak, A., O. Pylypenko,., K. Alexandrov. 2003. Structure of Rab GDP-dissociation inhibitor in complex with prenylated YPT1 GTPase. Science. 302:646–650.

43. Tian, W., J. Yu,., X. Luo. 2010. Structural and functional analysis of the YAP-binding domain of human TEAD2. Proc. Natl. Acad. Sci. USA. 107:7293–7298.

44. Hoffman, G. R., N. Nassar, and R. A. Cerione. 2000. Structure of the Rho family GTP-binding protein Cdc42 in complex with the multifunc-tional regulator RhoGDI. Cell. 100:345–356.

Biophysical Journal 104(10) 2246–2253

Figure

FIGURE 3 Altered Ig-fold structures of preLA-TD and D50LA-TD exposed to calcium. (A) We determined the dihedral angle between the two b-sheets of the Ig-fold, q, from the a-carbon of orthogonal amino acids (S429, S437, I497, and L478) within the full TD
FIGURE 4 Dynamics of the C-terminus of preLA-TD and D50LA-TD exposed to calcium. (A) We examined all pseudoequilibrated conformations of unfarnesylated preLA-TD produced in explicit-solvent REMD  simula-tions and acquired snaphopys of the lowest energy con
FIGURE 5 Simulation of farnesyl dynamics of preLA and D50LA TD. We preequilibrated the  far-nesyl group within the hydrophobic interior of the respective TDs and allowed equilibration in explicit solvent with 50 mM NaCl or 50 mM NaCl þ 2 mM CaCl 2

Références

Documents relatifs

The shapes obtained following the combination of the genetic algorithm reported in [20] and the topological optimization by the mixed relaxation/level set method are presented

da Fonseca, Brett Drury and André Melo, Conformational characterization of disulfide bonds: A tool for protein classification, Journal of Theoretical

Voici un tableau récapitulatif des patients admis au service de médecine nucléaire et suivis pour l’étude de l’apport repérage isotopique peropératoire

This tBLM model was assembled on various supports (e.g., glass and gold surface) modified with an amine layer for lipid bilayer reconstitution using vesicles con- taining

• GIS Sol: Soil inventory and monitoring programmes coordinated by Inra (InfoSol unit). RMQS

The Quantitative Methods for Psychology journal opens a new category of articles, the Outstanding data sets, whose purpose is to publish research paper as well as research data..

We shall start with a short description of the three main numerical methods in shape optimization, namely the boundary variation, the homogenization and the level set methods.. For

Using cell biology approaches, we show that recProt01230 is able to adhere to bovine host cells and interacts with proteins from the cell lysate and the &#34;membranes/