Protein- and Metal-dependent Interactions of a Prominent Protein in Mussel Adhesive Plaques

18
Protein- and Metal-dependent Interactions of a Prominent Protein in Mussel Adhesive Plaques * S Received for publication, April 12, 2010, and in revised form, June 18, 2010 Published, JBC Papers in Press, June 21, 2010, DOI 10.1074/jbc.M110.133157 Dong Soo Hwang ‡1,2 , Hongbo Zeng §1 , Admir Masic , Matthew J. Harrington , Jacob N. Israelachvili , and J. Herbert Waite** 3 From the Materials Research Laboratory, University of California, Santa Barbara, California 93106, the § Department of Chemical and Materials Engineering, University of Alberta, Edmonton, Alberta T6G 2V4, Canada, the Department of Biomaterials, Max Planck Institute for Colloids and Interfaces, 14424 Potsdam-Golm, Germany, and the Department of Chemical Engineering and **Marine Science Institute, University of California, Santa Barbara, California 93106 The adhesive plaques of Mytilus byssus are investigated increasingly to determine the molecular requirements for wet adhesion. Mfp-2 is the most abundant protein in the plaques, but little is known about its function. Analysis of Mfp-2 films using the surface forces apparatus detected no interaction between films or between a film and bare mica; however, addi- tion of Ca 2 and Fe 3 induced significant reversible bridging (work of adhesion W ad ≈ 0.3 mJ/m 2 to 2.2 mJ/m 2 ) between two films at 0.35 M salinity. The strongest observed Fe 3 -mediated bridging approaches the adhesion of oriented avidin-biotin complexes. Raman microscopy of plaque sections supports the co-localization of Mfp-2 and iron, which interact by forming bis- or tris-DOPA-iron complexes. Mfp-2 adhered strongly to Mfp-5, a DOPA-rich interfacial adhesive protein, but not to another interfacial protein, Mfp-3, which may in fact displace Mfp-2 from mica. In the presence of metal ions or Mfp-5, Mfp-2 adhesion was fully reversible. These results suggest that plaque cohesiveness depends on Mfp-2 complexation of metal ions, particularly Fe 3 and also by Mfp-2 interaction with Mfp-5 at the plaque-substratum interface. The mussel holdfast or byssus is emerging as an effective model system for studying the requirements for opportunistic underwater adhesion (1). The 3,4-dihydroxyphenylalanine (DOPA) 4 -rich mussel foot proteins (MFPs) of the adhesive footprint, namely Mfp-3 and Mfp-5, have been subjected to particular scrutiny (2–3) given the recent demonstration that DOPA is capable of mediating reversible adhesion to titania surfaces at forces of nearly 1 nanonewton/DOPA (4). Mussel- inspired catecholic polymers are engineered increasingly for moisture-resistant adhesive and coating applications (5– 8). At least five other proteins known mostly as MFPs are pres- ent in the byssal plaque and contribute presumably to its adhe- sive performance: Mfp-1, Mfp-2, Mfp-4, Mfp-6, the prepep- sinized collagens (9), and thread matrix protein (Fig. 1) (10). Of these, only two have reasonably well established functions: Mfp-1 complexed to Fe 3 provides a protective outer coating for the plaque and thread (11–12), and the prepepsinized col- lagens are fiber-forming collagens that mediate the fusion of thread and plaque (13). Other proteins have been partially char- acterized and sequenced, but their role in byssal structure is largely speculative. Mfp-2 is particularly intriguing because it is the most abundant protein of byssal plaques comprising 25% of the plaque by weight (14). Mfp-2 has a mass of 45 kDa and consists of 11 tandem repeats of an EGF motif, each of which resembles a knot-like structure stabilized by three disulfide bonds (Fig. 2) (15–16). DOPA content is comparatively low at 3–5 mol % with an average of two residues per EGF repeat and three to four residues at the N- and C-terminal ends of Mfp-2. Surface binding studies of Mfp-2 using attenuated total reflec- tance Fourier transform infrared spectrometry showed the pro- tein to be adsorbed rapidly to a variety of surfaces, but it was displaced effectively by Mfp-1 (17). In vitro, in the absence of protein cross-linkers such as tyrosinase, Mfp-2 exhibits little tendency for interaction with Mfp-1 (18). To better understand the role of Mfp-2 in the byssal adhesive plaque, we tested its adhesion alone and in the presence of added cations and other MFPs using the surface forces appara- tus (SFA). SFA is suited ideally to distinguish the proteins that coat surfaces from those that bridge from one surface to another, an important distinction in adhesion (9). Our studies show that Mfp-2 by itself coats mica but tends not to bridge between two mica surfaces. However, addition of Ca 2 and Fe 3 can mediate bridge formation between films of Mfp-2, and Mfp-2 can bind directly to Mfp-5 films. The strongest bridging between two adsorbed Mfp-2 films was induced by addition of low micromolar Fe 3 . High resolution resonance Raman microprobe analysis of plaques supports the co-localization of iron and Mfp-2 and suggests that iron binding may be necessary for cohesion of Mfp-2 in plaque. MATERIALS AND METHODS MFP Purification from Mussel Feet—Blue mussel (Mytilus edulis L.) feet were obtained in 500-g flash-frozen lots from Northeast Transport of Union, Maine. Mussel foot proteins * This work was supported in part by National Institutes of Health Grant R01 DE018468 and National Science Foundation Materials Research Science and Engineering Center Grant DMR05-20415. S The on-line version of this article (available at http://www.jbc.org) contains supplemental “Materials and Methods,” Schemes S1–S3, Figs. S1–S3, and additional references. 1 Both authors contributed equally to this work. 2 Recipient of an Otis Williams Fellowship from the Santa Barbara Foundation. 3 To whom correspondence should be addressed: Marine Science Institute, University of California at Santa Barbara, Santa Barbara, CA 93106. Tel.: 805-893-2817; Fax: 805-893-7998; E-mail: [email protected]. 4 The abbreviations used are: DOPA, 3, 4-dihydroxyphenylalanine; MFP, mussel foot protein; SFA, surface forces apparatus; bis-tris, 2-[bis(2- hydroxyethyl)amino]-2-(hydroxymethyl)propane-1,3-diol. THE JOURNAL OF BIOLOGICAL CHEMISTRY VOL. 285, NO. 33, pp. 25850 –25858, August 13, 2010 © 2010 by The American Society for Biochemistry and Molecular Biology, Inc. Printed in the U.S.A. 25850 JOURNAL OF BIOLOGICAL CHEMISTRY VOLUME 285 • NUMBER 33 • AUGUST 13, 2010 by guest on April 4, 2016 http://www.jbc.org/ Downloaded from by guest on April 4, 2016 http://www.jbc.org/ Downloaded from by guest on April 4, 2016 http://www.jbc.org/ Downloaded from by guest on April 4, 2016 http://www.jbc.org/ Downloaded from by guest on April 4, 2016 http://www.jbc.org/ Downloaded from by guest on April 4, 2016 http://www.jbc.org/ Downloaded from by guest on April 4, 2016 http://www.jbc.org/ Downloaded from by guest on April 4, 2016 http://www.jbc.org/ Downloaded from by guest on April 4, 2016 http://www.jbc.org/ Downloaded from by guest on April 4, 2016 http://www.jbc.org/ Downloaded from

Transcript of Protein- and Metal-dependent Interactions of a Prominent Protein in Mussel Adhesive Plaques

Protein- and Metal-dependent Interactions of a ProminentProtein in Mussel Adhesive Plaques*□S

Received for publication, April 12, 2010, and in revised form, June 18, 2010 Published, JBC Papers in Press, June 21, 2010, DOI 10.1074/jbc.M110.133157

Dong Soo Hwang‡1,2, Hongbo Zeng§1, Admir Masic¶, Matthew J. Harrington¶, Jacob N. Israelachvili�,and J. Herbert Waite**3

From the ‡Materials Research Laboratory, University of California, Santa Barbara, California 93106, the §Department of Chemicaland Materials Engineering, University of Alberta, Edmonton, Alberta T6G 2V4, Canada, the ¶Department of Biomaterials, MaxPlanck Institute for Colloids and Interfaces, 14424 Potsdam-Golm, Germany, and the �Department of Chemical Engineering and**Marine Science Institute, University of California, Santa Barbara, California 93106

The adhesive plaques of Mytilus byssus are investigatedincreasingly to determine the molecular requirements for wetadhesion. Mfp-2 is the most abundant protein in the plaques,but little is known about its function. Analysis of Mfp-2 filmsusing the surface forces apparatus detected no interactionbetween films or between a film and bare mica; however, addi-tion of Ca2� and Fe3� induced significant reversible bridging(work of adhesion Wad ≈ 0.3 mJ/m2 to 2.2 mJ/m2) between twofilms at 0.35 M salinity. The strongest observed Fe3�-mediatedbridging approaches the adhesion of oriented avidin-biotincomplexes. Raman microscopy of plaque sections supports theco-localization of Mfp-2 and iron, which interact by formingbis- or tris-DOPA-iron complexes. Mfp-2 adhered strongly toMfp-5, a DOPA-rich interfacial adhesive protein, but not toanother interfacial protein, Mfp-3, which may in fact displaceMfp-2 frommica. In the presence ofmetal ions orMfp-5,Mfp-2adhesion was fully reversible. These results suggest that plaquecohesiveness depends on Mfp-2 complexation of metal ions,particularly Fe3� and also by Mfp-2 interaction with Mfp-5 atthe plaque-substratum interface.

The mussel holdfast or byssus is emerging as an effectivemodel system for studying the requirements for opportunisticunderwater adhesion (1). The 3,4-dihydroxyphenylalanine(DOPA)4-rich mussel foot proteins (MFPs) of the adhesivefootprint, namely Mfp-3 and Mfp-5, have been subjected toparticular scrutiny (2–3) given the recent demonstration thatDOPA is capable of mediating reversible adhesion to titaniasurfaces at forces of nearly 1 nanonewton/DOPA (4). Mussel-inspired catecholic polymers are engineered increasingly formoisture-resistant adhesive and coating applications (5–8).

At least five other proteins known mostly as MFPs are pres-ent in the byssal plaque and contribute presumably to its adhe-sive performance: Mfp-1, Mfp-2, Mfp-4, Mfp-6, the prepep-sinized collagens (9), and threadmatrix protein (Fig. 1) (10). Ofthese, only two have reasonably well established functions:Mfp-1 complexed to Fe3� provides a protective outer coatingfor the plaque and thread (11–12), and the prepepsinized col-lagens are fiber-forming collagens that mediate the fusion ofthread andplaque (13).Other proteins have been partially char-acterized and sequenced, but their role in byssal structure islargely speculative.Mfp-2 is particularly intriguing because it isthe most abundant protein of byssal plaques comprising �25%of the plaque by weight (14). Mfp-2 has a mass of 45 kDa andconsists of 11 tandem repeats of an EGF motif, each of whichresembles a knot-like structure stabilized by three disulfidebonds (Fig. 2) (15–16). DOPA content is comparatively low at3–5 mol % with an average of two residues per EGF repeat andthree to four residues at the N- and C-terminal ends of Mfp-2.Surface binding studies of Mfp-2 using attenuated total reflec-tance Fourier transform infrared spectrometry showed the pro-tein to be adsorbed rapidly to a variety of surfaces, but it wasdisplaced effectively by Mfp-1 (17). In vitro, in the absence ofprotein cross-linkers such as tyrosinase, Mfp-2 exhibits littletendency for interaction with Mfp-1 (18).To better understand the role ofMfp-2 in the byssal adhesive

plaque, we tested its adhesion alone and in the presence ofadded cations and other MFPs using the surface forces appara-tus (SFA). SFA is suited ideally to distinguish the proteins thatcoat surfaces from those that bridge from one surface toanother, an important distinction in adhesion (9). Our studiesshow that Mfp-2 by itself coats mica but tends not to bridgebetween two mica surfaces. However, addition of Ca2� andFe3� canmediate bridge formation between films ofMfp-2, andMfp-2 can bind directly to Mfp-5 films. The strongest bridgingbetween two adsorbed Mfp-2 films was induced by addition oflow micromolar Fe3�. High resolution resonance Ramanmicroprobe analysis of plaques supports the co-localization ofiron andMfp-2 and suggests that iron bindingmay be necessaryfor cohesion of Mfp-2 in plaque.

MATERIALS AND METHODS

MFP Purification from Mussel Feet—Blue mussel (Mytilusedulis L.) feet were obtained in 500-g flash-frozen lots fromNortheast Transport of Union, Maine. Mussel foot proteins

* This work was supported in part by National Institutes of Health Grant R01DE018468 and National Science Foundation Materials Research Scienceand Engineering Center Grant DMR05-20415.

□S The on-line version of this article (available at http://www.jbc.org) containssupplemental “Materials and Methods,” Schemes S1–S3, Figs. S1–S3,and additional references.

1 Both authors contributed equally to this work.2 Recipient of an Otis Williams Fellowship from the Santa Barbara Foundation.3 To whom correspondence should be addressed: Marine Science Institute,

University of California at Santa Barbara, Santa Barbara, CA 93106. Tel.:805-893-2817; Fax: 805-893-7998; E-mail: [email protected].

4 The abbreviations used are: DOPA, 3, 4-dihydroxyphenylalanine; MFP,mussel foot protein; SFA, surface forces apparatus; bis-tris, 2-[bis(2-hydroxyethyl)amino]-2-(hydroxymethyl)propane-1,3-diol.

THE JOURNAL OF BIOLOGICAL CHEMISTRY VOL. 285, NO. 33, pp. 25850 –25858, August 13, 2010© 2010 by The American Society for Biochemistry and Molecular Biology, Inc. Printed in the U.S.A.

25850 JOURNAL OF BIOLOGICAL CHEMISTRY VOLUME 285 • NUMBER 33 • AUGUST 13, 2010

by guest on April 4, 2016

http://ww

w.jbc.org/

Dow

nloaded from

by guest on April 4, 2016

http://ww

w.jbc.org/

Dow

nloaded from

by guest on April 4, 2016

http://ww

w.jbc.org/

Dow

nloaded from

by guest on April 4, 2016

http://ww

w.jbc.org/

Dow

nloaded from

by guest on April 4, 2016

http://ww

w.jbc.org/

Dow

nloaded from

by guest on April 4, 2016

http://ww

w.jbc.org/

Dow

nloaded from

by guest on April 4, 2016

http://ww

w.jbc.org/

Dow

nloaded from

by guest on April 4, 2016

http://ww

w.jbc.org/

Dow

nloaded from

by guest on April 4, 2016

http://ww

w.jbc.org/

Dow

nloaded from

by guest on April 4, 2016

http://ww

w.jbc.org/

Dow

nloaded from

Mfp-1, Mfp-2, Mfp-3, and Mfp-5 were purified from frozenM. edulis feet according to published procedures (19–21) andare summarized in the supplemental Scheme S1. Sample puritywas assessed by acid urea-PAGE, amino acid analysis, andMALDI time-of-flightmass spectrometry. Themol %DOPA inpurifiedMfp-1,Mfp-2,Mfp-3, andMfp-5 was�12,�2, 22, and28%, respectively, determined by amino acid analysis after a 1-hhydrolysis in 6 N HCl at 158 °C. Purified samples were freeze-dried, resuspended in 50mM acetic acid, and thereafter dividedinto convenient aliquot volumes for storage in vials at �70 °Cprior to testing. Low pH and protection from light were neces-sary to reduce DOPA losses during handling and storage.Milli-Q water (Millipore, Bedford, MA) was used for all glass-ware cleaning and solution preparation.Force Versus Distance Profiles Measurement by the SFA—

The SFA technique has been used for many years to measureboth normal and lateral forces between surfaces in vaporsand liquids, e.g. van der Waals forces, electrostatic forces,adhesion forces, friction and lubrication forces, hydro-phobic interactions, and specific and nonspecific biologicalinteractions (22–25). A diagram of the instrument anddescription of the underlying theory can be found in thesupplemental “Materials and Methods.” SFA can measureaccurately the normal (attractive adhesion or repulsive) forcesF as low as 10 nanonewtons as a function of surface separationdistanceDwith a resolution of�1Ådetermined usingmultiplebeam interferometry.The normal force-distance profiles and adhesion forces (Fad)

of Mfp-2 were determined using an SFA in a configurationreported previously (9). Briefly, a thinmica sheet of 1–5�mwasglued onto a cylindrical silica disk (radius, r� 2 cm). 100 �l of astockMfp-2 solution (20�g/ml) diluted in 0.1 M sodiumacetatewith 0.25 M potassium nitrate at pH 5.5, was injected onto onemica surface. (Note that in SFA, potassium nitrate is used inplace of NaCl to reduce chloride ion-induced corrosion of thesemireflecting silver layers under the mica substrates. Salt con-centration was adjusted to half that of salinity of seawater.)For experiments with iron, 1 mM bis-tris was added to stabi-

lize the solubility of Fe3� (26). Although the test pH is signifi-cantly lower than seawater (pH 8.2), it was necessitated bythe poor solubility of iron and themussel proteins at higher pHin the SFA. The two curved and coatedmica surfaces were thenmounted in the SFA chamber in a crossed-cylinder geometry,which corresponds roughly to a sphere of radius R on a flatsurface based on the Derjaguin approximation: F(D) �2�RW(D), where F(D) is the force between the two curved sur-faces, andW(D) is the interaction energy per unit area betweentwo flat surfaces. Themeasured adhesion or “pull-off” force Fadis related to the adhesion energy per unit area Wad by Fad �2�RWad for rigid (undeformable) surfaceswithweakly adhesiveinteractions, and by Fad� 1.5�RWad (used in this study) for softdeformable surfaces with strong adhesive contact (27, 28). In atypical experiment, the separation distance D is monitored insitu using the fringes of equal chromatic order inmultiple beaminterferometry. All experiments were performed at room tem-perature (23 °C).Protein film deposition to one or both mica surfaces was

determined by whether asymmetric or symmetric testing was

FIGURE 1. Byssal plaque proteins of Mytilus. A mussel (M. galloprovincialis,inset) is shown attached to a sheet of mica. One of its plaques (red circle) isenlarged as a schematic drawing to illustrate the approximate distribution ofknown proteins.

FIGURE 2. Mfp-2. A, sequence is based on Mfp-2 from M. galloprovincialis(UniProtKB/Swiss-Prot entry Q25464) (13, 14). The 11 EGF repeats arealigned according to the invariant six cysteine residues/EGF. Tyrosine res-idues known to be occasionally or always modified to DOPA are denotedin bold. The boxed sequence in EGF repeat 7 is a calcium binding motif.Mfp-2 from M. edulis is 90% identical and is listed as (UniProtKB/Swiss-Protentries Q1XBT6 –Q1XBT8) but was never submitted as a published report.B, structure of one EGF domain based on Ref. 15, showing the cysteineresidues paired for disulfides, charges, and location of the DOPA residues(*). Below the structure is a complete string of 11 EGF repeats showing theDOPA and acidic clusters.

Major Plaque Protein Interactions

AUGUST 13, 2010 • VOLUME 285 • NUMBER 33 JOURNAL OF BIOLOGICAL CHEMISTRY 25851

by guest on April 4, 2016

http://ww

w.jbc.org/

Dow

nloaded from

planned. In the asymmetric mode, protein was applied to onemica surface only. In contrast, protein was applied to bothmicasurfaces in symmetric mode.Raman Spectroscopic Studies—Plaques from Mytilus gallo-

provincialis were embedded in PEG-2000 (Carl Roth Gmbh),and 20-�m thick longitudinal sections were microtomed.Plaque sections were washed thoroughly with several changesin distilled water to remove any remaining PEG, positioned ona quartz slide in distilled water and fixed under a quartz cover-slip. For Raman microspectroscopy, a continuous laser beamwas focused on the sample through a confocal Raman micro-scope (model CRM200,WITec, Ulm, Germany) equipped witha piezo scanner (model P-500, Physik Instrumente, Karlsruhe,Germany). The diode-pumped, 785-nm near-infrared laserexcitation (Toptica Photonics AG, Graefelfing, Germany) wasused in combination with a 100� oil immersed (Nikon, NA �

1.25) microscope objective. Laserpower ranging between 15 and 30milliwatts was used for all measure-ments. The spectra were acquiredusing an air-cooled charge-coupled device (DU401A-DR-DD,Andor, Belfast, North Ireland)behind a grating (300 g mm�1)spectrograph (Acton, PrincetonInstruments Inc., Trenton, NJ)with a 6 cm�1 spectral resolution.Software ScanCtrlSpectroscopyPlus(version 1.38, Witec) was used formeasurement setup. Raman spectrawere processed and analyzed withWitec Project software (version 2.02).Raman spectra were background-subtracted and lightly smoothedusing the first order polynomial func-tion and 9-point Savitzky-Golay filter(4th order polynomial), respectively.More details on background subtrac-tion are available in the supple-mental “Materials andMethods.”Color images were obtained from

Witec Project software using thebasis analysis and image color com-bination functions. In the case ofbasis analysis, the algorithm fitseach spectrum of the multigraphdata object with a linear combina-tion of the basis spectra (tree spec-tra, see Fig. 9E) using the leastsquares method. To solve theproblem of differing fluorescencebackground in various parts of thesample, the first derivative of bothmultigraph data object and ba-sis spectra was performed. Theweighting factors of various com-ponents obtained by fitting werestored in an image and combined

in a false color bitmap using the image color combinationfunction.A purified solution of Mfp-2 (1 mg/ml) in 0.1 M sodium ace-

tate with 0.25 M KNO3 and 1 mM bis-tris (pH 5.5) was mixedwith a small volume of 1 mM FeCl3 in 10 mM bis-tris (pH 5.5)and equilibrated for 10 min. The added iron was adjusted toachieve DOPA:Fe3� ratios of 2:1 and 20:1. The pH was thenraised to �8.0 with 0.1 M NaOH, and a droplet of the solutionwas dried by evaporation on a glass slide. Raman spectra weretaken of the protein-iron film at the edge of the film and ofMfp-2 solutions prior to iron complexation.

RESULTS

Self-interaction of Mfp-2—As the dominant plaque proteinbetween the byssal thread and adhesive interface (14), Mfp-2needs to be able to interact strongly with interfacial proteins at

FIGURE 3. Interaction between Mfp-2 films adsorbed to mica relative to their separation distance (D).A, (asymmetric) Mfp-2 adsorbed to one mica surface only; B, (symmetric) Mfp-2 adsorbed to both mica surfaces.Black, approach; gray, separation. The y axis on the left gives the measured force, F/R (normalized by the radiusof the surface), whereas the y axis on the right gives the corresponding adhesion energy per unit area (W)between two flat surfaces, defined by W � F/1.5�R (27, 28). The hard wall corresponds to the thickness of theprotein films and is indicated by an arrow (top left). Separation was after a brief (�1–2 min) contact; interactionswere unchanged by longer contacts.

FIGURE 4. Influence of added Ca2� (5 �M) on interaction between two symmetric Mfp-2 films adsorbed tomica. Contact times in A, B, and C were as shown. In D, another approach, brief contact, and separation wasperformed after C. Black, approach; gray, separation. The hard wall corresponds to the thickness of protein filmsand is denoted by an arrow (top left).

Major Plaque Protein Interactions

25852 JOURNAL OF BIOLOGICAL CHEMISTRY VOLUME 285 • NUMBER 33 • AUGUST 13, 2010

by guest on April 4, 2016

http://ww

w.jbc.org/

Dow

nloaded from

the substratum, with itself, andwith various proteins within themussel adhesive plaque (Fig. 1). We tested the adhesion ofMfp-2 to bare mica (asymmetric) and of two opposing Mfp-2films on mica (symmetric) using the surface forces apparatus.The ability of Mfp-2 to coat or bridge on mica surfaces isrevealed in the force-distance (F versus D) profile and by therepulsion associated with an initial approach to the “hard wall”followed by separation of the surfaces (note that the hardwall isdefined as the mica-mica separation at which the thickness of

the confined proteins becomesasymptotic with increased normalload or pressure). Adsorption ofMfp-2 tomica was confirmed by thehard wall distance shift from 5 to 10nmevident from the fringes of equalchromatic order signal shift andshape changes as shown previously(9). No apparent adhesion was mea-sured between Mfp-2 and baremica, nor could adhesion be de-tected between two Mfp-2 films(Fig. 3, A and B). Even a 1-h contacttime failed to induce adhesion inMfp-2.Interaction of Mfp-2 with Ca2�

and Fe3�—As significant levels ofcalcium and iron have been detectedin byssal plaques (�0.1% (w/w) (29)),it seemed appropriate to test theeffect of these ions on Mfp-2 adhe-sion. Weak adhesion between twosymmetric Mfp-2 coated surfacesoccurred after adding 5�MCa2� (Fig.4A). Adhesion improved slightly to�0.28 mJ/m2 by increasing the con-tact time for Mfp-2 films (Fig. 4B).Ca2�-mediated bridging of Mfp-2films was further subjected to fivecycles of approach and separationwithout significant loss in adhesionenergy (Fig. 4D).With respect to Fe3�, no adhe-

sion was initially detected afterintroducing 5 �M Fe3� betweensymmetricMfp-2 films, but a longercontact time (�1 h) resulted insignificant adhesion (�1.3 mJ/m2)(Fig. 5B). Reapproach and separa-tion after breaking the 1-h contactproduced the strongest observedadhesion (�2.2 mJ/m2) in ourstudies (Fig. 5C). A reversibleiron-mediated bridging of Mfp-2is suggested by the instantaneousenhanced adhesion of subsequentapproach-separationcycles.Study-ing the effect of pH on the iron-Mfp-2 interaction proved chal-

lenging because both iron and protein tend to precipitate athigher pH; however, it was possible to raise the pH to 6.8(Fig. 6, A–D). Remarkably, adhesion remained the same, butthe rate at which adhesion increased was faster (compare, forexample, the similar asymptotic adhesion after 1 h at pH 5.5and 6.8 in Figs. 5B and 6C, respectively). However, after 10min at pH 6.8, adhesion was 70% complete (Fig. 6B), whereasat pH 5.5, adhesion was not yet detectable. Higher pH alsoaffects the magnitude of adhesion after the 1-h separa-

FIGURE 5. Influence of added Fe3� (5 �M) on the interaction between two symmetric Mfp-2 filmsadsorbed to mica. Contact times in A–D were as indicated. Black, approach; gray, separation. The hard wallcorresponds to the thickness of protein films and is denoted by an arrow (top left).

FIGURE 6. Influence of Fe3� (5 �M) on the interaction between two symmetric Mfp-2 films on mica at pH6.8. A, brief contact following iron addition; B, 10-min contact; C, 1-h contact; D, brief contact and separationafter C; inset, zoom of approach highlights the persistent jump-in between 30 to 23 nm. Black, approach; gray,separation.

Major Plaque Protein Interactions

AUGUST 13, 2010 • VOLUME 285 • NUMBER 33 JOURNAL OF BIOLOGICAL CHEMISTRY 25853

by guest on April 4, 2016

http://ww

w.jbc.org/

Dow

nloaded from

tion and approach. At pH 6.8, it is only half that at 5.5(Fig. 6D).Interaction between Mfp-2 and Other MFPs—Mfp-3 and

Mfp-5 are present in the plaque footprints presumably as sur-face primers and exhibit moderate adhesion to mica as well asother surface chemistries (9).5 We hypothesized that Mfp-2should be able to adhere to either or both Mfp-3 and -5. In anasymmetric experiment,Mfp-2 was deposited on onemica faceand a film of either Mfp-3 or Mfp-5 was formed on the other.Mfp-2 showed strong adhesion toMfp-5 that took effect imme-diately and reached amaximumof 1.3mJ/m2 after a 1-h contact(Fig. 7C). There was no change in the hard wall, suggesting thatboth films remain on the mica.

In contrast, Mfp-2 showed no adhesion to Mfp-3 until after10 min of contact, after which it increased to 1.0 mJ/m2 at 60min. The observed hardwall at 10 and 60min is only half that at0 min, suggesting that the proteins are rearranging on the sur-face or that Mfp-2 is displaced by Mfp-3 (Fig. 8, A–C). Toexplore whether iron addition enhances adhesion betweenMfp-2 andMfp-3 orMfp-5, we repeated the asymmetric exper-iments in the presence of these metal ions (5 �M), but no sig-nificant effects were detected (supplemental Figs. S2 and S3).Consistent with a previous report, Mfp-1 did not exhibit

bridging in our studies. Similarly, whenMfp-1 andMfp-2 wereasymmetrically coated on mica, a hard wall distance of 10 nmsuggests the persistence of both films, but there was no adhe-5 H. Zeng, D. S. Hwang, and J. H. Waite, unpublished data.

FIGURE 7. Interaction between Mfp-5 and Mfp-2 films adsorbed to mica.Contact times were as shown in A–C. Black, approach; gray, separation. Thehard wall corresponds to the thickness of protein films and is denoted by anarrow (top left). FIGURE 8. Interaction between Mfp-3 and Mfp-2 films adsorbed to mica.

Contact times were as shown in A–C. Black, approach; gray, separation. Thehard wall corresponds to the thickness of protein films and is denoted byarrows (top left).

Major Plaque Protein Interactions

25854 JOURNAL OF BIOLOGICAL CHEMISTRY VOLUME 285 • NUMBER 33 • AUGUST 13, 2010

by guest on April 4, 2016

http://ww

w.jbc.org/

Dow

nloaded from

sion between them (Fig. 9). Addition of 5 �M Fe3� did not leadto detectable bridging.Raman Spectroscopy onMussel Byssal Plaque and Purified fp-2—

DOPA-iron coordination gives distinctive resonance Ramanspectra that can be combined with confocal microscopy tolocalize DOPA-iron complexes in biomaterials (11). Different

regions of the spectrum, namely CH stretching, Fe3�-DOPAcomplexation, and phenylalanine, were integrated after Ramanimaging of a thin (20-�m) section of the byssal plaque (Fig. 10,A–D). Notably, DOPA-iron coordination (500–650 cm�1) isdistributed throughout the plaque with highest intensity in thecuticle (Fig. 10C). Phenylalanine (1003 cm�1) is only prominentin the foam-like core (Fig. 10D). Average spectrawere extractedfrom the differentmorphological regions of the plaque (cuticle,core, and interface) (Fig. 10E) and were fitted to the image (Fig.10F). As expected, the plaque cuticle spectrum is essentiallyidentical to the thread cuticle spectrum published previously(10). Spectra of the plaque core, however, deviate subtly fromthe cuticle in some respects: a) a sharp peak at 1003 cm�1 asso-ciated with Phe (30); b) several shifts and intensity changes inthe region of DOPA-metal complex vibrations (500–650cm�1); c) changes in intensity of tyrosine-related peaks in theregion of 810–860 cm�1; and d) changes in the DOPA ringvibrations (1380–1520 cm�1). The Raman spectrum of theplaque-surface interface shows further differences from thecore. The adhesive protein primers (Mfp-3 and Mfp-5) arelocated this region, which exhibits the lowest intensities forDOPA-iron complexes (Fig 10E, green trace).Mixtures of Fe3� and purified Mfp-2 precipitated at pH � 8

also showed a strong resonance Raman signal indicative ofDOPA-iron coordination (Fig. 11with the triscatecholato-iron struc-ture). The Phe peak at 1003 cm�1 iseasy to spot in the uncomplexedprotein but is overwhelmed by reso-nance at 2:1 DOPA to iron. At 20:1DOPA to iron, in contrast, with lessdeveloped resonance Raman, thePhe peak ofMfp-2 is clearly evident.Indeed, apart from the buffer arti-fact at 1060 cm�1, the 20:1DOPA toiron spectrum of Mfp-2 is very sim-ilar to the Raman spectrum of theplaque. At 2:1 and 3:1 DOPA to ironratios, the spectra of Mfp-2 closelyresemble those of Mfp-1 followingthe same precipitation procedure(supplemental Fig. S4), but bothdiffer significantly from spectraobtained from the plaque in situ.

DISCUSSION

There has been a long-standingchallenge of how to best study pro-tein-protein interactions in sclero-tized biomaterials such as musselbyssal plaques. The protein precur-sors, although often quite abundantwhile stockpiled in secretory cells,become chemically modified fol-lowing release by covalent andmetal complexation-based cross-linking, which in turn leads tohardening. Traditional immunohis-

FIGURE 9. Interaction between Mfp-1 and Mfp-2 films adsorbed to mica.Separation (out) following a brief contact. Prolonged contact did not improveadhesion. An arrow (top left) denotes a hard wall.

FIGURE 10. Imaging of a byssal adhesive plaque section. A, scanning electron micrograph of a sectionedbyssal plaque from M. galloprovincialis. The boxed region was subjected to Raman microscopy. Raman imagesof the same region were integrated for CH-stretching (2850 –3010 cm�1) (B), iron-DOPA (490 – 696 cm�1)(C), and phenylalanine (980 –1020 cm�1) (D), respectively. E, average spectra of the three morphologicallydistinct domains in the plaque i.e. cuticle, core, and plaque-substrate interface. F, the distribution of thesespectra is visualized through least square fitting. Scale bar below Raman images is 10 �m.

Major Plaque Protein Interactions

AUGUST 13, 2010 • VOLUME 285 • NUMBER 33 JOURNAL OF BIOLOGICAL CHEMISTRY 25855

by guest on April 4, 2016

http://ww

w.jbc.org/

Dow

nloaded from

tochemical localization ofDOPA-containing proteins in a varietyof sclerotized structures has proven useless due to the speedwith which DOPA epitopes change following oxidation andmetal binding (31–32).In this study, we have explored two new approaches for

insights relating to protein interactions in sclerotized struc-tures: the Raman microscope and surface forces apparatus.Ramanmicroscopy enables analysis of biological specimens forchemical functionalities, secondary structure, and orienta-tional preference with a 1-�m spatial resolution, and the sur-face forces apparatus allows a facile means of quantifying inter-actions between polymers or between polymers and surfaces.By depositing paired polymer candidates on opposingmica sur-faces, the strength of each pairwise interaction can bemeasuredand compared. Given that byssal plaques adhere strongly tomica surfaces in seawater (9), mica is suitable as a support sur-face to which various MFPs can be adsorbed, and their interac-tions can be examined at salinities approaching seawater. pH5.5 hardly resembles seawater pH (8.2), but it is a reasonableapproximation for the intragranular pH of regulated proteinsecretion (33–34) and, as such, will resemble the pH of MFPssecreted onto a surface. Most of theMFPs and Fe3� precipitateat pH � 8; in addition, we observed in a previous study (3) thatplaque footprints preserved the 20 mol % DOPA content ofMfp-3 notwithstanding the pH of the surrounding seawater.This would be possible only if the footprints weremaintained atacidic pH or were full of antioxidants. Given the above factors,pH 5.5 was the most expedient for this study.Mfp-2 is the most abundant protein in the attachment

plaques of the mussel byssus and thus cannot afford to be aweak link in the structure. At the top of the plaque, it needs tointeract with the scaffolding collagens (prepepsinized collag-

ens) and associated matrix proteinsthat join the thread to the plaque(Fig. 1); at the bottom, Mfp-2 mustbind to the priming proteins (Mfp-3and/or Mfp-5) that secure the foot-print of the plaque to foreign sur-faces. An Mfp-2 film was incapableof adhering to bare mica (asymmet-ric). No improvement was obtainedby coating both mica surfaces withMfp-2 (symmetric). The resultsimply that Mfp-2 is not an adhesiveprotein, a result that resembles thebehavior of Mfp-1, a protein knownto function only as a coating in bys-sus (9, 11).Given that iron and calcium are

present in adhesive plaques at levels(�0.1 w/w%) significantly abovethose of seawater (29), we exploredthe effect of adding Ca2� and Fe3�

on adhesion between symmetricMfp-2 films. Mfp-2 showed weakcalcium-mediated adhesion con-sistent with electrostatic interac-tions, but the effect of ionic strength

needs further scrutiny. Notwithstanding the high positivecharge density and high pI of Mfp-2, there are clusters of neg-atively charged residues onEGF repeat 8 and atC andN termini(Fig. 2,A andB). These clusterswould be repelled by the anionicmica surface andpresumably serve as ligands forCa2� ions (Fig.2). Separate from these, the seventh EGF repeat contains a cal-cium binding motif with a known consensus sequence Cys3-x-Asn-x-x-x-x-Tyr-x-Cys4 (35) having a micromolar bindingconstant (36).Fe3� increased adhesion by a factor of 5 to 7 timesmore than

that of calciumduring the separation of symmetricMfp-2 films.Notably, the measured adhesion energy (Wad� Fad/1.5�R) was2.2 mJ/m2, which approaches the 10 mJ/m2 energy measuredfor the strongest known noncovalent protein-ligand interac-tion, i.e. biotin-avidin (28).Given the incremental coordination of Fe3� by low molecu-

lar weight catechols (from one to three) with increasing pH(37), the prediction that the adhesion of iron bridging inMfp-2would increase with pH seemed reasonable, but there was nochange. We suggest that in Mfp-2 as in Mfp-1, iron bindingalready has achieved the maximal tris-catecholate iron at pH5.5 (38). Although complex stoichiometry is fixed at the 3:1limit, its formation can be accelerated, and this is probablywhatwe observed in the enhanced rate of adhesion (11, 38).There is another pH-dependent result that merits scrutiny.

The reduced adhesion post-1 h separation and reapproach atpH 6.8 (Fig. 6D) was half of that occurring at pH 5.5. Given theroughly 2-fold thicker hard wall at pH 6.8, the higher rate ofcomplex formationmay render the filmmore stiff and less com-pliant than at pH 5.5. The larger recovered adhesion at pH 5.5may be stronger because the film can flow and rearrange toform new complexes, whereas at pH 6.8, it is more rigid.

FIGURE 11. Relative Raman intensities for Mfp-2 with different admixtures of iron. No added Fe3� (bot-tom) and with Fe3� added at DOPA:iron ratios of 20:1 (middle) and 2:1 (top). Fe3� was added to Mfp-2 at pH 5.5,after which the protein-iron complex was precipitated by raising the pH to 8. Tris-DOPA-iron complexes (asshown) are detected primarily by the ring and -O-iron resonances indicated (arrows). Excitation laser was at 532nm. Inset spectrum shows actual intensities normalized to CH band (2850 –3010 cm�1) to emphasize the Ramanenhancement by resonance. a.u., arbitrary units.

Major Plaque Protein Interactions

25856 JOURNAL OF BIOLOGICAL CHEMISTRY VOLUME 285 • NUMBER 33 • AUGUST 13, 2010

by guest on April 4, 2016

http://ww

w.jbc.org/

Dow

nloaded from

The strength of adhesion in Mfp-2 is reminiscent of iron-induced adhesion of symmetricMfp-1, which was attributed toformation of multiple catecholate iron complexes (11, 38).Curiously, even though all MFPs have DOPA, iron bridgingseemsMFP-specific; iron addition to asymmetricMfp-2-Mfp-1films had no effect. In this regard, it should be noted that thereis no evidence that Mfp-1 and -2 make direct contact in theplaque. These results indicate that the intrinsically poor cohe-sion betweenMfp-2 films, at least as measured by the SFA, canbe overcome by the addition of metal ions.Ramanmaps of plaque thin sections support a role for iron

in setting. This conclusion was also proposed by Sever et al.(39) based on EPR analysis of plaques. Given the strong phe-nylalanine signal coupled with the Fe3�-catechol resonancesignals localized in the plaque core, Mfp-2 is most likely theFe3� binding protein in this region. Mfp-2 has a significantPhe content (7–8 residues per protein), whereas Mfp-1,Mfp-3, andMfp-5 have none. Indeed, at DOPA:iron ratios of20:1, Mfp-2 and the plaque core have highly similar Ramanspectra. The observed difference between the resonancespectra of the plaque cuticle and core probably reflects theeffect of different protein/metal mixtures on the coordina-tion environment. Mfp-2 and Mfp-1 are both capable ofcoordinating Fe3�, but themussel apparently tunes the coor-dination chemistry and environment of the different pro-teins during processing by titrating the supply of iron to theprecursors in the secretory granules.The reliance of reversible Mfp-2-Mfp-2 interactions on two

different metal ions Ca2� and Fe3� may be a fascinating exam-ple of a biological “safety net.” As the two binding sites aredistinct, DOPA for iron and a specific binding motif for cal-cium, both may be present at any given interface. Assuming aparallel binding array, the stronger (iron-DOPA)may yield dur-ing a separation with a second weaker bridge (calcium binding)

providing guidance for contact recovery. Notwithstanding thatthis model should be tested directly by SFA, a high initial stiff-ness and gradual post-yield recovery is generally observed inbyssal threads under tension (40–41).At the interface between the plaque and the foreign substra-

tum, three MFPs have been detected, namely Mfp-3, Mfp-5,andMfp-6 (2, 3). Of these, onlyMfp-3 andMfp-5 are known tobe strongly adhesive (9).5 Indeed, Mfp-3 and Mfp-5 haveinspired a wide range of DOPA-like synthetic adhesive poly-mers (e.g. 6). The interaction of Mfp-5 and Mfp-2 was instan-taneous, very strong, and reversible. The binding functional-ities for this protein-protein interaction are not at all clear atpresent. In contrast, the interaction betweenMfp-2 andMfp-3is a very unconvincing one with no instantaneous adhesion. Atthis time, the best interpretation of the results is that Mfp-3completely displaces Mfp-2 from one mica surface to becomethe only adhesive protein to span the mica sheets. Indeed, thetime-dependent adhesion achieved between Mfp-3 and Mfp-2is no different than with Mfp-3 alone (9). Presumably, bindingto surface Mfp-3 is mediated by some other plaque protein(s).In summary, the interactions of Mfp-2 in the plaque are as

follows (Fig. 12); Mfp-2 binds strongly toMfp-5 at the interfaceand to other Mfp-2 proteins in the core in the presence of Fe3�

and/or Ca2�. Mfp-2 does not bind Mfp-1 nor apparently toMfp-3; binding toMfp-5 is not enhanced by adding iron. Otherinteractions remain to be investigated.

Acknowledgments—We thank C. J. Sun and E. Danner for partiallypurifying Mfp-2 and J. Yu and P. Fratzl for support.

REFERENCES1. Waite, J. H., Holten-Andersen, N., Jewhurst, S. A., and Sun, C. J. (2005) J.

Adhesion 81, 297–3172. Zhao, H., Robertson, N. B., Jewhurst, S. A., and Waite, J. H. (2006) J. Biol.

Chem. 281, 11090–110963. Zhao, H., and Waite, J. H. (2006) J. Biol. Chem. 281, 26150–291584. Lee, H., Scherer, N. F., andMessersmith, P. B. (2006) Proc. Natl. Acad. Sci.

U.S.A. 103, 12999–130035. Dalsin, J. L., Lin, L., Tosatti, S., Voros, J., Textor,M., andMessersmith, P. B.

(2005) Langmuir 21, 640–6466. Lee, B. P., Chao, C. Y., Nunalee, F. N., Motan, E., Shull, K. R., andMesser-

smith, P. B. (2006)Macromolecules 39, 1740–17487. Westwood, G., Horton, T. N., andWilker, J. J. (2007)Macromolecules 40,

3960–39648. Wang, J., Tahir, M. N., Kappl, M., Tremel,W., Metz, N., Barz, M., Theato,

P., and Butt, H. J. (2008) Advanced Materials 20, 3872–38769. Lin, Q., Gourdon, D., Sun, C., Holten-Andersen, N., Anderson, T. H.,

Waite, J. H., and Israelachvili, J. N. (2007) Proc. Natl. Acad. Sci. U.S.A. 104,3782–3786

10. Sagert, J., and Waite, J. H. (2009) J. Exp. Biol. 212, 2224–223611. Harrington,M. J.,Masic, A., Holten-Andersen, N.,Waite, J. H., and Fratzl,

P. (2010) Science 328, 216–22012. Holten-Andersen, N., Mates, T. E., Toprak, M. S., Stucky, G. D., Zok,

F. W., and Waite, J. H. (2009) Langmuir 25, 3323–332613. Waite, J. H., Qin, X. X., and Coyne, K. J. (1998)Matrix Biol. 17, 93–10614. Rzepecki, L. M., Hansen, K. M., andWaite, J. H. (1992) Biological Bulletin

183, 123–13715. Inoue, K., Takeuchi, Y., Miki, D., and Odo, S. (1995) J. Biol. Chem. 270,

6698–670116. Kohda, D., and Inagaki, F. (1992) Biochemistry 31, 11928–1193917. Suci, P. A., and Geesey, G. G. (2001) Colloids Surf. 22B, 159–16818. Fant, C., Elwing, H., and Hook, F. (2002) Biomacromolecules 3,

FIGURE 12. Summary of protein interactions in the byssal plaque ofMytilus. Mfp-3 and -5 are strongly surface active (black arrows). Mfp-2interacts directly with Mfp-5 and with itself in the presence of iron orcalcium. There is no interaction with Mfp-1, and interactions with prepep-sinized collagens (preCOLs), thread matrix protein-1 (tmp), and Mfp-4remain to be determined. The numerical values denote specific adhesionenergies at pH 5.5, and the value for the interaction between Mfp-3 andmica is from Ref. 9. Question mark denotes unknown interactions.

Major Plaque Protein Interactions

AUGUST 13, 2010 • VOLUME 285 • NUMBER 33 JOURNAL OF BIOLOGICAL CHEMISTRY 25857

by guest on April 4, 2016

http://ww

w.jbc.org/

Dow

nloaded from

732–74119. Waite, J. H. (1995)Methods Enzymol. 258, 1–2020. Waite, J. H., and Qin, X.-X. (2001) Biochemistry 40, 2887–289321. Papov, V. V., Diamond, T. V., Biemann, K., andWaite, J. H. (1995) J. Biol.

Chem. 270, 20183–2019222. Israelachvili, J. N., and Adams, G. E. (1978) J. Chem. Soc. Faraday Trans. I

74, 975–100123. Israelachvili, J. N. (1987) Proc. Natl. Acad. Sci. U.S.A. 84, 4722–472424. Israelachvili, J. N., and McGuiggan, P. M. (1990) J. Mater. Res. 5,

2223–223125. Israelachvili, J. N. (1992) Intermolecular and Surface Forces, 2nd Ed., Ac-

ademic Press, San Diego26. Taylor, S.W., Chase, D. B., Emptage, M. H., Nelson, M. J., andWaite, J. H.

(1996) Inorg. Chem. 35, 7572–757727. Johnson, K. L., Kendall, K., and Roberts, A. D. (1971) Proc. R. Soc. London

324, 301–31328. Helm, C. A., Knoll, W., and Israelachvili, J. N. (1991) Proc. Natl. Acad. Sci.

U.S.A. 88, 8169–817329. Sun, C., and Waite, J. H. (2005) J. Biol. Chem. 280, 39332–3933630. Movasaghi, Z., Rehman, S., and Rehman, I. U. (2007) Appl. Spectrosc.

Review 42, 493–54131. Anderson, K. E., and Waite, J. H. (2000) J. Exp. Biol. 203, 3065–307632. Robinson, M. W., Colhoun, L. M., Fairweather, I., Brennan, G. P., and

Waite, J. H. (2001) Parasitology 123, 509–51833. De Young, M. B., Nemeth, E. F., and Scarpa, A. (1987) Arch. Biochem.

Biophys. 254, 222–23334. Njus, D., Sehr, P. A., Radda, G. K., Ritchie, G. A., and Seeley, P. J. (1978)

Biochemistry 17, 4337–434335. Campbell, I. D., and Bork, P. (1993) Curr. Opin. Struct. Biol. 3, 385–39236. Rand, M. D., Lindblom, A., Carlson, J., Villoutreix, B. O., and Stenflo, J.

(1997) Protein Science 6, 2059–207137. Avdeef, A., Sofen, S. R., Bregante, T. L., and Raymond, K. N. (1978) J. Am.

Chem. Soc. 100, 5362–537038. Zeng, H., Hwang, D. S., Israelachvili, J., andWaite, J. H. (2010) Proc. Natl.

Acad. Sci. U.S.A., 10.1073/pnas.100741610739. Sever, M. J., Weisser, J. T., Monahan, J., Srinivasan, S., and Wilker, J. J.

(2004) Angew. Chem. Int. Ed. 43, 448–45040. Harrington, M. J., and Waite, J. H. (2007) J. Exp. Biol. 210, 4307–431841. Carrington, E., and Gosline, J. M. (2004)AmericanMalacological Bulletin

18, 135–142

Major Plaque Protein Interactions

25858 JOURNAL OF BIOLOGICAL CHEMISTRY VOLUME 285 • NUMBER 33 • AUGUST 13, 2010

by guest on April 4, 2016

http://ww

w.jbc.org/

Dow

nloaded from

1

PROTEIN- AND METAL-DEPENDENT INTERACTIONS OF A PROMINENT PROTEIN IN MUSSEL ADHESIVE PLAQUES*

Dong Soo Hwang1, Hongbo Zeng2, Admir Masic3, Matthew J. Harrington3, Jacob N. Israelachvili4 and J. Herbert Waite5

1Materials Research Laboratory, University of California, Santa Barbara, CA 93106, USA 2Dept Chem. & Mat. Eng., University of Alberta, Edmonton, AB, T6G 2V4, Canada 3Dept Biomaterials, Max Planck Institute for Colloids and Interfaces, Potsdam-Golm, Germany 4Department of Chemical Engineering, University of California, Santa Barbara, CA 93106 5Marine Science Institute, University of California, Santa Barbara, CA 93106

Supporting Data Scheme S1: Mfps purification Scheme S2: Raman microscopy data processing Scheme S3: Surface Forces Apparatus diagram Figure S1: Interaction of mfp-5 and –2 by SFA Figure S2: Interaction of mfp-3 and –2 by SFA Figure S3: Confocal Raman microscopy of precipitated mfp-1 and mfp-2 complexed to Fe3+

Mussel Foot Protein Purification. The purification of each Mfp is described in

detail elsewhere. Salient steps and solvents are summarized here and in Scheme S1with particular emphasis on innovations. It should be emphasized that these purifications were adapted from earlier procedures used extensively for histones; mussel proteins and histones resemble one another in their high charge density and pI, as well as a dearth of secondary structure in solution. Mfps used for this study were prepared from Mytilus edulis feet. Mfp-1 and mfp-2 were extracted from dissected phenol glands using about 100 mL/5g cold 5% (v/v) acetic acid with 0.1 mM leupeptin and 0.1 mM pepstatin. Glands were homogenized with 50 mL tissue grinders from Kontes (Vineland, NJ), and homogenates were spun in a refrigerated (5˚ C) centrifuge 20,000xg for 40 min to produce a supernatant S1 and pellet P1. Mfp-1 and -2 were further purified from S1 by the addition of 1.5 % (v/v) perchloric acid (resulting insolubles were removed by centrifugation as before). Mfp-1 and -2 were harvested from solution by the dropwise addition of 2 volumes of cold acetone exactly as described in Waite (1995). The concentrate was subjected to gel filtration on a Shodex KW 803 eluted with 5% acetic acid at 0.25 ml/min at room temperature and monitored at 280 nm (Ohkawa et al., 2004). Mfp-1 (108 kDa) typically elutes just after the void volume, whereas mfp-2 (45 kDa), depending on the sample volume injected, elutes as a peak about 10 min after mfp-1. Fractions under the peaks were examined for their mfp content using acid-urea gel electrophoresis (Waite, 1995). Those fractions of mfp-1 without mfp-2, or mfp-2 without mfp-1, respectively, were pooled, freeze-dried, reconstituted in a small volume of 5% acetic acid (~1 mL), and polished by C8 reverse-phase HPLC using an acetonitrile gradient in water with 0.1% trifluoroacetic acid. Both proteins elute at about 22% acetonitrile. Following HPLC, the proteins were freeze-dried and stored at –80 ˚C. Mfp-3 and –5 were obtained by further extraction of P1. Mfp-3 was extracted by homogenization (tissue grinder) of pellets with cold 5% acetic acid with 8M urea (50

2

mL/5g). Centrifugation (20,000xg @ 40 min) of this resulted in a pellet P2 and supernatant S2. S2 was collected and 30 % weight/volume ammonium sulfate was added and stirred 1 h at room temperature. Insolubles were removed by centrifugation (as above) and the supernatant was dialyzed against Q-water at 5 ˚C with dialysis tubing MWCO 1,000 (Spectrum Industries). Mfp-3 forms a floc and settles during dialysis. This was harvested by a 5min spin at 15K xg on a microfuge and redissolved in a small volume of 5% acetic acid (Papov et al., 1996). Mfp-5 was extracted from P2 by homogenization (tissue grinder) of pellets using cold 5% acetic acid with 6M guanidine·HCl (50 mL/5g). Centrifugation (20,000xg @ 40 min) of this resulted in a pellet P3 and supernatant S3. S3 was collected and 30 % weight/volume ammonium sulfate was added and stirred 1 h at room temperature. Insolubles were removed by centrifugation (as above), and the supernatant was dialyzed against Q-water at 5 ˚C with dialysis tubing MWCO 1,000. Mfp-5 forms a floc and settles during dialysis. This was harvested by a 5min spin at 15K x g on a microfuge and redissolved in a small volume of 5% acetic acid (Waite & Qin, 2000). Mfp-3 and mfp-5 were polished by C8 HPLC at a flow rate of 1 ml/min using a linear gradient of acetonitrile in water with 0.1 % trifluoroacetic acid. Elution times are well separated.

Scheme S1. Isolation strategy of mfps from Mytilus feet. Every mfp precursor can be isolated from a different step of the same starting lot. See supporting text for details on purification to homogeneity. Waite, J. H. (1995) Meth. Enzymol. 258, 1-20. Ohkawa, K., Nishida, A., Yamamoto, H., Waite, J.H. (2004) Biofouling 20, 101-115.

Raman spectroscopic studies - Plaques from Mytilus galloprovincialis were embedded in PEG-2000 (Carl Roth Gmbh), and 20 µm thick longitudinal sections were microtomed. Plaque sections were washed thoroughly with several changes in distilled

3

water to remove any remaining PEG, positioned on a quartz slide in distilled water and fixed under a quartz cover slip. For Raman microspectroscopy, a continuous laser beam was focused on the sample through a confocal Raman microscope (model CRM200, WITec, Ulm, Germany) equipped with a piezo-scanner (model P-500, Physik Instrumente, Karlsruhe, Germany). The diode-pumped 785 nm near infrared (NIR) laser excitation (Toptica Photonics AG, Graefelfing, Germany) was used in combination with a 100µm oil immersed (Nikon, NA = 1.25) microscope objective. Laser power ranging between 15-30 mW was used for all measurements. The spectra were acquired using an air-cooled CCD (DU401A-DR-DD, Andor, Belfast, North Ireland) behind a grating (300 g mm-1) spectrograph (Acton, Princeton Instruments Inc., Trenton, NJ, USA) with a 6 cm-1 spectral resolution. Software ScanCtrlSpectroscopyPlus (version 1.38, Witec) was used for measurement setup. Raman spectra were processed and analyzed with Witec Project Plus software (Version 2.02).

Single Raman spectra were collected from different regions of the plaque (coating, foam and interface) by averaging 30 acquired spectra, each with a 1 s integration time. Acquired spectra were used as the single spectra basis for further analysis (Scheme 2). A 2-dimensional multi-spectral data set was collected by scanning the surface of the cross section in steps of 0.5-µm, recording spectra with an integration time of 1 s per each scan (pixel). All Raman spectra were background-subtracted and lightly smoothed using the first order polynomial function and 9-point Savitzky-Golay filter (4th order polynomial), respectively.

Raman images in Fig.9 B, C and D were generated using a sum filter, which integrates the intensity of the signal for a defined wave number range of interest after subtracting the background as a linear baseline from the upper to lower border. The ranges of interest were the bands for CH stretching (2850-3010cm-1), Fe-dopa (490-696 cm-1), and phenylalanine (980-1020 cm-1). The color images were obtained from Witec Project software using the basis analysis and image color combination functions. In the case of basis analysis, the algorithm fits each spectrum of the multi-spectral data set with a linear combination of the basis single spectra (tree spectra in figure 9E) using the least squares method. To solve the problem of differing fluorescence background in various parts of the sample, the first derivative of both multi-spectral data set and basis single spectra was performed. The weighting factors of various components obtained by fitting were stored in an image and combined in a false color bitmap using the image color combination function. Supplementary Scheme S2 summarizes the Raman data processing procedure.

4

Scheme S2: Flowchart of data processing in Raman microscopy.

For Raman microscopy of purified mfps, a purified solution of mfp-2 (1 mg/ml) in 0.1M sodium acetate, 0.25M KNO3, 1mM bis-tris (pH 5.5) was mixed with a solution of FeCl3 in 10mM Bis-Tris (pH 5.5) at ratios of dopa:Fe3+ of 2:1 and 20:1 and allowed to equilibrate for 10 minutes. The protein precipitated as the pH was raised to ~8.0 with 0.1M NaOH. The droplet was allowed to evaporate on a glass slide, and Raman spectra (30 accumulations with 1 s integration time each) were taken from the precipitated protein residue at the edges. Raman spectra were also measured from the protein solution prior to iron complexation and from solutions of mfp-1 and mfp-2 in which Fe3+ was not limiting for comparison.

Force vs distance profiles measurement by the Surface Forces Apparatus (SFA)- The SFA technique has been used for many years to measure both normal and lateral forces between surfaces in vapors and liquids, e.g., van der Waals forces, electrostatic forces, adhesion forces, friction and lubrication forces, hydrophobic interactions, specific and non-specific biological interactions (22-25). A typical SFA experiment setup for measuring normal forces between two surfaces is shown in Scheme S3. SFA can accurately measure normal (attractive adhesion or repulsive) forces F as low

Scheme S3. Surface forces diagram. Experiment setup using a Surface Forces Apparatus (SFA) to measure the normal interactions between two surfaces, (a) schematic of SFA experiment setup for normal force measuremment, (b) illustration of typical force-distance profile (positive force values present repulsion, and negative values show attraction), (c) illustration of the experimental geometry for the study of interfactions between two protein layers.

as 10 nN as a function of surface separation distance D with a resolution of less than 1 Å monitored in situ using the fringes of equal chromatic order (FECO) in multiple beam

5

interferometry. The force between the two surfaces is measured according to Hooke’s law, F = kΔx , Where k is the spring constant supporting the lower surface, andΔx = Dactual −Dapplied is the difference between applied and actual surface separations. The normal force-distance profiles and adhesion forces (Fad) of mfp-2 were determined using an SFA in a configuration reported previously (9). Briefly, a thin mica sheet of 1-5 m was glued onto a cylindrical silica disk (radius R=2 cm). 100 µL of a stock mfp-2 solution (20 µg/mL) diluted in 0.1 M sodium acetate with 0.25 M potassium nitrate at pH 5.5, was injected onto one mica surface. [n.b. In the SFA, potassium nitrate is used in place of NaCl to reduce chloride ion induced corrosion of the semi-reflecting silver layers under the mica substrates.].

For experiments with Fe, 1mM bis-tris was added to stabilize the solubility of Fe3+ (26). Although the test pH is significantly lower than seawater pH 8.2, it is consistent with the pH the adhesive secretion by the foot (unpublished observations) and was necessitated by the poor solubility of Fe and the mussel proteins at higher pH in the SFA. The two curved and coated mica surfaces were then mounted in the SFA chamber in a crossed-cylinder geometry, which roughly corresponds to a sphere of radius R on a flat surface based on the Derjaguin approximation: F(D) = 2πRW(D), where F(D) is the force between the two curved surfaces and W(D) the interaction energy per unit area between two flat surfaces. The measured adhesion or “pull-off” force Fad is related to the adhesion energy per unit area Wad by Fad=2πRWad for rigid (undeformable) surfaces with weakly adhesive interactions, and by Fad=1.5πRWad (used in this study) for soft deformable surfaces with strong adhesive contact (27, 28). All experiments were performed at room temperature (23 °C). We applied proteins to mica surfaces according to the whether asymmetric and symmetric testing mode was to be used. In the asymmetric mode, protein was applied to one mica surface only. In contrast, protein was applied to both mica surfaces in symmetric mode.

6

Figure S1. Interaction of mfp-5 and mfp-2 films in the presence of 5µM Fe3+. Contact times in min (0, A), 10 (B), 60 (C)

(A)

(C)

(B)

7

Figure S2. Interaction of mfp-3 and mfp-2 films in the presence of 5µM Fe3+. Contact times in min (0, A), 10 (B), 60 (C). Blue –approach, red - separation. Contact position was the same.

(A)

(C)

(B)

8

Figure S3. A. Raman spectra of purified mfp-1 and mfp-2 precipitated by raising pH to from 5 to 8.0 in the presence of Fe3+ at an iron-Dopa ratio of 1:2. The two spectra are essentially identical. Peaks denoted by * at 930 and 1050 cm-1 are buffer artifacts. B. Three Raman spectra of crystals formed upon drying the buffer 0.1M sodium acetate, 0.25M KNO3, 1mM bis-tris (pH 5.5) on a glass slide. Inset is full scale.

A

Israelachvili and J. Herbert WaiteDong Soo Hwang, Hongbo Zeng, Admir Masic, Matthew J. Harrington, Jacob N.

Adhesive PlaquesProtein- and Metal-dependent Interactions of a Prominent Protein in Mussel

doi: 10.1074/jbc.M110.133157 originally published online June 21, 20102010, 285:25850-25858.J. Biol. Chem. 

  10.1074/jbc.M110.133157Access the most updated version of this article at doi:

 Alerts:

  When a correction for this article is posted• 

When this article is cited• 

to choose from all of JBC's e-mail alertsClick here

Supplemental material:

  http://www.jbc.org/content/suppl/2010/06/21/M110.133157.DC1.html

  http://www.jbc.org/content/285/33/25850.full.html#ref-list-1

This article cites 40 references, 16 of which can be accessed free at

by guest on April 4, 2016

http://ww

w.jbc.org/

Dow

nloaded from