new supporting online material for - science · 2011. 6. 1. · 2. ftir data 3. 2d-ir data 4. model...

24
www.sciencemag.org/cgi/content/full/332/6034/1206/DC1 Supporting Online Material for Residue-Specific Vibrational Echoes Yield 3D Structures of a Transmembrane Helix Dimer Amanda Remorino, Ivan V. Korendovych, Yibing Wu, William F. DeGrado, Robin M. Hochstrasser* *To whom correspondence should be addressed E-mail: [email protected] Published 3 June 2011, Science 332, 1206 (2011) DOI: 10.1126/science.1202997 This PDF file includes: Materials and Methods SOM Text Figs. S1 to S10 Tables S1 and S2 Full Reference List

Upload: others

Post on 20-Oct-2020

2 views

Category:

Documents


0 download

TRANSCRIPT

  • www.sciencemag.org/cgi/content/full/332/6034/1206/DC1

    Supporting Online Material for

    Residue-Specific Vibrational Echoes Yield 3D Structures of a Transmembrane Helix Dimer

    Amanda Remorino, Ivan V. Korendovych, Yibing Wu, William F. DeGrado, Robin M. Hochstrasser*

    *To whom correspondence should be addressed E-mail: [email protected]

    Published 3 June 2011, Science 332, 1206 (2011)

    DOI: 10.1126/science.1202997

    This PDF file includes:

    Materials and Methods SOM Text Figs. S1 to S10 Tables S1 and S2 Full Reference List

  • 1

    Supporting Online Material

    Content Summary:

    1. Material and methods

    2. FTIR data

    3. 2D-IR data

    4. Model to extract the coupling constants

    5. Fit of the FTIR and 2D-IR spectra

    6. Structure determination

    1. Materials and Methods:

    Peptide synthesis. Twelve different peptides were synthesized by introducing 13

    C=18

    O labels on

    each of the eleven residues L10 to L20 of CGGKPIWWVL10VGVLGGLLLL20TILVLAMWKK

    in addition to an unlabeled peptide. N-Fmoc-1-13

    C=18

    O labeled amino acids (Gly, Leu, Val) were

    prepared from the corresponding N-Fmoc-1-13

    C labeled amino acids (Cambridge Isotope

    Laboratories, Andover, MA) and H218

    O according to the literature procedures (15) allowing for

    at least 90% isotopic enrichment in 18

    O, as evidenced by ESI-MS spectra of the products. The

    TM portion of human IIb integrin (965-990) was additionally flanked with Lys residues and a

    CGG linker. The peptides were synthesized on a PTI Symphony automated peptide synthesizer

    using standard Fmoc protocols on a 0.05 mmol scale using a Fmoc-PAL-PEG -PS resin (Applied

    Biosystems) with a substitution level of 0.21 mmol/g. Activation of the free amino acids (5 fold

    excess) was achieved with 0.95 equiv (relative to the amino acid) excess of HATU in the

    presence of 10 equiv of diisopropylethylamine (DIEA). The reaction solvent contains 25%

    dimethylsulfoxide (DMSO) and 75% N-methylpyrrolidone (NMP) (HPLC grade, Aldrich). Side

    chain deprotection and simultaneous cleavage from the resin was performed using a mixture of

    trifluoroacetic acid (TFA)/triethylsilane/water/ethanedithiol (94:2.5:2.5:1 v/v) at room

    temperature, for 3 hours. After filtration most of the solvent was evaporated using a stream of

  • 2

    N2. The crude peptides collected from precipitation with cold diethyl ether (Aldrich) were dried

    in vacuo. The peptides were then purified on a preparative reverse phase HPLC system (Varian

    ProStar 210) with a C4 preparative column (Vydac) using a linear gradient of buffer A (0.1%

    TFA in Millipore water) and buffer B (6:3:1 2- propanol:acetonitrile:water) containing 0.1%

    TFA. The identities of the purified peptides were confirmed by MALDI-TOF mass spectroscopy

    on a Voyager Biospectrometry Workstation (PerSeptive Biosystems), and their purity was

    assessed using HP1100 analytical HPLC system (Hewlett Packard) with an analytical C-4

    column (Vydac) and a linear A/B gradient.

    Stock solutions in trifluoroethanol (TFE) were prepared from the lyophilized powder. The

    samples for 2D-IR measurements were prepared by mixing the stock of peptides with the 100

    mM stock of dodecyl phosphatidylcholine (DPC), the solvent was removed in the stream of

    nitrogen and the resulting films were dried in vacuo overnight to remove all of the organic

    solvent leftover. The peptide-detergent film was then dissolved in D2O (20 mM phosphate

    buffer, pH (uncorrected) 7.4) to the final concentration of peptide of 4 mM and the detergent of

    200 mM.

    Analytical Ultracentrifugation. Equilibrium sedimentation was used primarily to determine the

    association state of the peptides and to provide an estimate of the association constants. The

    experiments were performed in a Beckman XL-I analytical ultracentrifuge (Beckman Coulter)

    using six-channel carbon-epoxy composite centerpieces at 25 °C. Peptides were co-dissolved in

    TFE (Sigma) and DPC (Avanti Polar Lipids). The organic solvent was removed under reduced

    pressure to generate a thin film of peptide/detergent mixture, which was then dissolved in buffer

    previously determined to match the density of the detergent component (10 mM phosphate

    buffer (pH= 7.4), containing 500 mM tris(2-carboxyethyl)phosphine (TCEP) in 52% D2O). The

    final concentration of DPC is 15 mM in all of the samples. Samples were prepared in a total

    peptide concentration of 38 µM. Data at different measurement speeds (35, 40, 45, 48 and 50

    krpm) were analyzed by global curve-fitting of radial concentration gradients (measured using

    optical absorption) to the sedimentation equilibrium equation for monomer-dimer equilibria

    among the peptides included in the solution. Peptide partial specific volumes were calculated

    using previously described methods (30) and residue molecular weights corrected for the 52%

    D2O exchange expected for the density-matched buffer. The solvent density (1.059 g/ml) was

  • 3

    measured using a Paar densitometer. Sedimentation equilibrium data were fit using Igor Pro

    (Wavemetrics) to the following equation:

    Abs E ac0alexp 2

    2RTMa (r

    2 r02)

    2a

    c0a2

    Kalexp

    2

    2RT2Ma (r

    2 r02)

    where E = baseline (zero concentration) absorbance, coa is the molar concentration of monomeric

    IIb TM peptide at ro, ea is the molar extinction coefficient for IIb TM peptide at 280 nm, l is

    the optical path length, =2*rpm, R= 8.3144 107 erg K

    -1mole

    -1 , T is temperature in K, Ma is

    the buoyant molecular weight of monomeric IIb; Ka is the homodimeric dissociation constant for

    IIb TM peptide.

    Molecular weight was obtained from the buoyant molecular weight using:

    Mw M(1 v_

    )

    where M is the buoyant molecular weight,

    v_

    is the partial specific volume and

    is the solution

    density.

    Figure S1: Sedimentation equilibrium

    profile at 280 nm of IIb TM peptide (38

    µM) in density matched DPC micelles (15

    mM) in phosphate buffer (10 mM, pH

    7.4). The partial specific volume and the

    solution density were fixed at 0.80057 mL/g

    and 1.059 g/mL. The data was analyzed

    using a global fitting routine. The molecular

    weight was held at 3300 and the data were fit

    to dissociation constant of 0.013

    (peptide/detergent molar units).

  • 4

    FTIR Spectra. The FTIR spectra (Nicolet 6700) were corrected with a DPC and D2O

    background subtraction. The second step involved the isolation of the isotopically substituted

    amide I transitions from those of the 12

    C=16

    O main band. This was a difficult task given that in

    the majority of the diluted cases, no peaks could be identified in the 13

    C=18

    O region. For this

    purpose, the isotopically substituted region (1575-1615 cm-1

    ) was devoid of the 13

    C=18

    O amide I

    transition known from the 2D-IR spectra and the remaining baseline was fitted with a Gaussian.

    This Gaussian baseline was subtracted from the original spectrum resulting in the data shown in

    Fig. 2 and the ones shown here. The accuracy of the subtractions was evaluated by comparison

    of the obtained FTIR baseline to that of the background of the 2D-IR spectra.

    2D-IR Spectra. The 2D-IR spectra were obtained by methods previously detailed (6). The echo

    signal field generated at frequency t following a sequence of three infrared pulses was

    measured as a function of the initial coherence frequency, . Each 2D-IR spectrum of vs. t

    is recorded at a particular choice of the waiting time delay, T, between the second and third

    pulses. The positive maxima in the 2D-IR are displaced to higher t frequencies because of the

    interference with the negative v=1→2 portion.

  • 5

    2. FTIR data:

    The DPC and D2O subtracted FTIR spectra of all the 100% and 10% 13

    C=18

    O labeled samples

    can be seen in Fig. S2.

    Figure S2: Subtracted FTIR

    spectra. FTIR spectra of (A)

    100% and (B) 10% 13

    C=18

    O

    labeled samples. The

    differences in OD between

    samples shown in panel A is

    given mostly by the variability

    in the amount of dissolution of

    the peptides after

    lyophilization. The amide I

    main band peak appears at 1656

    cm-1

    and a secondary peak at ca. 1635 cm-1

    is attributed to small amounts of exposed amide I modes (31). The

    transition at approx 1675 cm-1

    is trifluoro acetic acid (TFA) whose further removal was avoided to minimize

    aggregation.

    The normalized FTIR spectra are also presented in Fig. S3 in order to compare the line shapes of

    all the different samples. For most of the 10% labeled samples no 13

    C=18

    O bands could be

    observed in the FTIR spectra but were present in the 2D-IR spectra.

    Figure S3: Normalized FTIR

    spectra: Normalized FTIR

    spectra of (A) 100% and (B) 10% 13

    C=18

    O labeled samples.

    The procedure used to isolate these transitions is shown in Fig. S4.

    Figure S4: FTIR spectra

    background subtraction. Method

    used to separate the 13

    C=18

    O band

    from the tail of the main band for (A)

    G12 100% and (B) 10% 13

    C=18

    O

    labeled samples. The spectrum devoid

    of the 13

    C=18

    O transition (full circles)

    was fitted with a Gaussian (red) that

    was subtracted from the full spectrum

    (empty circles).

  • 6

    3. 2D-IR data:

    An example of a 2D-IR spectrum showing the complete amide I region is presented in Fig. S5. In

    particular the one for 100% 13

    C=18

    O labeled G12 is shown.

    Figure S5: 2D-IR correlation spectrum of 100% 13

    C=18

    O labeled G12. The amide I main band

    transition appears at =1656 cm-1

    . At ca. =1635

    cm-1

    the amide I transition of hydrated residues can

    be seen. The transition at ca. =1613 cm-1

    is from 13

    C=16

    O that exists due to natural abundance of 13

    C

    and to incomplete conversion of G12 from 13

    C=16

    O

    to 13

    C=18

    O (ca. 5%). All the amide I transitions

    present a non diagonal peak at lower energies (ca.

    10 cm-1

    ) that is constant with waiting time (T)

    indicating that it is not produced by a chemical

    exchange phenomenon.

    The isotope labeled region of the 2DIR spectra of all the samples are presented in Fig. S6. The

    population decay time (T1) of the main band (12

    C=16

    O residues) was calculated from transient

    grating experiments. The intensity of the 13

    C=18

    O labeled peaks in the 2D-IR spectra decayed at

    the same rate with waiting time (T).

  • 7

  • 8

    Figure S6: 2D-IR correlation spectra. The isotope region of the 2D-IR correlation spectrum of (bottom) 10%,

    (middle) 20% and (top) 100% 13

    C=18

    O labeled samples. The dashed yellow line indicates the peak frequency

    assigned to 13

    C=18

    O used for the analysis.

    4. Model to extract the coupling constants:

    Neglecting the mixed mode anharmonicity the 5x5 hamiltonian is:

    j

    ji

    i

    j

    i

    H

    22000

    2200

    02200

    000

    000

    in the basis set of the local sites i,0 , 0,j , i2,0 , 0,2 j and ji, . The 0,i state represents

    a one quantum excitation in a given labeled residue in one helix and the j,0 in the other

    whereas 0,2i represents a two quanta excitation in one helix, j2,0 in the other and ji, a

    one quantum excitation in each helix. The eigenvalues and eigenvectors were calculated

    numerically for all the possible combinations of i and j. We can write the one quantum

    eigenstates as:

    0,2

    sin,02

    cos0,,0 22 jeiejcicii

    ji

    0,2

    cos,02

    sin0,,0 22 jeiejcicii

    ji

    Where ij 2tan and ie . We express the two quanta eigenstates in a symbolic

    form as:

    ijcjcicSS

    ji

    S

    j

    S

    i ,0,22,0 22

    ijcjcicSS

    ji

    S

    j

    S

    i ,0,22,0 22

  • 9

    ijcjcicAA

    ji

    A

    j

    A

    i ,0,22,0 22

    The transition dipoles of the eigenstates can be written as a function of those of the local sites

    that are the ones that are related to the real structure. We assume that the transition dipoles in the

    different helices have the same magnitude:

    1,0,0,0 ji

    The transition dipoles for the 0+ , 0- and from one quantum states to two quanta states are:

    jjii cc ,0,01,0,0 ˆˆ

    jjii cc ,0,01,0,0 ˆˆ

    jSijiSjjiSijjSiiS cccccccc ,02,021,0, ˆ2ˆ2

    jSijiSjjiSijjSiiS cccccccc ,02,021,0, ˆ2ˆ2

    jAijiAjjiAijjAiiA cccccccc ,02,021,0, ˆ2ˆ2

    jSijiSjjiSijjSiiS cccccccc ,02,021,0, ˆ2ˆ2

    jSijiSjjiSijjSiiS cccccccc ,02,021,0, ˆ2ˆ2

    jAijiAjjiAijjAiiA cccccccc ,02,021,0, ˆ2ˆ2

    The pathways involved in our experiments (19) for T=300 fs can be written in terms of the

    eigenstates in the following way:

  • 10

    The equivalent pathways starting with a coherence of the asymmetric mode also exist. Assuming

    that the experiment involves two distinct timescales the responses can be expressed as an

    inhomogeneous average of the homogeneous components. We think this is a good approximation

    as we do not identify significant changes in the spectra during the waiting time T. The

    orientational prefactors weigh each pathway with the projection of the dipole direction onto the

    incident field. Each energy combination will produce a different orientational prefactor which is

    then averaged with the inhomogeneous distribution. In the following equations the triangular

    brackets represent the inhomogeneous frequency average. Off diagonal disorder is neglected. All

    the experiments were done with xxxx polarization. Assuming that the molecules are fixed during

    the experiment and that the distribution of angles between the transition dipoles is narrow we

    write the response functions in the following way:

  • 11

    Positive peaks (indicated in red)

    0,0,,0,0

    4

    ,0

    21

    3Re

    15

    1,,,,

    1

    t

    TT

    ttii

    eTRTR

    0,0,0,0,

    4

    ,0

    54

    3Re

    15

    1,,,,

    1

    t

    TT

    ttii

    eTRTR

    0,0,,0,0,0,0

    2

    ,0

    2

    ,0

    13

    ˆˆ21Re

    15

    1,,

    1

    t

    TT

    tii

    eTR

    0,0,0,0,,0,0

    2

    ,0

    2

    ,0

    14

    ˆˆ21Re

    15

    1,,

    1

    t

    TT

    tii

    eTR

    0,0,,0,0,0,0

    2

    ,0

    2

    ,0

    21

    ˆˆ21Re

    15

    1,,

    1

    t

    TTT

    tii

    eeTR

    0,0,0,0,

    ,0,0

    2

    ,0

    2

    ,0

    25

    ˆˆ21Re

    15

    1,,

    1

    t

    TTT

    tii

    eeTR

    Negative peaks (indicated in blue):

    ,,,0,0

    2

    ,,0

    2

    ,

    2

    ,0

    3

    ˆˆ21Re

    15

    1,,

    1

    AtA

    AA

    TT

    tAii

    eTR

    ,,,0,0

    2

    ,,0

    22

    0

    3

    ˆˆ21Re

    15

    1,,

    1

    StS

    SS

    TT

    tSii

    eTR

    ,,,0,0

    2

    ,,0

    22

    0

    3

    ˆˆ21Re

    15

    1,,

    1

    StS

    SS

    TT

    tSii

    eTR

    ,,0,0,

    2

    ,,0

    2

    ,

    2

    ,0

    6

    ˆˆ21Re

    15

    1,,

    1

    AtA

    AA

    TT

    tAii

    eTR

    ,,0,0,

    2

    ,,0

    22

    0

    6

    ˆˆ21Re

    15

    1,,

    1

    StS

    SS

    TT

    tSii

    eTR

  • 12

    ,,0,0,

    2

    ,,0

    22

    0

    6

    ˆˆ21Re

    15

    1,,

    1

    StS

    SS

    TT

    tSii

    eTR

    ,,,0,0

    ,,0,,0,,0,,0,,,0,0,,,0,0

    23

    ˆˆˆˆˆˆˆˆˆˆˆˆRe

    15

    1,,

    1

    AtA

    AAAAAAAA

    TTT

    tAii

    eeTR

    ,,,0,0

    ,,0,,0,,0,,0,,,0,0,,,0,0

    23

    ˆˆˆˆˆˆˆˆˆˆˆˆRe

    15

    1,,

    1

    StS

    SSSSSSSS

    TTT

    tSii

    eeTR

    ,,,0,0

    ,,0,,0,,0,,0,,,0,0,,,0,0

    23

    ˆˆˆˆˆˆˆˆˆˆˆˆRe

    15

    1,,

    1

    StS

    SSSSSSSS

    TTT

    tSii

    eeTR

    ,,0,0,

    ,,0,,0,,0,,0,,,0,0,,,0,0

    27

    ˆˆˆˆˆˆˆˆˆˆˆˆRe

    15

    1,,

    1

    AtA

    AAAAAAAA

    TTT

    tAii

    eeTR

    ,,0,0,

    ,,0,,0,,0,,0,,,0,0,,,0,0

    27

    ˆˆˆˆˆˆˆˆˆˆˆˆRe

    15

    1,,

    1

    StS

    SSSSSSSS

    TTT

    tSii

    eeTR

    ,,0,0,

    ,,0,,0,,0,,0,,,0,0,,,0,0

    27

    ˆˆˆˆˆˆˆˆˆˆˆˆRe

    15

    1,,

    1

    StS

    SSSSSSSS

    TTT

    tSii

    eeTR

    The averaging due to the inhomogeneous distribution of energies is

    ji

    jjii

    j

    jj

    i

    ii

    eii

    eTR

    jit

    TT

    jit

    2

    2

    2

    2

    21221

    1

    20000

    2

    0

    2

    0

    1

    12

    13Re

    15

    1,,

    For the case of uncorrelated frequencies, 0 , equal distribution of energies on both sites and

    d being the diagonal trace displaced by c from the center ( dt c ):

    2

    2

    2

    2

    122

    0000

    2

    0

    2

    0

    12

    13Re

    15

    1,

    j

    jj

    i

    ii

    ecii

    eTR

    jidd

    TT

    jid

  • 13

    This same evaluation is done for the rest of the pathways. The FTIR spectrum involves one

    quantum transitions only:

    2

    2

    2

    2

    22

    00

    2

    0

    00

    2

    0

    2

    1Re

    3

    1 jjj

    i

    ii

    eii

    Sji

    jiFTIR

    It should be noted that the spectra only depend on the product of coupling constant and cosine of

    the angle between the dipoles. This fact limits the information obtained from the spectra to the

    absolute value of the coupling constant. Another detail to be observed is that, within our

    approximations, decreases with delocalization reaching a lower limit of one half for the

    completely delocalized case. The static frequency correlation component () is also distributed

    between the symmetric and asymmetric modes reaching a lower limit of 21 when the coupling

    is much greater than the inhomogeneous width.

    The width of the FTIR and the diagonal trace of the 2D-IR spectra have a quadratic dependence

    with the coupling constant. This dependence presents a higher slope in the 2D-IR diagonal trace

    which also resolves the symmetric and asymmetric transitions at lower couplings when the

    transition is mostly homogeneous. The frequency separation between symmetric and asymmetric

    peaks in dimers with disorder is larger than two times the coupling constant when the

    inhomogeneous width is comparable to the coupling. This indicates that when static disorder is

    present fitting coupled bands with two components and extracting the coupling constant from the

    separation between them can lead to overestimations.

    The difference in phase between the FTIR spectrum and the 2D-IR diagonal trace widths in 10%

    13C=

    18O labeled samples shown in Fig. 3 of the main text is attributed to the variation of

    homogeneous and inhomogeneous components. When the lineshape is dominated by the

    inhomogeneous average shown in the model, there is no significant difference between the 2D-

    IR and FTIR lineshapes, whereas if it is dominated by the homogeneous component, the 2DIR

    trace is 12 narrower than the FTIR. This makes the phases of the 2D-IR and FTIR i+4

    patterns to be different.

  • 14

    The diagonal 2D-IR trace presents significant advantages over the FTIR spectra. It is an

    inhomogeneous average of a product of Lorenztians whereas the FTIR spectrum consists of the

    average of a single Lorenztian. Therefore the 2D-IR diagonal trace has significantly better

    spectral resolution than FTIR when the homogeneous broadening is dominant. This fact provides

    the 2D-IR diagonal trace with better sensitivity to coupling. To relate to a distribution of

    structures, knowledge of the effect of structures on the frequency would be needed (32).

    5. Fit of the FTIR and 2D-IR spectra

    The model described above was used to fit the FTIR and 2D-IR diagonal traces of 100% and

    10% 13

    C=18

    O labeled samples and extract values for the absolute value of the coupling constant

    for each pair of adjacent residues. Demanding consistency between the FTIR and the 2D-IR

    spectra provides robustness to the underlying theoretical description because they present

    different functionalities of the same parameters. In particular, it constrains the homogeneous to

    inhomogeneous ratio by properly describing the relative widths of FTIR and 2D-IR spectra

    simultaneously. Also, the uncertainties in the coupling constant (see text) are increased by ca.

    20% by dropping consistency between the FTIR and 2D-IR. The results are presented in Fig S7.

  • 15

  • 16

    Figure S7: FTIR and 2D-IR diagonal trace fits. Lineshapes of (column 1) normalized FTIR 100%, (column 2)

    normalized FTIR 10%, (column 3) normalized 2D-IR 100% and (column 4) normalized 2D-IR 10% 13

    C=18

    O labeled

    samples. The experimental data is shown with full black lines and the fits to the model explained above are shown in

    blue dashed lines. Some samples (e.g. G12) exhibit background transitions that are more evident in the 10% 13

    C=18

    O

    labeled samples than in the 100% ones, because the signal to background ratio is reduced. These transitions are

    believed to be Fermi resonances to combination modes. In the cases in which any of these peaks would interfere

    with the transition under study they were fitted with a Gaussian (green) while the transition of interest was fitted

    with the model developed above (blue) resulting in the total spectrum (red).

  • 17

    6. Structure determination:

    In order to create a complete family of possible two-fold symmetric ideal helical dimers the

    procedure shown in Fig. S8 was used.

    Fig. S8: Sampling the of two-fold symmetric ideal

    helix dimers space. The axes of the two fold

    symmetric ideal helices lies along the z axis. The

    phase around the z-axis was varied between -180 and

    180 degrees in steps of 10 degrees which is equivalent

    to approximately a 0.5 Å displacement on the surface

    of the helix. The translation along the z axis leaves

    the crossing point of the helix on the xy plane. This

    displacement was sampled in a range of 20 Å in steps

    of 0.5 Å. The helices were rotated around the x axis

    by the crossing angle divided by 2 which was

    sampled between -180 and 180 degrees in steps of 5

    degrees. Finally, each helix was displaced in opposite

    directions along the x axis by half the interhelical distance which was sample from 6 Å to 10 Å in steps of 0.25 Å.

    For each structure the vibrational coupling constant was calculated assuming that the dipole

    direction of the 13

    C=18

    O substituted dipole is the same as the unperturbed 12

    C=16

    O one. Through

    perturbation theory the angle between unperturbed ( m ) and perturbed ( 'm ) transition dipoles

    can be calculated as:

    n

    mnmn

    m

    mmm

    cos

    65

    11cos

    '

    '

    where mn is the coupling in cm-1

    , mn is the angle between the m and n transition dipoles and the

    energy gap is 65 cm-1

    . For central residues in an alpha helix these angles are smaller than 3.

    These yielded a family of 1.9 x 106 structures 48 of which possessed vibrational coupling

    constants whose absolute value agreed with the experiment by a 2 analysis with 75%

    confidence. This group of structures had a crossing angle of -63±13 and interhelical distance of

    8.5±0.3 Å.

    A new module named IR has been implemented into the Xplor-NIH (25,26) program in order to

    use IR restraints in protein structure calculation. The structure was calculated by simulated

    annealing in torsion angle space using experimental IR constraints together with backbone

    dihedral angle constraints and hydrogen bonds for ideal -helical secondary structure elements.

  • 18

    The 48 preselected structures obtained by the method described above were each taken as

    starting points of simulated anneal runs for 100 structures. Every structure was first minimized

    by 500 steps to remove bad contacts, bathed at a high temperature (3000 K) for 1000 steps,

    followed by cooling to a low temperature (10 K) for another 1000 steps and subjected to Powell

    minimization for final 2000 steps. We found the force constants for the IR constraints (in cm-1

    )

    by reducing a coarse sampling (5-100 Kcal mol-1

    cm2) to a smaller range (10 to 40 kcal mol

    -

    1cm

    2), in which the structural RSMD values converged and the overall secondary structure was

    not distorted (Table 1).

    Table S1: Refinement statistics for protein structures in presence of IR constraints with different weight factors

    Weight factor 10

    kcal mol-1

    cm2

    20*

    kcal mol-1

    cm2

    40

    kcal mol-1

    cm2

    Structure statistics

    Average RMSD. to the mean structure (Å)

    Residue 10-21 0.92 ± 0.23 0.78 ± 0.25 0.85 ± 0.11

    All residues 1.28 ± 0.20 1.14 ± 0.21 1.16 ± 0.09

    Fitting

    Slope 0.95 ± 0.06 0.97 ± 0.05 0.97 ± 0.05

    R 0.968 0.973 0.971

    RMS (cm-1

    ) 0.713 0.693 0.705

    Energy (kcal)

    Total energy 35.7 ± 1.3 38.5 ± 2.5 43.6 ± 4.8

    IR energy 0.6 ± 0.4 1.1 ± 0.6 2.3 ± 1.7

    *: the one was used in the final calculation.

    Importantly, the RMSD between the mean structures computed with force constants between 10

    and 40 kcal/mol is smaller than the RMSD computed for the individual members of the ensemble

    at a given value of the force constant. This can be understood, because the structural restraints

  • 19

    maintain helical geometry and prevent repulsive interactions between the helices, while the

    experimental restraints provide the bulk of the attractive potential. The final value of the force

    constant in the simulated annealing target function for IR (20 kcal mol-1

    cm2) was chosen based

    on the smallest RMSD and the best agreement between the experimental coupling constants with

    those back calculated from the structure (Table S2 and Fig. S9).

    Table S2. Pairwise RMSD. (Å) between the mean structures obtained with different weights for IR constraints.

    10

    kcal mol-1

    cm2

    20*

    kcal mol-1

    cm2

    40

    kcal mol-1

    cm2

    10

    kcal mol-1

    cm2

    0 0.20 0.38

    20*

    kcal mol-1

    cm2

    0 0.42

    40

    kcal mol-1

    cm2

    0

    Figure S9: Pairwise backbone RMSD, calculated for 20

    best structures by superimposing residues 10-21. Magenta,

    red and purple traces represent values with an IR-restrained

    force constant of 10, 20 and 40 kcal mol-1

    cm2, respectively.

    Although the convergence is best for N- and C-terminus

    with an IR-restrained force constant of 40 kcal mol-1

    cm2,

    the central part where the IR constraints are actually

    applied has the best converge for 20 kcal mol-1

    cm2.

    The final values for the IR constraint is 20 kcal mol-1

    cm2; the rest of the restraints were standard

    (50 kcal mol-1 Å-2

    for hydrogen bond restraints; 5 kcal mol-1

    rad-2

    for dihedral angle restraints; 4

    kcal mol-1 Å-4

    for the quartic van der Waals repulsion term). Of the 4800 obtained structures, the

    20 lowest energy structures form an ensemble to represent the structure. The distribution of

  • 20

    crossing angles and interhelical distances and the RMSD. per residue is presented in Fig. S10.

    Figure S10: Simulated

    annealing results: (left)

    distribution of crossing

    angles and interhelical

    distances for the 20

    structures yielded by the

    constrained simulated

    annealing. The red dot

    belongs to the structure of

    lowest energy. (right) RMSD

    per residue for the 20 yielded

    structures.

  • References and Notes 1. P. Hamm, M. Lim, R. M. Hochstrasser, Structure of the amide I band of peptides measured by

    femtosecond nonlinear-infrared spectroscopy. J. Phys. Chem. B 102, 6123 (1998). doi:10.1021/jp9813286

    2. S. Sul, D. Karaiskaj, Y. Jiang, N.-H. Ge, Conformations of N-acetyl-L-prolinamide by two-dimensional infrared spectroscopy. J. Phys. Chem. B 110, 19891 (2006). doi:10.1021/jp062039h Medline

    3. I. V. Rubtsov, R. M. Hochstrasser, Vibrational dynamics, mode coupling, and structural constraints for acetylproline-NH2. J. Phys. Chem. B 106, 9165 (2002). doi:10.1021/jp020837b

    4. P. Hamm, M. Lim, W. F. DeGrado, R. M. Hochstrasser, The two-dimensional IR nonlinear spectroscopy of a cyclic penta-peptide in relation to its three-dimensional structure. Proc. Natl. Acad. Sci. U.S.A. 96, 2036 (1999). doi:10.1073/pnas.96.5.2036 Medline

    5. S. Woutersen, P. Hamm, Structure determination of trialanine in water using polarization sensitive two-dimensional vibrational spectroscopy. J. Phys. Chem. B 104, 11316 (2000). doi:10.1021/jp001546a

    6. Y. S. Kim, J. Wang, R. M. Hochstrasser, Two-dimensional infrared spectroscopy of the alanine dipeptide in aqueous solution. J. Phys. Chem. B 109, 7511 (2005). doi:10.1021/jp044989d Medline

    7. R. R. Ernst, G. Bodenhausen, A. Wokaun, Principles of Nuclear Magnetic Resonance in One and Two Dimensions (Oxford Univ. Press, Oxford, 1987).

    8. R. Li et al., Dimerization of the transmembrane domain of integrin αIIb subunit in cell membranes. J. Biol. Chem. 279, 26666 (2004). doi:10.1074/jbc.M314168200 Medline

    9. Y. S. Kim, L. Liu, P. H. Axelsen, R. M. Hochstrasser, 2D IR provides evidence for mobile water molecules in beta-amyloid fibrils. Proc. Natl. Acad. Sci. U.S.A. 106, 17751 (2009). doi:10.1073/pnas.0909888106 Medline

    10. A. Ghosh, J. Qiu, W. F. Degrado, R. M. Hochstrasser, Proc. Natl. Acad. Sci. U.S.A. 108, 6115 (2011).

    11. W. Zhuang, T. Hayashi, S. Mukamel, Coherent multidimensional vibrational spectroscopy of biomolecules: Concepts, simulations, and challenges. Angew. Chem. Int. Ed. Engl. 48, 3750 (2009). doi:10.1002/anie.200802644 Medline

    12. S. Krimm, J. Bandekar, Vibrational spectroscopy and conformation of peptides, polypeptides, and proteins. Adv. Protein Chem. 38, 181 (1986). doi:10.1016/S0065-3233(08)60528-8 Medline

    13. C. Fang, A. Senes, L. Cristian, W. F. DeGrado, R. M. Hochstrasser, Amide vibrations are delocalized across the hydrophobic interface of a transmembrane helix dimer. Proc. Natl. Acad. Sci. U.S.A. 103, 16740 (2006). doi:10.1073/pnas.0608243103 Medline

    14. K. R. MacKenzie, J. H. Prestegard, D. M. Engelman, A transmembrane helix dimer: structure and implications. Science 276, 131 (1997). doi:10.1126/science.276.5309.131 Medline

    http://dx.doi.org/10.1021/jp9813286http://dx.doi.org/10.1021/jp062039hhttp://www.ncbi.nlm.nih.gov/entrez/query.fcgi?cmd=Retrieve&db=PubMed&list_uids=17020375&dopt=Abstracthttp://dx.doi.org/10.1021/jp020837bhttp://dx.doi.org/10.1073/pnas.96.5.2036http://www.ncbi.nlm.nih.gov/entrez/query.fcgi?cmd=Retrieve&db=PubMed&list_uids=10051590&dopt=Abstracthttp://dx.doi.org/10.1021/jp001546ahttp://dx.doi.org/10.1021/jp044989dhttp://www.ncbi.nlm.nih.gov/entrez/query.fcgi?cmd=Retrieve&db=PubMed&list_uids=16851862&dopt=Abstracthttp://dx.doi.org/10.1074/jbc.M314168200http://www.ncbi.nlm.nih.gov/entrez/query.fcgi?cmd=Retrieve&db=PubMed&list_uids=15067009&dopt=Abstracthttp://dx.doi.org/10.1073/pnas.0909888106http://www.ncbi.nlm.nih.gov/entrez/query.fcgi?cmd=Retrieve&db=PubMed&list_uids=19815514&dopt=Abstracthttp://dx.doi.org/10.1002/anie.200802644http://www.ncbi.nlm.nih.gov/entrez/query.fcgi?cmd=Retrieve&db=PubMed&list_uids=19415637&dopt=Abstracthttp://dx.doi.org/10.1016/S0065-3233(08)60528-8http://dx.doi.org/10.1016/S0065-3233(08)60528-8http://www.ncbi.nlm.nih.gov/entrez/query.fcgi?cmd=Retrieve&db=PubMed&list_uids=3541539&dopt=Abstracthttp://dx.doi.org/10.1073/pnas.0608243103http://www.ncbi.nlm.nih.gov/entrez/query.fcgi?cmd=Retrieve&db=PubMed&list_uids=17075037&dopt=Abstracthttp://dx.doi.org/10.1126/science.276.5309.131http://www.ncbi.nlm.nih.gov/entrez/query.fcgi?cmd=Retrieve&db=PubMed&list_uids=9082985&dopt=Abstract

  • 15. J. Marecek et al., A simple and economical method for the production of 13C,18O-labeled Fmoc-amino acids with high levels of enrichment: Applications to isotope-edited IR studies of proteins. Org. Lett. 9, 4935 (2007). doi:10.1021/ol701913p Medline

    16. Single-letter abbreviations for the amino acid residues are as follows: G, Gly; L, Leu; V, Val.

    17. K. Kwac, M. Cho, Molecular dynamics simulation study of N-methylacetamide in water. II. Two-dimensional infrared pump-probe spectra. J. Chem. Phys. 119, 2256 (2003). doi:10.1063/1.1580808

    18. S. Mukamel, Principles of Nonlinear Optical Spectroscopy (Oxford Univ. Press, New York, 1995).

    19. J. Wang, R. M. Hochstrasser, Characteristics of the two-dimensional infrared spectroscopy of helices from approximate simulations and analytic models. Chem. Phys. 297, 195 (2004). doi:10.1016/j.chemphys.2003.10.013

    20. R. Wertheimer, R. Silbey, On excitation transfer and relaxation models in low-temperature systems. Chem. Phys. Lett. 75, 243 (1980). doi:10.1016/0009-2614(80)80505-7

    21. P. Mukherjee, I. Kass, I. T. Arkin, M. T. Zanni, Picosecond dynamics of a membrane protein revealed by 2D IR. Proc. Natl. Acad. Sci. U.S.A. 103, 3528 (2006). doi:10.1073/pnas.0508833103 Medline

    22. A. M. Woys et al., 2D IR line shapes probe ovispirin peptide conformation and depth in lipid bilayers. J. Am. Chem. Soc. 132, 2832 (2010). doi:10.1021/ja9101776 Medline

    23. C. Fang, Two-dimensional infrared measurements of the coupling between amide modes of an α-helix. Chem. Phys. Lett. 382, 586 (2003). doi:10.1016/j.cplett.2003.10.111

    24. H. Torii, M. Tasumi, Model calculations on the amide-I infrared bands of globular proteins. J. Chem. Phys. 96, 3379 (1992). doi:10.1063/1.461939

    25. C. Schwieters, J. Kuszewski, G. Mariusclore, Using Xplor-NIH for NMR molecular structure determination. Prog. Nucl. Magn. Reson. Spectrosc. 48, 47 (2006). doi:10.1016/j.pnmrs.2005.10.001

    26. C. D. Schwieters, J. J. Kuszewski, N. Tjandra, G. M. Clore, The Xplor-NIH NMR molecular structure determination package. J. Magn. Reson. 160, 65 (2003). doi:10.1016/S1090-7807(02)00014-9 Medline

    27. B. W. Berger et al., Consensus motif for integrin transmembrane helix association. Proc. Natl. Acad. Sci. U.S.A. 107, 703 (2010). doi:10.1073/pnas.0910873107 Medline

    28. R. F. S. Walters, W. F. DeGrado, Helix-packing motifs in membrane proteins. Proc. Natl. Acad. Sci. U.S.A. 103, 13658 (2006). doi:10.1073/pnas.0605878103 Medline

    29. C. Kolano, J. Helbing, M. Kozinski, W. Sander, P. Hamm, Watching hydrogen-bond dynamics in a β-turn by transient two-dimensional infrared spectroscopy. Nature 444, 469 (2006). doi:10.1038/nature05352 Medline

    30. H. Durchschlag, P. Zipper, Calculation of the partial volume of organic compounds and polymers. Prog. Colloid Polym. Sci. 94, 20 (1994). doi:10.1007/BFb0115599

    http://dx.doi.org/10.1021/ol701913phttp://www.ncbi.nlm.nih.gov/entrez/query.fcgi?cmd=Retrieve&db=PubMed&list_uids=17958432&dopt=Abstracthttp://dx.doi.org/10.1063/1.1580808http://dx.doi.org/10.1016/j.chemphys.2003.10.013http://dx.doi.org/10.1016/0009-2614(80)80505-7http://dx.doi.org/10.1073/pnas.0508833103http://www.ncbi.nlm.nih.gov/entrez/query.fcgi?cmd=Retrieve&db=PubMed&list_uids=16505377&dopt=Abstracthttp://dx.doi.org/10.1021/ja9101776http://www.ncbi.nlm.nih.gov/entrez/query.fcgi?cmd=Retrieve&db=PubMed&list_uids=20136132&dopt=Abstracthttp://dx.doi.org/10.1016/j.cplett.2003.10.111http://dx.doi.org/10.1063/1.461939http://dx.doi.org/10.1016/j.pnmrs.2005.10.001http://dx.doi.org/10.1016/S1090-7807(02)00014-9http://dx.doi.org/10.1016/S1090-7807(02)00014-9http://www.ncbi.nlm.nih.gov/entrez/query.fcgi?cmd=Retrieve&db=PubMed&list_uids=12565051&dopt=Abstracthttp://dx.doi.org/10.1073/pnas.0910873107http://www.ncbi.nlm.nih.gov/entrez/query.fcgi?cmd=Retrieve&db=PubMed&list_uids=20080739&dopt=Abstracthttp://dx.doi.org/10.1073/pnas.0605878103http://www.ncbi.nlm.nih.gov/entrez/query.fcgi?cmd=Retrieve&db=PubMed&list_uids=16954199&dopt=Abstracthttp://dx.doi.org/10.1038/nature05352http://www.ncbi.nlm.nih.gov/entrez/query.fcgi?cmd=Retrieve&db=PubMed&list_uids=17122853&dopt=Abstracthttp://dx.doi.org/10.1007/BFb0115599

  • 31. S. T. R. Walsh et al., The hydration of amides in helices; a comprehensive picture from molecular dynamics, IR, and NMR. Protein Sci. 12, 520 (2003). doi:10.1110/ps.0223003 Medline

    3. J. R. Schmidt, S. A. Corcelli, J. L. Skinner, Ultrafast vibrational spectroscopy of water and aqueous N-methylacetamide: Comparison of different electronic structure/molecular dynamics approaches. J. Chem. Phys. 121, 8887 (2004). doi:10.1063/1.1791632 Medline

    http://dx.doi.org/10.1110/ps.0223003http://www.ncbi.nlm.nih.gov/entrez/query.fcgi?cmd=Retrieve&db=PubMed&list_uids=12592022&dopt=Abstracthttp://www.ncbi.nlm.nih.gov/entrez/query.fcgi?cmd=Retrieve&db=PubMed&list_uids=12592022&dopt=Abstracthttp://dx.doi.org/10.1063/1.1791632http://www.ncbi.nlm.nih.gov/entrez/query.fcgi?cmd=Retrieve&db=PubMed&list_uids=15527353&dopt=Abstract