You are on page 1of 6

Investigation of protein conformation and interactions

with salts via X-ray absorption spectroscopy


Craig P. Schwartza,b, Janel S. Uejioa,b, Andrew M. Duffina,b, Alice H. Englanda,b, Daniel N. Kellya,
David Prendergastc, and Richard J. Saykallya,b,1
a
Department of Chemistry, University of California, Berkeley, CA 94720; bChemical Sciences Division, Lawrence Berkeley National Laboratory, Berkeley,
CA 94720; and cMolecular Foundry, Lawrence Berkeley National Laboratory, Berkeley, CA 94720

Edited by Michael L. Klein, Temple University, Philadelphia, PA, and approved June 30, 2010 (received for review May 11, 2010)

Nitrogen K-edge spectra of aqueous triglycine were measured using It has also been proposed that absolute free energies of hydra-
liquid microjets, and the effects of Hofmeister-active salts on the tion also underlie Hofmeister effects (8, 9), contending that if
spectra were observed. Spectra simulated using density functional absolute hydration-free energies of anionic and cationic groups
theory, sampled from room temperature classical molecular match, then ion pairing will occur, signifying strong ion-protein
dynamics trajectories, capture all major features in the measured interactions (10). Cremer and coworkers have performed a series
spectra. The spectrum of triglycine in water is quite similar to that of studies wherein protein analogues were combined with various
in the presence of chaotropic sodium bromide (and other halides), amounts of salt, causing the lower critical solution temperature
which raises the solubility of proteins. However, a new feature is (LCST)—the critical temperature below which the mixture is
found when kosmotropic Na2 SO3 , which lowers solubility, is miscible in all portions—to change as a function of salt concen-
present; this feature results from excitations of the nitrogen atom tration and identity (3, 11–13). They observed that kosmotropes
in the terminal amino group of triglycine. Both direct interactions lower the LCST more than do chaotropes. Based on fitting of
between this salt and the protonated amino terminus, as well as cor- their measurements combined with thermodynamic arguments,
responding changes in the conformational dynamics of the system, they argued that the way the salts affect the LCST is dependent
contribute to this new feature. These molecular measurements upon the Hofmeister activity of the anion. Chaotropes (such as
support a different mechanism for the Hofmeister effect than has Br− ) were believed to directly interact with the protein analogue
previously been suggested based on thermodynamic measure- near the nitrogen sites, as well as to interact with the exposed
ments. It is also shown that near edge X-ray absorption fine struc- hydrophobic surfaces. Kosmotropes, on the other hand, were
ture (NEXAFS) is sensitive to strong direct interaction between believed to polarize the water interacting with the nitrogen-
certain salts and charged peptides. However, by investigating the and oxygen-containing groups as well as to interact with hydropho-
sensitivity of NEXAFS to the extreme structural differences between bic groups. However, these conclusions were based on light scat-
model β-sheets and α-helices, we conclude that this technique is re- tering measurements and thermodynamic data, so the interactions
latively insensitive to secondary structure of peptides and proteins. had to be inferred, rather than being observed directly. Herein,
we show evidence that suggests kosmotropes directly interact with
X-ray absorption near edge structure ∣ molecular dynamics ∣ the nitrogen backbone of peptides based on a molecular probe,
density functional theory ∣ eXcited electron and Core Hole
casting doubt on the proposed mechanism behind the thermody-
namic data.

O ver 100 years ago Hofmeister discovered that adding salts to


egg white protein could alter its solubility (1). Different con-
centrations of salt would cause the protein to precipitate out of
Here, we investigate triglycine as a model peptide via near
edge X-ray absorption fine structure (NEXAFS) experiments
and theoretical methods, seeking to further characterize the
solution with the amount specifically related to the ion identities. interactions of proteins with selected salts. Nitrogen K-edge (ex-
This ordering of salt-protein interactions has since become citation from the N 1s orbital) spectra of triglycine in aqueous
known as the “Hofmeister series” (2), and the associated “Hof- solution were measured, without cosolutes, with chaotropic coso-
meister effects” now extend to many different phenomena, lute Na2 SO3 , and with cosolute NaBr, seeking to quantify
including protein denaturing, optical rotation of sugars, and observed differences in the spectra of triglycine due to the
bacterial growth rates (3). It is known that anions have a stronger presence of the salts. NEXAFS probes unoccupied antibonding
effect on protein solubility than do cations and the general anion and Rydberg states, which are highly sensitive to intermolecular
ordering is interactions (14–16). This follows similar recent studies using
X-ray spectroscopy to characterize interactions between ions
CO3 2− > SO4 2− > SO3 2− > Cl− > Br− > I− > SCN− : and small biomolecules (16, 17).
There are now several methods available for obtaining
The species on the left are generally referred to as “kosmo- core-level spectra of liquids; we use small diameter jets of water
tropes,” or “structure makers,” and the species on the right solutions (∼30 microns) with windowless coupling to a synchro-
are generally referred to as “chaotropes,” or “structure breakers” tron beamline. This approach is particularly useful for studying
(3, 4). Originally, these terms derived from the entropies of complex molecules, because the high velocity and continuous
hydration and referred to making and breaking the water struc- renewal of the liquid jet minimizes spectral contributions due
ture, which was believed to be the cause of the Hofmeister effects. to sample damage (18). The chemical information that can be
However, this model of ascribing Hofmeister effects to bulk water extracted from such measurements is limited in many respects
changes has come under criticism, given more recent experimen- by the accuracy of theoretical methods for computing core-level
tal findings (3). It is now believed that interactions either directly
between a protein and the ions or interactions mediated by a
single solvation shell underlie the effects (3, 5). Some groups have Author contributions: R.J.S. designed research; C.P.S., J.S.U., A.M.D., A.H.E., and D.N.K.
recently presented measurements to support the original conten- performed research; C.P.S. and D.P. analyzed data; and C.P.S. wrote the paper.

tion that Hofmeister effects resulted from changes in the bulk The authors declare no conflict of interest.
water structure, but their models treated water as a two-state This article is a PNAS Direct Submission.
system, which has come under intense recent criticism (6, 7). 1
To whom correspondence should be addressed. E-mail: saykally@berkeley.edu.

14008–14013 ∣ PNAS ∣ August 10, 2010 ∣ vol. 107 ∣ no. 32 www.pnas.org/cgi/doi/10.1073/pnas.1006435107


spectra. Here, we use the recently developed eXcited electron alanine in an α-helix (Fig. 1D) and β-sheet (Fig. 1E). At neutral
and Core Hole (XCH) method, which employs density functional pH, triglycine will exist primarily in the zwitterionic structure
theory (DFT) to accurately calculate NEXAFS spectra (19, 20). shown. The nitrogen atoms are labeled 1–3; this scheme will
We simulated the atomic configurations of triglycine with and be used throughout this work to reference specific nitrogen atoms
without additional salts in aqueous solution at 300 K using within triglycine, with N1 referring to the nitrogen closest to the
classical molecular dynamics. Sampling a trajectory in our DFT carboxylic group, N2 the middle nitrogen, and N3 the terminal
simulation captures the impact of nuclear motion on the calcu- nitrogen. In the case of polyalanine in both the antiparallel
lated spectrum (21). This technique has been used previously to β-sheet and the α-helix, every monomer of alanine is conforma-
interpret the spectra of solvated amino acids and the prototypical tionally identical due to the coordinates used.
aromatic molecule, pyrrole in aqueous solutions (22, 23). To
explore possible spectral signatures of secondary structure, we Experimental Results and DFT Calculations. The experimental and
also used rigid structural models of polyalanine in both an theoretical results for the nitrogen K-edge spectra of solvated
antiparallel β-sheet and an α-helix using structural parameters triglycine are shown in Fig. 2. The experimental spectrum consists
from the literature (24) to calculate the respective N-edge of a large feature at approximately 401 eV and another large
X-ray absorption spectra. broad feature centered at approximately 406 eV, followed by a
There have been many studies of proteins and amino acids in large decay with another feature possibly located at 412 eV.
solid form using NEXAFS; every naturally occurring individual Our simulations provide excellent agreement, capturing the es-
amino acid has had its core-level spectra characterized at the C, sential features of the experimental spectra. Furthermore, the
N, and O K edges (25), and a variety of polypeptides and protein character of transitions can easily be assigned by examining
spectra have been measured under similar conditions (26–28). the spectra of the individual nitrogen atoms. The first peak is
These data have been used to predict (using a building-block due entirely to N1 and N2. That N1 and N2 have similar spectra
approach) the corresponding spectrum of a protein based on its is not surprising, given that the local environment surrounding
amino acid composition with reasonable success (28). Based on each amide nitrogen in the triglycine backbone should be quite
such studies, it has been suggested that the difference in π  -reso- similar. That N1 and N2 are similar is also expected based on the
“building-block” approach that is used throughout much of the

CHEMISTRY
nances between a model dipeptide and a protein could be due to
spectral differences between an α-helix and a planar structure (26, NEXAFS literature (27, 29). N3 largely gives rise to the broad
27). In this work, we show preliminary evidence that the N K-edge feature around 406 eV. It is not surprising that there are not
spectrum will be relatively unchanged in these conformations. many sharp features, as this is a large and quite flexible molecule
with freedom to sample molecular phase space, and the spectral
Results and Discussion features shift considerably in both energy and intensity with pep-
Molecular Structures. The structure of triglycine is shown in Fig. 1A, tide motions.
along with that of the sulfite anion (Fig. 1B) and polyalanine When sodium bromide is added to the solution of triglycine,
(Fig. 1C), along with the structure used in calculations of poly- the spectrum does not change significantly, as shown in Fig. 3.
The spectrum again consists of a strong feature at approximately
401 eV, followed by another strong feature centered at 406 eVand
a possible weak feature at 412 eV. The experimental result again
is closely modeled by the simulations. We note that the spectra of
N1 and N2 at 401 eV remain similar despite the interactions with
bromide. Sodium is generally repelled from the protonated nitro-
gens and therefore is not of much spectral importance, based on
classical molecular dynamics (MD), including our own simula-

Fig. 1. Schematic drawing of triglycine (A), the sulfite anion (B), and poly-
alanine (C). The structure of the α-helix and antiparallel β-sheet are shown in
D and E, respectively, with carbon in yellow, hydrogen in blue, oxygen in red,
and nitrogen in bluish gray. The unit cell used in the calculations is shown.
Triglycine is in its zwitterionic form at the pH of this study. The nitrogen Fig. 2. Experimental nitrogen K-edge NEXAFS spectrum of solvated
atoms of triglycine are labeled; this numbering scheme will be used through- triglycine (black, solid line) and the overall calculated spectrum for triglycine
out this paper. Note the two leftmost oxygen atoms on triglycine are degen- (purple, ••••). The calculated spectrum due to individual nitrogen atoms N1
erate, as are all the oxygens of the sulfite anion and every alanine monomer. (blue, — ••• —), N2 (red, – – –), and N3 (green, — • —) are shown.

Schwartz et al. PNAS ∣ August 10, 2010 ∣ vol. 107 ∣ no. 32 ∣ 14009
Molecular Dynamics and Spectroscopic Analysis. In order to extract
details of solvation structure, radial distribution functions (RDFs)
comparing the nitrogen atoms and ion proximity were calculated.
These are shown for sodium bromide and triglycine with respect
to bromide–nitrogen distances in Fig. 5A and for sodium sulfite
and triglycine with respect to oxygen (sulfite)–nitrogen distances
in Fig. 5B. Neither for sodium bromide nor sodium sulfite does
the sodium atom have a propensity for directly interacting
with the triglycine nitrogen atoms; generally it is repelled from
the first solvation shell. In contrast, sodium strongly interacts with
the sulfite, which, in turn, strongly interacts with the protonated
nitrogen terminus, leading to a large density of sodium as a second
nearest neighbor.
We first discuss Fig. 5A. The only nitrogen that is strongly
associated with bromide ions is N3, the protonated terminus
of the zwitterion. However, this interaction produces only a mod-
est spectral consequence, comprising a small increase in intensity
at approximately 403 eV and a small decrease in intensity at
approximately 406 eV, as compared to trigylcine without the
sodium bromide present.
The RDF of sodium sulfite (oxygen) with the nitrogen atoms of
triglycine (nitrogen) is shown in Fig. 5B. It is important to note
Fig. 3. Experimental nitrogen K-edge NEXAFS spectrum of solvated trigly-
that the sulfite ion is almost an order of magnitude more likely to
cine with sodium bromide (black, solid line) and the overall calculated be located directly adjacent to hydrogen atoms bonded to any of
spectrum for triglycine with sodium bromide (purple, ••••). The calculated the nitrogen atoms than is bromide based on the radial distribu-
spectrum due to individual nitrogen atoms N1 (blue, — ••• —), N2 (red, tion functions, primarily due to a higher charge. Furthermore,
– – –), and N3 (green, — • —) are shown. sodium sulfite often exhibits a high degree of ion pairing, parti-
cularly for the sulfite near N3 and to a lesser extent that near
tions (30). This spectrum is generally very similar to that of N1. These large differences in local environments around the
triglycine without sodium bromide, in that no new features are nitrogen atoms, in particular with regards to N1 and N2, lead
present. Further NEXAFS measurements show a similar result, to the overall spectral transition being slightly broader, as the
namely, no new features or effects with all the other sodium
halide salts.
The addition of sodium sulfite to the solution of triglycine has a
larger spectral effect than does sodium bromide. Whereas the
spectrum shown in Fig. 4 still exhibits the same features, centered
at 401 eV, 406 eV, and 412 eV, a new feature arises at approximately
403 eV. Although some absorption intensity is contributed from
all nitrogen atoms at 406 eV, the new feature is due primarily
to the addition of sodium sulfite and its affect on the amino group
N3. We note that a new feature was also found at approximately
403 eV in our measurements in the case of potassium sulfate.

Fig. 5. Radial distribution functions from simulations of (A) triglycine with


Fig. 4. Experimental nitrogen K-edge NEXAFS spectrum of solvated trigly- sodium bromide and (B) sodium sulfite with triglycine. The radial distribution
cine with sodium sulfite (black, solid line) and the overall calculated spectrum functions shown are between the nitrogen atoms of triglycine and (A) Br− as
for triglycine with sodium sulfite (purple, ••••). The calculated spectrum due well as the radial distribution function between the nitrogen atoms of trigly-
to individual nitrogen atoms N1 (blue, — ••• —), N2 (red, – – –), and N3 cine and (B) oxygen (from SO3 2− ). N1, N2, and N3 are shown in blue (solid
(green, — • —) are shown. line), red (••••), and green (– – –), respectively, for both A and B.

14010 ∣ www.pnas.org/cgi/doi/10.1073/pnas.1006435107 Schwartz et al.


1s → lowest unoccupied molecular orbital (LUMO) transition bones lie roughly in a plane; i.e., they are not twisted. However,
is no longer energetically degenerate for N1 and N2. The 1s → the sodium sulfite solution exhibits a different conformation of
LUMO transition shifts in energy as the local environment shifts triglycine than when sodium bromide is present. Furthermore,
around the N2 atom; this is particularly related to the position of the rms deviation increases when triglycine is simulated with
the sulfite and the N-H bond length. We note that there are other either no salt present or with sodium bromide, as compared to
factors that contribute, but a single cause cannot be isolated at sodium sulfite; this implies the existence of less conformational
the present time. The new peak at 403 eV appears to be caused variations in the presence of sodium sulfite. This is in agreement
primarily by the N3 group interacting with the sulfite ion. A large with the traditional interpretation of Hofmeister effects, namely,
intensity at that energy is not found in all of the calculated spec- that kosmotropic solvent molecules such as SO3 2− will make it
tra, but only for a fraction of the calculated snapshots. However, harder for a protein to bend to form a cavity (the protein becomes
many snapshots have similar 1s → LUMO transition energies less conformationally flexible) (3). The chaotropic sodium bro-
(i.e., the absolute energy of many transitions is approximately mide induces a very slight increase in the conformational flexibil-
403 eV). This implies a similar local environment for a variety ity of triglycine, as compared to the peptide without cosolute, in
of snapshots, indicating that sulfite induces a certain amount agreement with the Hofmeister series for solvent cavity formation
of local structure in or around triglycine. If the environment (3). Although this overall protein conformational flexibility may
was shifting constantly, one would not expect this peak to appear. be important for phenomena related to the Hofmeister series,
Although the RDFs characterize the distance between trigly- NEXAFS appears to have limited sensitivity to this effect, based
cine and ions, they provide no dynamical information. In order to on our calculations. Conformational flexibility can still be impor-
determine structural variations, the rms fitted backbone of trigly- tant with regard to peak width, however.
cine from our simulations is shown in Fig. 6, with the standard The reason for this limited spectroscopic sensitivity can be
deviation in the rms fitting corresponding to the spheroid size. elucidated by examining the states probed in the experiment.
Example states are shown for the LUMO þ 1 in Fig. 7. As seen
It is important to note that for all of these structures, the back-
in this image, the state is fairly well localized on a given segment
of the polypeptide chain. When N1 is excited, the excitation is not

CHEMISTRY
delocalized very far along the backbone, although it does reach
the carbon of the carboxylic group. When N2 is excited, the case
that would be most representative of the middle of a large pro-
tein, the state is almost entirely localized on N2 and atoms at
most two bonds away. There is no significant off-chain mixing.
The state is extremely localized. The N3 excited state mixes more
than any other nitrogen, with the LUMO þ 1 delocalized all
the way back up the chain of trigylcine to N2. It also mixes with
neighboring water molecules, and for that reason, N3 is likely to
be much more sensitive to its surroundings than are the other
nitrogen molecules.
The applicability of our findings to the general elucidation of
Hofmeister effects appears to be indirect, in particular, with re-
gard to the restriction of conformational flexibility, as the kosmo-
tropic sulfite ion is largely interacting with the protonated
(zwitterionic) nitrogen group, wherein the anion pairs with so-
dium ions, causing the carboxylic acid group to loop around
and interact with the sodium. In other words, the present confor-
mational findings are probably specific to short oligopeptides;
however, one could envision such salts affecting proteins when
multiple charged groups are in close spatial proximity, as will of-
ten be the case in actual proteins. We note that it was recently
predicted that iodide and bromide ions would not directly inter-
act with the protein backbone, but rather would instead interact
with hydrophobic regions (30); we have here experimentally and
theoretically confirmed that bromide does not strongly interact
with a hydrophilic polypeptide. The calculations in the previous
work indicate that the destabilizing effect of bromide in the Hof-

Fig. 6. Result of rms fitting the backbone of triglycine to the average


structure for a given MD simulation. The carbon atoms are in blue, and Fig. 7. Isosurfaces (orange and green) are shown for the LUMO þ 1 of the
the nitrogen atoms are in purple. The results correspond to the simulations excited state of triglycine solvated with sodium sulfite and water, corre-
of (A) triglycine, (B) triglycine and sodium bromide, and (C) triglycine and sponding to 10% of the total integrated value. Many of the waters have been
sodium sulfite. The standard deviation of the fit for each atom corresponds removed for clarity. The images show an excitation on N1, N2, and N3 from
to the size of each atom. Note how similar A and B are as compared to C and left to right. Despite being a delocalized state, the state does not extend
the often larger size of the spheroids in A and B. throughout the entire molecule.

Schwartz et al. PNAS ∣ August 10, 2010 ∣ vol. 107 ∣ no. 32 ∣ 14011
meister effect is due to attraction to hydrophobic regions of the due solely to protein secondary structure, and other effects, such
protein. as bond lengths and angles, may play a larger role. We note that
Previous studies have addressed the thermodynamics of the this spectral insensitivity should be expected given the lack of mix-
Hofmeister series, but molecular level insights have been much ing present when N2 is excited, the nitrogen most representative
harder to gain. Cremer and coworkers have identified systematic of bulk protein structure. In summary, NEXAFS is, as known,
thermodynamic dependencies of the LCST based on specific largely a local probe of chemical environments.
anion identity (3, 12, 13). Their conclusion—that the more kos- Finally, if one were to consider that vibrational effects will
motropic anions would interact with the protein via mediating likely broaden the observed spectral features, it is reasonable
waters and that chaotropic anions would interact with a protein to assume that the differences between the α-helix and the
directly—is refuted by the results presented here. In order for β-sheet will be further suppressed; even the minor peak at
their conclusion to be correct, the interaction would have to ∼403 eV not present in the α-helix will likely be obscured as a
not affect the Hofmeister series in any way, a seemingly unlikely, consequence of motions. It strongly indicates that although subtle
though possible, outcome. However, it is important to note that changes in NEXAFS spectra can result from differences in con-
their measurements were taken at salt concentrations that were formation, the overall spectral sensitivity to conformation will be
significantly higher than those studied here, and it is possible that minor. However, if one could orient the proteins (for example,
different effects occur as concentration approaches the solubility relative to a surface), the NEXAFS spectrum should then exhibit
limit. Although this therefore does not prove that the mechanism a large linear dichroism effect in the case of the α-helix that would
proposed on thermodynamic arguments by Cremer and cowor- not be found with a β-sheet and perhaps this effect could be used
kers is incorrect, it should emphasize the difficulty of trying to to identify the protein conformational state. Nevertheless, pro-
infer molecular phenomena from measurements that are inher- tein conformational information will be quite difficult to extract
ently macroscopic, and should cast serious doubt on their mole- from bulk nitrogen K-edge spectra.
cular level interpretation.
Methods
Polyalanine. Finally, we have investigated the case of “extreme” Samples. Gly-Gly-Gly (Triglycine), NaBr, and Na2 SO3 were obtained commer-
cially from Sigma-Aldrich, with stated purities of at least 98%. All water used
conformational differences in polyalanine. As shown in Fig. 8,
had a resistivity of 18 MΩ∕cm. Samples were used without further purifica-
we have compared the calculated nitrogen K-edge spectrum of tion. The aqueous samples used in the liquid jets contained concentrations
polyalanine in an antiparallel β-sheet and in an α-helix (as shown of triglycine that were 33.3 g∕L of water (∼0.13 M). One of the samples
in Fig. 1) for the solid. In these frozen conformations, spectral comprised solely triglycine, another contained triglycine and 33.3 g∕L
differences will be magnified, as there will be no motions to blur (∼0.33 M) of NaBr, and the third contained triglycine and 33.3 g∕L of
spectral features—in essence, this represents a maximum possible Na2 SO3 (∼0.26 M). These correspond to approximately a 2.5∶1 NaBr∶
conformational effect on the core-level spectrum. triglycine ratio and a 2∶1 Na2 SO3 ∶triglycine ratio. We note that these
Nevertheless, the spectra are quite similar between the anti- concentrations are well below those wherein precipitation would occur in
parallel β-sheet and the α-helix. Both feature a strong absorption the liquid jets, which would lead to jet clogging.
band at ∼401 eV and peaks roughly corresponding to experimen-
NEXAFS Spectroscopy of Solvated Triglycine. Total electron yield (TEY) NEXAFS
tal scattering peaks at ∼408 eV and ∼413 eV as well as a strong
were recorded at the nitrogen K edge (∼400 eV). These measurements were
transition at ∼404 eV. The antiparallel β-sheet features an performed at Beamline 8.0.1 of the Advanced Light Source (ALS) at Lawrence
absorption at ∼403 eV that is not present in the α-helix. This Berkeley National Laboratory. A detailed description of the experimental sys-
can be compared with several experimental spectra from the tem has been published previously (18). Briefly, an intense (>1011 photons∕s),
literature, polyisoleucine (a polypeptide), lysozyme (a protein), high resolution (E∕ΔE > 4;000) tunable X-ray beam is generated from an
and 2,5-diketopiperazine (a simple cyclic glycine dimer) (26, 27). undulator. The synchrotron light is then focused (∼50 μm spot size) onto a
The lack of absorption at ∼404 eV for 2,5-diketopiperazine was small liquid jet (∼30 μm diameter), and the TEY is collected to obtain spectra
thought to be due to 2,5-diketopiperazine not being in an α-helix, of the bulk liquid (15). The jet is produced by using a syringe pump (Teledyne-
compared to proteins and peptides that should largely be in Isco) to pressurize the liquid behind a fused silica capillary tip and travels
parallel to the polarization of the incident radiation. Almost immediately
α-helices. Our calculations show the opposite result, with a fea-
(<0.5 mm and <15 μs) after leaving the tip orifice, the liquid is intersected
ture found at ∼404 eV for the β-sheet and not the α-helix. This by the X-ray beam, wherein the sample is close to room temperature
indicates that the cause of this absorption feature is probably not (∼23 °C) (22). The liquid jet is then condensed in a separate cryogenically
cooled section of the chamber to maintain low pressures (∼10−4 torr) in
the interaction region. The signal is normalized to that from a gold mesh
located up-beam of the chamber, and the energy is calibrated based on
specific metal impurities located in the beamline optics. The spectra are
area-normalized and smoothed.

Calculations
Core-Level Spectra. Our theoretical approach has been described
previously (19, 21). We use periodic boundary conditions and a
plane-wave pseudopotential formalism. PWSCF is used to calcu-
late the zone-center electronic structure (31). We use the
Perdew–Burke–Ernzehof (PBE) form of the generalized gradient
approximation to the exchange-correlation potential within
density functional theory (32–35). Core-hole matrix elements
between atomic core levels and unoccupied valence pseudostates
were calculated by reconstructing the electronic structure of the
core region of the excited atom within the atomic frozen core
Fig. 8. Experimental nitrogen K-edge NEXAFS spectrum of solid lysozyme
approximation. Transition amplitudes are estimated in the sin-
(purple, solid line), 2,5-diketopiperazine (black, ••••) and polyisoleucine gle-particle and dipole approximations. The lowest energy
(green, — ••• —) from the literature (see text). These are contrasted with core-hole excited state is modeled by explicit inclusion of the core
calculated spectra at the nitrogen K edge of solid polyalanine in an antipar- hole (via a modified pseudopotential) and the excited electron
allel β-sheet (red, solid line) and an α-helix (blue, ••••). (XCH). Excitations to states above this first excited state are

14012 ∣ www.pnas.org/cgi/doi/10.1073/pnas.1006435107 Schwartz et al.


approximated using the unoccupied Kohn–Sham eigenstates 6 Naþ and Br− ions were added to the triglycine water simulation,
computed from the XCH self-consistent potential (19, 21). and in the final case 4 Naþ ions and 2 SO3 2− ions were added to
The spectra were aligned to experiment. These calculations were the triglycine water simulation. The Naþ and Br− were parame-
performed on the Franklin supercomputer at National Energy terized with the default parameters in AMBER, and SO3 2− was
Research Scientific Computing Center (NERSC). All spectra parameterized using the GAFF (37). We note that this is higher
were stretched along the energy axis by 25% to improve agree- than the experimental concentration, but this was done to empha-
ment with measurements, due to known bandwidth underestima- size any effects that may exist. For statistical purposes, these MD
tion within DFT (22, 23). simulations were continued for a total of 20 ns and sampled every
50 fs. These simulations were used for the RDFs and the rms
Numerical Broadening. All calculated spectra are numerically broa-
positions shown later. The rms positions of the atoms were
dened using Gaussians of 0.2 eV full width at half maximum.
determined by rms fitting the backbone of the triglycine structure
Others often resort to ad-hoc energy-dependent broadening
schemes (36). We use this relatively small and uniform broaden- to the average structure, with the average structure first rms fitted
ing with the aim of simulating and distinguishing electronic and to the first coordinate set to remove rotational effects. The results
vibrational effects explicitly, thereby arriving at a predictive were similar if rotational effects were not removed initially.
computational approach.
Spectrum of Polyalanine. The coordinates of polyalanine in both an
Solvated Triglycine MD Simulations. We have modeled the nuclear ideal α-helix and an ideal antiparallel β-sheet were taken from the
degrees of freedom in these systems using MD performed at literature (24). Because all of the nitrogen atoms are effectively
300 K using a Langevin thermostat, with the generalized AMBER degenerate and motions were not treated, only a single NEXAFS
force field (GAFF) (37). The resulting distribution of nuclear spectrum was calculated for both the α-helix and antiparallel
coordinates was sampled at 40-ps intervals to minimize correla- β-sheet at the nitrogen K edge.
tion between snapshots. This was done for 100 configurations on
each nitrogen (therefore, three calculations per configuration) ACKNOWLEDGMENTS. We thank Wanli Yang for excellent user support of
(37, 38). These coordinates are subsequently used in DFT beamline 8.0.1. This work was supported by the Director, Office of Basic

CHEMISTRY
Energy Sciences, Office of Science, U.S. Department of Energy (DOE) under
calculations, which were by far the most significant part of the Contract DE-AC02-05CH11231 through the Lawrence Berkeley National
computational cost. The system sizes were approximately Laboratory Chemical Sciences Division, the Molecular Foundry, and the
4;000 Å3 and contained approximately 120 water molecules. In Advanced Light Source. Computational resources were provided by NERSC,
one case the triglycine was simulated with water, in another case a DOE Advanced Scientific Computing Research User Facility.

1. Hofmeister F (1888) Zur Lehre von der Wirkung der Salze. Arch Exp Pathol Pharmakol 19. Uejio JS, Schwartz CP, Saykally RJ, Prendergast D (2008) Effects of vibrational
24:247–260. motion on core-level spectra of prototype organic molecules. Chem Phys Lett
2. Kunz W, Henle J, Ninham BW (2004) ‘Zur Lehre von der Wirkung der Salze’ (about the 467:195–199.
science of the effect of salts): Franz Hofmeister’s historical papers. Curr Opin Colloid 20. Prendergast D, Galli G (2006) X-ray absorption spectra of water from first principles
Interface Sci 9:19–37. calculations. Phys Rev Lett 96:215502.
3. Zhang YJ, Cremer PS (2006) Interactions between macromolecules and ions: The 21. Schwartz CP, Uejio JS, Saykally RJ, Prendergast D (2009) On the importance of
Hofmeister series. Curr Opin Chem Biol 10:658–663. nuclear quantum motions in near edge x-ray absorption fine structure spectroscopy
4. Luo EP, Chai YQ, Yuan R (2007) Unsymmetrical tetradentate Schiff bases copper (II) of molecules. J Chem Phys 130:184109.
22. Schwartz CP, et al. (2009) Auto-oligomerization and hydration of pyrrole revealed by
complex as neutral carrier for thiocyanate-selective electrode. Anal Lett 40:369–386.
x-ray absorption spectroscopy. J Chem Phys 131:114509.
5. Batchelor JD, Olteanu A, Tripathy A, Pielak GJ (2004) Impact of protein denaturants
23. Uejio JS, et al. (2010) J Phys Chem B 114:4702–4709.
and stabilizers on water structure. J Am Chem Soc 126:1958–1961.
24. Vener MV, Egorova AN, Fomin DP, Tsirelson VG (2009) Hierarchy of the non-covalent
6. Smith JD, et al. (2005) Unified description of temperature-dependent hydrogen-bond
interactions in the alanine-based secondary structures. DFT study of the frequency
rearrangements in liquid water. Proc Natl Acad Sci USA 102:14171–14174.
shifts and electron-density features. J Phys Org Chem 22:177–185.
7. Nucci NV, Vanderkooi JM (2008) Effects of salts of the Hofmeister series on the 25. Zubavichus Y, Shaporenko A, Grunze M, Zharnikov M (2005) Innershell absorption
hydrogen bond network of water. J Mol Liq 143:160–170. spectroscopy of amino acids at all relevant absorption edges. J Phys Chem A 109:
8. Hess B, van der Vegt NFA (2009) Cation specific binding with protein surface charges. 6998–7000.
Proc Natl Acad Sci USA 106:13296–13300. 26. Zubavichus Y, Shaporenko A, Grunze M, Zharnikov M (2007) NEXAFS spectroscopy of
9. Collins KD (2006) Ion hydration: Implications for cellular function, polyelectrolytes, homopolypeptides at all relevant absorption edges: Polyisoleucine, polytyrosine, and
and protein crystallization. Biophys Chem 119:271–281. polyhistidine. J Phys Chem B 111:9803–9807.
10. Mason PE, et al. (2009) Specificity of ion-protein interactions: Complementary and 27. Zubavichus Y, Shaporenko A, Grunze M, Zharnikov M (2008) Is X-ray absorption
competitive effects of tetrapropylammonium, guanidinium, sulfate, and chloride ions. spectroscopy sensitive to the amino acid composition of functional proteins? J Phys
J Phys Chem B 113:3227–3234. Chem B 112:4478–4480.
11. Cho YH, et al. (2008) Effects of Hofmeister anions on the phase transition temperature 28. Stewart-Ornstein J, et al. (2007) Using intrinsic X-ray absorption spectral differences to
of elastin-like polypeptides. J Phys Chem B 112:13765–13771. identify and map peptides and proteins. J Phys Chem B 111:7691–7699.
12. Zhang Y, et al. (2007) Effects of Hofmeister anions on the LCST of PNIPAM as a function 29. Stöhr J (1992) NEXAFS Spectroscopy (Springer, Berlin) p 403.
of molecular weight. J Phys Chem C 111:8916–8924. 30. Heyda J, Vincent JC, Tobias DJ, Dzubiella J, Jungwirth P (2010) Ion specificity at the
13. Zhang YJ, Furyk S, Bergbreiter DE, Cremer PS (2005) Specific ion effects on the peptide bond: Molecular dynamics simulations of N-methylacetamide in aqueous salt
water solubility of macromolecules: PNIPAM and the Hofmeister series. J Am Chem solutions. J Phys Chem B 114:1213–1220.
31. Baroni S, Corso AD, Gironcoli SD, Giannozzi P (2008) PWSCF v4.0.2.
Soc 127:14505–14510.
32. Hohenberg P, Kohn W (1964) Inhomogeneous electron gas. Phys Rev B 136:B864–B871.
14. Aziz EF, Ottosson N, Faubel M, Hertel IV, Winter B (2008) Interaction between liquid
33. Kohn W, Sham LJ (1965) Self-consistent equations including exchange and correlation
water and hydroxide revealed by core-hole de-excitation. Nature 455:89–91.
effects. Phys Rev 140:A1133–A1138.
15. Wilson KR, et al. (2001) X-ray spectroscopy of liquid water microjets. J Phys Chem B
34. Perdew JP, et al. (1992) Atoms, molecules, solids, and surfaces—applications of
105:3346–3349.
the generalized gradient approximation for exchange and correlation. Phys Rev B
16. Uejio JS, et al. (2008) Characterization of selective binding of alkali cations with 46:6671–6687.
carboxylate by x-ray absorption spectroscopy of liquid microjets. Proc Natl Acad Sci 35. Perdew JP, Burke K, Ernzerhof M (1996) Generalized gradient approximation made
USA 105:6809–6812. simple. Phys Rev Lett 77:3865–3868.
17. Aziz EF, et al. (2008) Cation-specific interactions with carboxylate in amino acid and 36. Naslund LA, et al. (2005) X-ray absorption spectroscopy study of the hydrogen bond
acetate aqueous solutions: X-ray absorption and ab initio calculations. J Phys Chem B network in the bulk water of aqueous solutions. J Phys Chem A 109:5995–6002.
112:12567–12570. 37. Case DA, et al. (2006) AMBER 9. (University of California, San Francisco) v9.0.
18. Wilson KR, et al. (2004) Investigation of volatile liquid surfaces by synchrotron x-ray 38. Wang JM, Wang W, Kollman PA, Case DA (2006) Automatic atom type and bond type
spectroscopy of liquid microjets. Rev Sci Instrum 75:725–736. perception in molecular mechanical calculations. J Mol Graph Model 25:247–260.

Schwartz et al. PNAS ∣ August 10, 2010 ∣ vol. 107 ∣ no. 32 ∣ 14013

You might also like