Common attributes of native-state structures of proteins, disordered proteins, and amyloid

May 2, 2006
103 (18) 6883-6888

Abstract

We show that a framework derived from the common character of globular proteins can be used to understand the design of protein sequences, the behavior of intrinsically unstructured proteins, and the formation of amyloid fibrils in a unified manner. Our studies provide compelling support for the idea that protein native-state structures, the structures adopted by intrinsically unstructured proteins on binding as well as those of amyloid aggregates, all reside in a physical state of matter in which the free energy landscape is sculpted not by the specific sequence of amino acids, but rather by considerations of geometry and symmetry. We elucidate the key role played by sequence design in selecting the structure of choice from the predetermined menu of putative native-state structures.
It is well known that the sequence of amino acids comprising a protein encodes its native-state structure. The protein folding problem, the determination of the native-state structure given a sequence of amino acids, however, has remained unsolved, in part because of the sheer complexity of the problem: the huge number of degrees of freedom associated with the protein atoms and the surrounding water molecules as well as the history dependence implicit in an evolutionary process. It was recently proposed (1, 2) that considerations of symmetry and geometry determine the limited menu of folded conformations from which a protein can choose for its native-state structure (3). Protein structures belong to a novel phase of matter associated with the marginally compact phase of short tubes with a thickness specially self-tuned to be comparable with the range of attractive interactions promoting the compaction. This phase is a finite size effect and exists only for relatively short tubes. That it is poised near a phase transition provides a simple explanation for the flexibility of native-state structures. The marginally compact phase is stabilized by the interplay of the hydrophobic (H) effect and hydrogen-bond formation. The structures that one finds in it are modular in construction, being made up of two principal building blocks, helices and planar sheets; the degeneracy is greatly reduced so that the number of the resulting energy minima is relatively small.
The principal theme of our work is to present a unification of several seemingly distinct aspects of proteins within the unified framework sketched above, which is derived from the common characteristics of all proteins. We will tackle three issues: (i) protein design (412); (ii) intrinsically unstructured proteins (IUPs), which interact with different molecular partners and adopt relatively rigid conformations in the presence of natural ligands (1319); and (iii) the propensity of proteins to misfold and aggregate leading to the formation of amyloid fibrils, which are implicated in debilitating human diseases (2030) such as Alzheimer’s, light-chain amyloidosis, type II diabetes, and spongiform encephalopathies. Our unified picture leads to a prediction of a free energy landscape with two distinct classes of structures. The amyloid phase is dominated by β-strands linked to each other in a variety of forms, whereas the native-state structure menu as well as the structures adopted by IUPs on binding is an assembly of α-helices (AHs) and β-structures. Our results underscore the stunning simplicity provided by the fixed backdrop of the protein folds determined by physical law in the context of which sequences and functionalities are shaped by evolution.

Results and Discussion

Design of Protein Folds with Hydrophobic–Polar (HP) Sequences.

Previous studies have shown that the large number of common attributes of globular proteins reflect a deeper underlying unity in their behavior (1, 2, 31). At odds with conventional belief, a consequence of our hypothesis is that the gross features of the energy landscape of proteins result from the amino acid aspecific common features of all proteins. This landscape is presculpted by general considerations of geometry and symmetry with ≈1,000 broad minima corresponding to putative native-state structures. These structures comprise the menu of folds, and the role of a sequence is not only to select the native structure from this menu but also to be fit in the environment of other cell products. The key point is that for each of these minima the desirable funnel-like behavior (32, 33) is already achieved at the homopolymer level in the marginally compact part of the phase diagram.
Nature has a choice of 20 aa for the design of protein sequences. The design procedure favors the appropriate native-state structure over the other structures and leads to a funnel-like energy landscape conducive for rapid and reproducible folding of that particular protein. The presculpted free energy landscape greatly facilitates the design process. Indeed, within our model, we find that a crude scheme with just two kinds of amino acids, which take into account the H (propensity to be buried) and polar (P) (desire to be exposed to the water) character of the amino acids (5, 34), is sufficient to carry out a successful design of sequences with a variety of target structures. Note that sequence heterogeneity also can be introduced by assigning different local bending energy penalties to different types of residues.
We turn to an investigation of the design of HP sequences of 48 aa, a length comparable with that of some of the shortest proteins, allowing for design into realistic target structures. We use the same model described in Methods with eR uniformly chosen to be equal to 0.2 and eW = −0.4 for H–H pairwise interactions, measured in units in which the local hydrogen bond energy is chosen to be of unit strength. eW is set to 0 if at least one of the interacting residues is P. Fig. 1 shows the “folded native” states of the de novo designed HP sequences of 48 residues (they are the lowest energy conformations obtained in multiple annealing simulations). The target structures were chosen among tertiary folds adopted by small globular proteins, namely the three-helix bundle, a helix placed on a β-sheet, and a sandwich of β-sheets. Strikingly, the resulting folded structures of the designed sequences bear a high resemblance to real protein native states shown in Fig. 1 df.
Fig. 1.
Comparison of de novo designed and real native protein structures. (ac) Folded native states of designed HP sequences. The sequences are all 48 residues long. H and P amino acids are shown in yellow and blue, respectively. The folded native states are the lowest energy conformations, which were obtained several times in multiple simulated annealing simulations that start from an open conformation. In each simulation, one starts at a temperature higher than the folding transition temperature and slowly decreases it to a temperature lower than the transition temperature, followed by a zero temperature quench in a total number of 600 × 106 to 800 × 106 Monte Carlo moves. (df) Real native protein structures, labeled by their Protein Data Bank ID codes, that resemble the designed structures shown in ac. The H amino acids (Ile, Leu, Phe, Met, Val, Trp, Cys, Pro, and Tyr) are shown in yellow. Note that in conformation b and 1pga (e) the helix orientation with respect to the β-sheet is different.
The design procedure is extremely simple and uses distinct HP patterns for AHs and β-sheets. We find that a chain fragment has a substantially high propensity to form an AH if the H amino acids are separated by two or three P residues as one proceeds along the chain, whereas a β-strand is favored if the H residues are separated by one P residue. Strikingly, exactly the same HP patterns have been used in experiments that allow the successful design of de novo proteins and amyloid-like fibrils (5, 10, 11, 35).
After an initial sequence was produced by simply matching the HP patterns described above with the secondary-structure elements in the target structure, some common-sense tailoring of the sequences ensures that one obtains the desired topology of the ground state. For instance, in the case shown in Fig. 1b, we intentionally assigned only one H residue in the rightmost strand to promote the partial exposure of this strand. Strikingly, nature seems to adopt the same strategy as can be seen in the protein with the Protein Data Bank ID code 1pga shown in Fig. 1 df. Nevertheless, the HP sequences of real proteins are not the same as those of the designed sequences. Even though the HP patterns are seen frequently in real native-state protein structures (36), wild-type sequences do not match the patterns exactly.
We find, for instance, few H residues that are exposed in the native structure of 1pga, but this is not the case in our designed structure. An attempt at folding the real HP sequence of 1pga yields a partially folded structure with the AH and one of the β-hairpins formed correctly, whereas the second β-hairpin remains unstructured. This result suggests that more than two types of amino acids may be needed to fold the protein correctly along with the introduction of sequence heterogeneity in the local bending energy, providing a more detailed effective representation of side-chain steric hindrance.
Fig. 2 shows that the free energy landscapes associated with the designed sequences are characterized by several broad minima, one of which corresponds to the folded state shown in Fig. 1 ac. The effective free energy is computed at the folding transition temperature Tf. (In two-state systems the folding transition temperature is commonly defined as the temperature at which the two states are equally populated. In the presence of more than two states, we resort to the method used here of probing the specific heat maximum, which yields a closely related estimate.)
Fig. 2.
Contour plots of the effective free energy of the three designed HP sequences shown in Fig. 1 at the temperature corresponding to the specific heat maximum. In the plots, Tf = 0.251 (a), 0.22 (b), and 0.246 (c) for the three cases. The effective free energy, defined as F(rmsd, E) = −ln P(rmsd, E), is obtained as a function of the total energy E and the rms deviation rmsd (in Å) to the folded state shown in Fig. 1 from the histogram P(rmsd, E) collected in equilibrium Monte Carlo simulations at constant temperature. The spacing between consecutive levels in each contour plot is 1 and corresponds to a free energy difference of kB, where is the temperature in physical units. The darker the color, the lower the free energy value. The typical conformations shown correspond to minima in the free energy landscape in the simulations. Note that the unfolded conformations (Top) are compact with well formed H contacts and are structured (37).
Interestingly, for the two cases shown in Fig. 1 a and b, we find that there is a competing minimum corresponding to a structure that resembles a mirror image of the folded state (see Fig. 2 a and b). For the case of the β-sandwich (Fig. 2c), the competing minima are structures that are β-sandwiches but with different packing of β-sheets or with unequal numbers of strands on each side. Our simple model, without any side chains, is unable to discriminate between these minima. For the case shown in Fig. 2b, we also find that there is a minimum corresponding to a structure (Bottom Right) that has a somewhat lower energy than the folded state. However, this minimum is not populated at the folding transition temperature. In all three cases, the unfolded state (the upper minimum in the free energy landscape) is characterized by disordered yet compact structures with well formed H contacts but with fewer hydrogen bonds than in the folded state.
Our calculations vividly demonstrate that, in our model, where the menu of folds is predetermined by symmetry and geometry, a simple two-letter code for the amino acids is sufficient for the design process to select interesting native structures from this menu. Our results suggest that the incorporation of 20 types of amino acids with a range of hydrophobicities and other attributes should lead to many sequences adopting the same fold as in experiment. More importantly, the rich repertoire of amino acids allows one to shape sequences that not only fold readily into their native state when in isolation, but also allow specific and varied interactions with other proteins and cell products, promoting solubility within the crowded cellular environment and avoiding deleterious aggregation. Remarkably, similar results have been obtained previously for HP sequences in simple cubic lattice models (38), suggesting that the geometric constraints imposed by hydrogen bonding are mimicked by the lattice geometry constraints. This hypothesis is further confirmed by the observation that secondary structure content is enhanced by compactness (39). The use of more realistic, yet simple, off-lattice models allows one to capture both the role of hydrogen bonding and the tube constraint, which are general properties of any polypeptide chain, in determining a marginally compact phase within which the free energy landscape is presculpted.

IUPs.

Natively unstructured proteins or IUPs (13, 15) are biological molecules that under physiological conditions do not exhibit extensive structural order in solution, but often display local and limited residual structure (19). A key feature of IUPs is their high conformational flexibility, which allows them to interact with different molecular partners and adopt relatively rigid conformations in the presence of natural ligands, thus undergoing a loss of conformational entropy upon binding (16). Note, however, that many IUPs carry out their functionality while remaining in the unstructured state. Furthermore, the conformation in the bound state is determined not only by the amino acid sequence but also by the structure of the interacting partner (17). Our Monte Carlo simulations are performed within a cubic box of side 50.0 Å on a chain of 24 amino acids. The relatively large box size ensures that the entropy loss due to confinement is a relatively small contribution to the overall free energy. The value of the hydrophobicity homopolymer parameter, eW, is fixed at −0.08 and the curvature energy penalty eR at 0.3 leading to a ground state of a single straight AH (1): these values have been chosen to allow the homopolymer to ensure that a three-stranded β-sheet (3S) is a local minimum.
It is important to note that for an IUP, the sequence heterogeneity, and more specifically the presence of P residues, leads to no intrinsic ordering under physiological conditions (14, 18). The case considered here is much simpler because we have just one kind of amino acid. Even though the ground state in the absence of any target geometry is ordered, namely a single helix, this simple model system can still be used for IUP modeling by operating at a temperature above the folding transition temperature of the isolated chain. There are local minima present in the free energy landscape, and the “target fold” (or “target structure”) is chosen from among those minima. Our analysis is greatly simplified because we do not include any sequence heterogeneity except for the interaction between the IUP and the target geometry. Our goal is to focus on understanding how this interaction results in the IUP adopting one of the best-fit structures from the menu predetermined at the homopolymer level (1, 2) by geometrical considerations.
The partner of the IUP is represented in a simple coarse-grained framework through three carefully chosen contact points lying on the inner part of the bottom face of the box, each of which has a special affinity to a specific amino acid of the IUP, mimicking a molecular recognition mechanism in the crudest way. The positions of the centers of interaction are chosen based on the geometry of the target fold to maximize the affinity of the IUP to one particular presculpted structure. Also, the bottom wall serves to capture the steric hindrance of the ligand/substrate partner and the associated loss of conformational entropy induced on the IUP. Specific binding centers that are distributed in full three-dimensional space would simplify the design task considerably, but in an unrealistic manner. The range of the contact interaction (rb) is set to 2.0 Å and the interaction energy of a contact between a binding center and its specific amino acid partner, eb, is chosen to be three times the energy of a local hydrogen bond. We present the results of simulations run for two different binding center patterns chosen to promote the folding of the IUP into an AH and a 3S.
Plots of the specific heat as a function of temperature (Fig. 3) show two peaks unlike the single peak observed in the absence of the binding. The higher temperature peak is associated with the adsorption transition of the polymer chain to the bottom wall containing the binding pattern, whereas that at the lower temperature marks the folding transition into the target structure. The relative positions of the peaks in the different cases can be understood by recognizing that the folding transition temperature is increased in the presence of the two patterns, because the entropy of the disordered state is reduced by adsorption to the binding substrate with respect to the bulk case, whereas the intrachain energy of the disordered state is not significantly influenced by the presence of the pattern.
Fig. 3.
Temperature dependence of the specific heat for the process of binding and folding of a homopolymer in the presence of the pattern encoding the AH and the 3S. The behavior of the specific heat of the system in the absence of the substrate is shown as well.
Fig. 4 underscores the role of the binding pattern in shaping the energy landscape of the IUP. Contour plots of the effective free energy as a function of the total number of hydrogen bonds and the total number of H contacts are compared at the same temperature, T = 0.2, for three different cases: no binding pattern present, 3S pattern, and AH pattern. In the first case (note that even in the absence of the binding pattern, the IUP is kept confined within the cubic box) the chosen temperature is slightly above the folding transition temperature (see Fig. 3), and the denatured disordered state is the most populated one. The ground state (single AH) is also populated, and the 3S conformation appears as a competitive local minimum as is the case for a completely free chain.
Fig. 4.
Contour plot of the effective free energy at temperature T = 0.2 for the three different cases: no pattern (a), 3S pattern (b), and AH pattern (c). Histograms P(Nw, Nh) used in the plots have been collected from equilibrium Monte Carlo simulations keeping the two temperatures constant. The effective free energy is defined here as a function of the number of H contacts Nw (y axis) and the total number of hydrogen bonds Nh (x axis), F(Nw, Nh) = −ln P(Nw, Nh). The spacing between consecutive levels in each contour plot corresponds to a free energy difference of kB, where is the temperature in physical units. The binding pattern of the IUP partner is shown using transparent spheres whose radius is equal to the range of their interaction with the corresponding amino acids of the IUP.
In the cases of 3S and AH patterns, T = 0.2 is below the folding transition temperature so that only the free energy minimum corresponding to the target conformation is populated. The competing conformations (single helix for the 3S, β-sheet for the AH) are entirely absent, being incompatible with the chosen binding patterns. We also found a conformation competing with the 3S similar to a Greek-key motif (data not shown), which fits the contact pattern equally well. However, at the folding temperature the 3S target fold is the global free energy minimum.
Our results fit nicely with the known behavior of IUPs if we identify T = 0.2 with native physiological conditions. Without a pattern, the IUP is mostly denatured, but it gets ordered in the presence of a pattern. The type of ordered structure can change and is controlled by the specificity of the pattern. Within our model the ordered structures are present as local minima even in the absence of the pattern. It has been suggested that the ordered conformation attained after binding to the substrate is visited in a preferential way by the protein even in the absence of the substrate (40). Moreover, some amount of ordering can be induced for IUPs without the presence of the substrate by changing environmental conditions (18). In this respect, one should regard the temperature in our simple approach as an effective parameter mimicking such changes in a crude manner.
Note that the distinction between globular proteins with a well defined native structure and the IUPs, which are natively unfolded, is often not clear-cut, because in many cases only a segment of the whole protein is natively unfolded. In this respect, the chain length used in our simulation, although very short when representing an entire protein sequence, ought to be a more reasonable model of the natively unfolded stretches.

Amyloid Fibril Formation.

Several neurodegenerative human diseases such as Alzheimer‘s, spongiform encephalopathies, and light-chain amyloidosis involve the deposition of plaque-like material in tissue arising from the aggregation of specific proteins and peptides (20–22, 24, 25). Irrespective of the native conformations of the proteins involved in such pathologies, which might well be rich in α-helical structure in their normal state, all such pathogenic materials are believed to share three basic common features: (i) fibrillar structure as evident from electron microscopic and atomic force microscope images, (ii) birefringence after congo-red staining, and (iii) cross β-structure as determined from x-ray diffraction patterns (20) and solid-state NMR spectroscopy (22), and very recently solved at the atomic level (30). The fibrillar structures may have various morphologies with different underlying molecular structures that can be controlled by subtle variations in the growth conditions (29). Recent experiments have shown that many other proteins not involved in such diseases are also able to aggregate in vitro forming fibrillar structures sharing the same properties (24, 25). Species formed early in the aggregation of such “artificial” fibrils have been even shown to be cytotoxic themselves (23). This finding suggests (24) that the tendency for proteins to aggregate is a generic property of polypeptide chains.
We have found (2) that small homopeptides have an inherent tendency to form amorphous β-sandwiches. To investigate the role of sequence heterogeneity in amyloid formation, we performed the “experiment” of cutting the sequence (Fig. 1a) with the three-helix bundle fold as its native state into six equal small fragments. Each fragment has the same HP sequence of PPHPPHHP, and they form helices when isolated from each other. Fig. 5 shows that when put together, these small peptides aggregate to form a cross-linked “triangular” β-helix arrangement, which is stabilized within our model by the repeated pattern of H residues being placed in the middle of the sides of the triangle. This triangular β-helix arrangement is akin to one of the all-β folds (single-stranded, left-handed parallel β-helix) present in databases of known protein structures. Strikingly, it also was recently proposed as a molecular model for the aggregated disease-causing isoform (PrPSc) of the prion protein (26). In other more common structural models for amyloid fibrils, two or more β-sheets are layered in a “sandwich” structure (22, 30). Interestingly, the HP patterns are known to influence amyloid formation (27, 36, 41), and the repeated HP pattern used in our specific example is indeed crucial in stabilizing the triangular β-structure in Fig. 5a. Our results reinforce previous suggestions (24) regarding the generic tendency of multiple chains of amino acids to form aggregated β-structures. We find that three main ingredients, i.e., sterics, hydrophobicity, and backbone hydrogen bonds, are sufficient to cause multiple fragments, irrespective of their native conformations when isolated, to form extensive β-structures, either sandwiches of β-sheets or β-helices. Such structures might easily constitute the building blocks of cross-β amyloid fibrils, with hydrogen bonds parallel to the fibril axis, because of their modularity. Our results also suggest that sequence heterogeneity, at the level of HP residue classes, may help to discriminate the nature of the β-structure into which the fragments assemble.
Fig. 5.
Aggregated β-structure of small peptides. (a and a’) Amyloid-like aggregation of six polymer chains, each of length 8. The chains are obtained by cutting the sequence shown in Fig. 1a into six equal fragments, and they have the same HP sequence of PPHPPHHP. Conformation b is a different view of a. The structure shown is the lowest energy conformation obtained in multiple annealing simulations that start from a configuration in which the chains are open and separated. In the simulations, the chains are confined in a cubic of box size 80 Å. The typical number of Monte Carlo moves in an annealing simulation is ≈1 billion (109) steps. We have checked that the aggregated forms shown in a and b are visited and remain stable also in constant temperature simulations started from open conformations for T = 0.2 and 0.21. At higher temperatures, more disordered forms, still rich in β-sheets, dominate. (b and b’) Real triangular β-helix motif (labeled by its Protein Data Bank ID code and residue fragment) that resembles the structure shown at the top.
In a recent report (28), Dobson noted that there is increasing evidence that fibrillar aggregates are not esoteric species associated with a small number of proteins, but instead are a generic form of polypeptide structure that results from the dominance of interactions involving the main chain common to all such molecules. By contrast, the structures of the normally soluble forms of proteins are dominated by the specific packing of the side chains that distinguish one protein sequence from another. One therefore can think of the amyloid diseases as resulting from the “reversion” of the highly evolved biologically functional forms of peptides and proteins into an alternative and unwelcome structural state that exists as a result of the inherent physicochemical nature of polypeptide chains. Our results vividly demonstrate the underlying simplicity of the protein problem, as envisioned by Dobson (28): both the structures of the normally soluble forms of proteins as well as those of fibrillar aggregates result “from the dominance of interactions involving the main chain common to all such molecules,” and the role of side chains of a soluble protein is to select a particular presculpted minimum.

Conclusion

There are many remarkable common characteristics of globular proteins. These characteristics include their ability to fold reproducibly and rapidly, the limited number of native-state folds, the simple modular nature (42) of native-state structures built of helices and almost planar sheets, and the unfortunate tendency of proteins to aggregate into amyloid. We have developed a framework for understanding these common characteristics and show how seemingly disparate phenomena such as the behavior of IUPs, sequence design, and amyloid formation all can be studied within the same framework. Our studies suggest a stunning simplicity to the protein problem and show how considerations of geometry and symmetry lead not only to the menu of native-state structures but also to amyloid structures. Sequence design is shown to play a vital role in selecting the structure of choice from this predetermined menu.

Methods

Following refs. 1 and 2, we consider chain molecules made up of amino acids each represented by its Cα atom [it has been shown (43) that one can approximately reconstruct the locations of all of the backbone atoms with just the knowledge of the Cα positions]. Hydrophobicity is incorporated by means of a pairwise attraction between H amino acids of magnitude eW. The special local direction at each amino acid location along the chain defined by the position of the neighboring amino acids is captured by employing a tubelike description, which leads naturally to the emergence of secondary motifs. The model penalizes sharp local turns of the backbone by means of a bending energy penalty of constant magnitude eR along the chain.
The energetics and geometry of hydrogen bonds are encapsulated in the model based on a statistical analysis of protein native structures. There are strong amino acid aspecific constraints on the relative orientation of the intrinsic, Frenet, coordinate systems associated with the Cα atoms of amino acids between which hydrogen bonds are formed. The independence of such constraints on the types of the hydrogen-bonded residues leads to a significant simplification. The local (nonlocal) hydrogen-bond energies are defined to be −1 (−0.7) with a cooperative energy (44) of −0.3 assigned to any pair of consecutive hydrogen bonds either in a β-sheet or an AH.
Our Monte Carlo simulations are carried out with moves commonly used in stochastic chain dynamics (45) with the standard pivot and crankshaft conformational rearrangements. Our simulations of the binding process for disordered proteins, confined within a cubic box, used two additional moves: a reptation-like move to facilitate decorrelation of the chain conformations and a translation of the chain by 2.5 Å along a randomly chosen direction at regular time intervals (every 24 elementary moves). The chain translation is useful for preventing the chain from getting stuck in the vicinity of the box boundaries, where the acceptance rate decreases because of geometric constraints. The Metropolis acceptance/rejection test is used with a thermal weight exp(−E/T), where E is the energy of the conformation and T is an effective temperature.

Abbreviations:

IUP
intrinsically unstructured protein
H
hydrophobic
P
polar:
HP
H–P
AH
α-helix
3S
three-stranded β-sheet

Acknowledgments

We thank Hue-Sun Chan, Michael Hecht, Stefano Lise, George Rose, and Istvan Simon for stimulating discussions. This work was supported by Programmi di Ricerca Scientifica di Rilevante Interesse Nazionale 2005 No. 2005027330, the National Aeronautics and Space Administration, National Science Foundation Integrative Graduate Education and Research Traineeship Grant DGE-9987589, the National Science Foundation Materials Research Science and Engineering Centers at Penn State, and the Natural Science Council of Vietnam.

References

1
T. X. Hoang, A. Trovato, F. Seno, J. R. Jayanth, A. Maritan Proc. Natl. Acad. Sci. USA 101, 7960–7964 (2004).
2
J. R. Banavar, T. X. Hoang, A. Maritan, F. Seno, A. Trovato Phys. Rev. E 70, 041905 (2004).
3
C. Chothia Nature 357, 543–544 (1992).
4
W. F. DeGrado, Z. R. Wasserman, J. D. Lear Science 243, 622–628 (1989).
5
S. Kamtekar, J. M. Schiffer, H. J. Xiong, J. M. Babik, M. H. Hecht Science 262, 1680–1685 (1993).
6
J. W. Bryson, S. F. Betz, H. S. Lu, D. J. Suich, H. X. Zhou, K. T. O’Neil, W. F. DeGrado Science 270, 935–941 (1995).
7
B. I. Dahiyat, S. L. Mayo Science 278, 82–87 (1997).
8
P. Koehl, M. Levitt J. Mol. Biol 293, 1161–1181 (1999).
9
P. Koehl, M. Levitt J. Mol. Biol 293, 1183–1193 (1999).
10
W. X. Wang, M. H. Hecht Proc. Natl. Acad. Sci. USA 99, 2760–2765 (2002).
11
Y. Wei, S. Kim, D. Fela, J. Baum, M. H. Hecht Proc. Natl. Acad. Sci. USA 100, 13270–13273 (2003).
12
B. Kuhlman, G. Dantas, G. C. Ireton, G. Varani, B. L. Stoddard, D. Baker Science 302, 1364–1368 (2003).
13
P. E. Wright, H. J. Dyson J. Mol. Biol 293, 7960–7964 (1999).
14
A. K. Dunker, J. D. Lawson, C. J. Brown, R. M. Williams, P. Romero, J. S. Oh, C. J. Oldfield, A. M. Campen, C. M. Ratcliff, K. W. Hipps, et al. J. Mol. Graphics Model 19, 26–59 (2001).
15
A. K. Dunker, Z. Obradovic Nat. Biotechnol 19, 805–806 (2001).
16
H. J. Dyson, P. E. Wright Curr. Opin. Struct. Biol 12, 54–60 (2002).
17
P. Tompa Trends Biochem. Sci 27, 527–533 (2002).
18
V. N. Uversky Eur. J. Biochem 269, 2–12 (2002).
19
M. Fuxreiter, I. Simon, P. Friedrich, P. Tompa J. Mol. Biol 338, 1015–1026 (2004).
20
J. W. Kelly Curr. Opin. Struct. Biol 8, 101–106 (1998).
21
S. B. Prusiner Proc. Natl. Acad. Sci. USA 95, 13363–13383 (1998).
22
A. T. Petkova, Y. Ishii, J. J. Balbach, O. N. Antzutkin, R. D. Leapman, F. Delaglio, R. Tycko Proc. Natl. Acad. Sci. USA 99, 16742–16747 (2002).
23
M. Bucciantini, E. Giannoni, F. Chiti, F. Baroni, L. Formigli, J. S. Zurdo, N. Taddei, G. Ramponi, C. M. Dobson, M. Stefani Nature 416, 507–511 (2002).
24
C. M. Dobson Nat. Rev. Drug Discov 2, 154–160 (2003).
25
F. Chiti, M. Stefani, N. Taddei, G. Ramponi, C. M. Dobson Nature 424, 805–808 (2003).
26
C. Govaerts, H. Wille, S. B. Prusiner, F. E. Cohen Proc. Natl. Acad. Sci. USA 101, 8342–8347 (2004).
27
K. F. Dubay, A. P. Pawar, F. Chiti, J. Zurdo, C. M. Dobson, M. Vendruscolo J. Mol. Biol 341, 1317–1326 (2004).
28
C. M. Dobson Science 304, 1259–1262 (2004).
29
A. T. Petkova, R. D. Leapman, Z. Guo, W. M. Yau, M. P. Mattson, R. Tycko Science 307, 262–265 (2005).
30
R. Nelson, M. R. Sawaya, M. Balbirnie, A. O. Madsen, C. Riekel, R. Grothe, D. Eisenberg Nature 435, 773–778 (2005).
31
J. D. Bernal Nature 143, 663–667 (1939).
32
P. G. Wolynes, J. N. Onuchic, D. Thirumalai Science 267, 1619–1620 (1995).
33
K. A. Dill, H. S. Chan Nat. Struct. Biol 4, 10–19 (1997).
34
K. F. Lau, K. A. Dill Macromolecules 22, 3986–3997 (1989).
35
M. W. West, W. Wang, J. Patterson, J. D. Mancia, J. R. Beasley, M. H. Hecht Proc. Natl. Acad. Sci. USA 96, 11211–11216 (1999).
36
M. W. West, M. H. Hecht Protein Sci 4, 2032–2039 (1995).
37
Z. Shi, C. A. Olson, G. D. Rose, R. L. Baldwin, N. R. Kallenbach Proc. Natl. Acad. Sci. USA 99, 9190–9195 (2002).
38
K. Yue, K. A. Dill Proc. Natl. Acad. Sci. USA 92, 146–150 (1995).
39
H. S. Chan, K. A. Dill Proc. Natl. Acad. Sci. USA 87, 6388–6392 (1990).
40
C. J. Tsai, B. Ma, Y. Y. Sham, S. Kumar, R. Nussinov Proteins 44, 418–427 (2001).
41
B. M. Broome, M. H. Hecht J. Mol. Biol 296, 961–968 (2000).
42
N. C. Fitzkee, P. J. Fleming, H. Gong, N. Panasick, T. O. Street, G. D. Rose Trends Biochem. Sci 30, 73–80 (2005).
43
R. Kazmierkiewicz, A. Liwo, H. A. Scheraga J. Comput. Chem 23, 715–723 (2002).
44
A. Liwo, R. Kazmierkiewicz, C. Czaplewski, M. Groth, S. Oldziej, R. J. Rackovski, M. R. Pincus, H. A. Scheraga J. Comput. Chem 19, 259–276 (1998).
45
A. D. Sokal Nucl. Phys. B 47, 172–179 (1996).

Information & Authors

Information

Published in

The cover image for PNAS Vol.103; No.18
Proceedings of the National Academy of Sciences
Vol. 103 | No. 18
May 2, 2006
PubMed: 16624879

Classifications

Submission history

Received: October 4, 2005
Published online: May 2, 2006
Published in issue: May 2, 2006

Keywords

  1. amyloid formation
  2. presculpted free energy landscape
  3. protein design
  4. protein folding

Acknowledgments

We thank Hue-Sun Chan, Michael Hecht, Stefano Lise, George Rose, and Istvan Simon for stimulating discussions. This work was supported by Programmi di Ricerca Scientifica di Rilevante Interesse Nazionale 2005 No. 2005027330, the National Aeronautics and Space Administration, National Science Foundation Integrative Graduate Education and Research Traineeship Grant DGE-9987589, the National Science Foundation Materials Research Science and Engineering Centers at Penn State, and the Natural Science Council of Vietnam.

Authors

Affiliations

Trinh X. Hoang
Institute of Physics and Electronics, Vietnamese Academy of Science and Technology, 10 Dao Tan, Hanoi, Vietnam;
Luca Marsella
International School for Advanced Studies (SISSA), Via Beirut 2-4, I-34014 Trieste, Italy;
Dipartimento di Fisica “G. Galilei,” Università di Padova, Via Marzolo 8, I-35131 Padova, Italy;
Antonio Trovato
Dipartimento di Fisica “G. Galilei,” Università di Padova, Via Marzolo 8, I-35131 Padova, Italy;
Sezione Istituto Nazionale di Fisica Nucleare, Università di Padova, I-35131 Padova, Italy; and
Flavio Seno
Dipartimento di Fisica “G. Galilei,” Università di Padova, Via Marzolo 8, I-35131 Padova, Italy;
Sezione Istituto Nazionale di Fisica Nucleare, Università di Padova, I-35131 Padova, Italy; and
Jayanth R. Banavar [email protected]
Department of Physics, 104 Davey Lab, Pennsylvania State University, University Park, PA 16802
Dipartimento di Fisica “G. Galilei,” Università di Padova, Via Marzolo 8, I-35131 Padova, Italy;
Sezione Istituto Nazionale di Fisica Nucleare, Università di Padova, I-35131 Padova, Italy; and

Notes

To whom correspondence may be addressed. E-mail: [email protected] or [email protected]
Communicated by Morrel H. Cohen, Rutgers, The State University of New Jersey, Bridgewater Township, NJ, March 9, 2006
Author contributions: T.X.H., L.M., A.T., F.S., J.R.B., and A.M. designed research; T.X.H. and L.M. performed research; T.X.H., L.M., A.T., F.S., J.R.B., and A.M. analyzed data; and T.X.H., L.M., A.T., J.R.B., and A.M. wrote the paper.

Competing Interests

Conflict of interest statement: No conflicts declared.

Metrics & Citations

Metrics

Note: The article usage is presented with a three- to four-day delay and will update daily once available. Due to ths delay, usage data will not appear immediately following publication. Citation information is sourced from Crossref Cited-by service.


Citation statements




Altmetrics

Citations

Export the article citation data by selecting a format from the list below and clicking Export.

Cited by

    Loading...

    View Options

    View options

    PDF format

    Download this article as a PDF file

    DOWNLOAD PDF

    Login options

    Check if you have access through your login credentials or your institution to get full access on this article.

    Personal login Institutional Login

    Recommend to a librarian

    Recommend PNAS to a Librarian

    Purchase options

    Purchase this article to access the full text.

    Single Article Purchase

    Common attributes of native-state structures of proteins, disordered proteins, and amyloid
    Proceedings of the National Academy of Sciences
    • Vol. 103
    • No. 18
    • pp. 6777-7198

    Media

    Figures

    Tables

    Other

    Share

    Share

    Share article link

    Share on social media