Skip to main content
  • Submit
  • About
    • Editorial Board
    • PNAS Staff
    • FAQ
    • Rights and Permissions
    • Site Map
  • Contact
  • Journal Club
  • Subscribe
    • Subscription Rates
    • Subscriptions FAQ
    • Open Access
    • Recommend PNAS to Your Librarian
  • Log in
  • My Cart

Main menu

  • Home
  • Articles
    • Current
    • Latest Articles
    • Special Features
    • Colloquia
    • Collected Articles
    • PNAS Classics
    • Archive
  • Front Matter
  • News
    • For the Press
    • Highlights from Latest Articles
    • PNAS in the News
  • Podcasts
  • Authors
    • Information for Authors
    • Purpose and Scope
    • Editorial and Journal Policies
    • Submission Procedures
    • For Reviewers
    • Author FAQ
  • Submit
  • About
    • Editorial Board
    • PNAS Staff
    • FAQ
    • Rights and Permissions
    • Site Map
  • Contact
  • Journal Club
  • Subscribe
    • Subscription Rates
    • Subscriptions FAQ
    • Open Access
    • Recommend PNAS to Your Librarian

User menu

  • Log in
  • My Cart

Search

  • Advanced search
Home
Home

Advanced Search

  • Home
  • Articles
    • Current
    • Latest Articles
    • Special Features
    • Colloquia
    • Collected Articles
    • PNAS Classics
    • Archive
  • Front Matter
  • News
    • For the Press
    • Highlights from Latest Articles
    • PNAS in the News
  • Podcasts
  • Authors
    • Information for Authors
    • Purpose and Scope
    • Editorial and Journal Policies
    • Submission Procedures
    • For Reviewers
    • Author FAQ

New Research In

Physical Sciences

Featured Portals

  • Physics
  • Chemistry
  • Sustainability Science

Articles by Topic

  • Applied Mathematics
  • Applied Physical Sciences
  • Astronomy
  • Computer Sciences
  • Earth, Atmospheric, and Planetary Sciences
  • Engineering
  • Environmental Sciences
  • Mathematics
  • Statistics

Social Sciences

Featured Portals

  • Anthropology
  • Sustainability Science

Articles by Topic

  • Economic Sciences
  • Environmental Sciences
  • Political Sciences
  • Psychological and Cognitive Sciences
  • Social Sciences

Biological Sciences

Featured Portals

  • Sustainability Science

Articles by Topic

  • Agricultural Sciences
  • Anthropology
  • Applied Biological Sciences
  • Biochemistry
  • Biophysics and Computational Biology
  • Cell Biology
  • Developmental Biology
  • Ecology
  • Environmental Sciences
  • Evolution
  • Genetics
  • Immunology and Inflammation
  • Medical Sciences
  • Microbiology
  • Neuroscience
  • Pharmacology
  • Physiology
  • Plant Biology
  • Population Biology
  • Psychological and Cognitive Sciences
  • Sustainability Science
  • Systems Biology

Both the cis-trans equilibrium and isomerization dynamics of a single proline amide modulate β2-microglobulin amyloid assembly

Vladimir Yu. Torbeev and Donald Hilvert
PNAS December 10, 2013 110 (50) 20051-20056; https://doi.org/10.1073/pnas.1310414110
Vladimir Yu. Torbeev
Laboratory of Organic Chemistry, ETH Zurich, CH-8093 Zurich, Switzerland
  • Find this author on Google Scholar
  • Find this author on PubMed
  • Search for this author on this site
Donald Hilvert
Laboratory of Organic Chemistry, ETH Zurich, CH-8093 Zurich, Switzerland
  • Find this author on Google Scholar
  • Find this author on PubMed
  • Search for this author on this site
  • For correspondence: hilvert@org.chem.ethz.ch
  1. Edited by Ronald T. Raines, University of Wisconsin–Madison, Madison, WI, and accepted by the Editorial Board October 29, 2013 (received for review June 1, 2013)

  • Article
  • Figures & SI
  • Info & Metrics
  • PDF
Loading

Significance

β2-Microglobulin is an abundant and normally soluble protein. In patients undergoing chronic dialysis, however, it forms insoluble amyloid plaques, leading to medical complications. It has been suggested that the conformational transformation of soluble protein monomers into polymeric amyloids is mediated by isomerization of a single amino acid, namely, proline 32. In this study, we probed the role of this amino acid by chemically synthesizing uniquely tailored protein analogs containing noncanonical amino acids at position 32. Our results show that both the chemical equilibrium and rate of cis-trans isomerization of proline 32 are critical for the solubility of β2-microglobulin and its self-assembly into morphologically distinct amyloid fibrils. These insights may aid ongoing efforts to provide remedies against dialysis-related amyloidosis.

Abstract

The human protein β2-microglobulin (β2m) aggregates as amyloid fibrils in patients undergoing long-term hemodialysis. Isomerization of Pro32 from its native cis to a nonnative trans conformation is thought to trigger β2m misfolding and subsequent amyloid assembly. To examine this hypothesis, we systematically varied the free-energy profile of proline cis-trans isomerization by replacing Pro32 with a series of 4-fluoroprolines via total chemical synthesis. We show that β2m’s stability, (un)folding, and aggregation properties are all influenced by the rate and equilibrium of Pro32 cis-trans isomerization. As anticipated, the β2m monomer was either stabilized or destabilized by respective incorporation of (2S,4S)-fluoroproline, which favors the native cis amide bond, or the stereoisomeric (2S,4R)-fluoroproline, which disfavors this conformation. However, substitution of Pro32 with 4,4-difluoroproline, which has nearly the same cis-trans preference as proline but an enhanced isomerization rate, caused pronounced destabilization of the protein and increased oligomerization at neutral pH. More remarkably, these subtle alterations in chemical composition—incorporation of one or two fluorine atoms into a single proline residue in the 99 amino acid long protein—modulated the aggregation properties of β2m, inducing the formation of polymorphically distinct amyloid fibrils. These results highlight the importance of conformational dynamics for molecular assembly of an amyloid cross-β structure and provide insights into mechanistic aspects of Pro32 cis-trans isomerism in β2m aggregation.

  • protein conformation
  • polymorphism
  • amyloidogenesis
  • native chemical ligation

Human β2-microglobulin (β2m) is a component of the class I major histocompatibility complex (1). Although the WT protein is generally stable at neutral pH, in kidney disease patients undergoing long-term hemodialysis, it forms insoluble amyloid deposits causing a condition known as dialysis-related amyloidosis (2). Collagen and glycosaminoglycans facilitate fibril formation at neutral pH (3, 4), as do organic solvents (5), divalent metal ions (Cu2+) (6, 7), and surfactants (8). Amyloids can also be induced at low pH (<3) (9) or by “seeding” with stabilized preformed amyloids (10). Despite more than two decades of research, however, the molecular mechanism by which soluble β2m is converted into insoluble fibrils remains poorly understood (11).

Wild-type β2m (WT β2m) contains a highly conserved, cis-configured proline at position 32 that is required for maintaining the soluble native structure (Fig. 1) (12). Several lines of evidence, including detailed folding studies and mutagenesis experiments on β2m, have suggested that cis-to-trans isomerization of Pro32 serves as a trigger to misfolding and subsequent aggregation (6, 13). In β2m amyloids, this residue adopts a trans conformation (14, 15), and additives that increase the equilibrium concentration of β2m conformers containing trans-Pro32 have been shown to promote amyloid formation (11). A trans-proline amide is also observed in the X-ray structure of the hexameric species that is formed upon treatment of the H13F variant with Cu2+ ions (7). Although replacement of Pro32 with glycine, which has a stronger propensity to adopt a trans amide conformation than proline, facilitates fibril elongation in seeding experiments (13), population of a trans-amide at residue 32 is not sufficient in itself to induce fibrillation. Mutation of Pro32 to alanine and valine, which have much higher trans amide bond propensities (>99.9%) than proline, does not result in enhanced amyloid formation (11). Other factors, such as protein conformational dynamics (11), are probably important for the initiation and assembly processes.

Fig. 1.
  • Download figure
  • Open in new tab
  • Download powerpoint
Fig. 1.

Human β2-microglobulin (β2m). (A) Primary amino acid sequence of the 99-residue β2m with Pro32 highlighted in red and the two cysteine-containing ligation sites underlined. (B) Native structure of β2m showing the BC loop (cyan) containing cis-Pro32 (red) (based on PDB ID 2YXF) (C) Structure of proline and the three fluorine-containing proline analogs used in this study: (2S, 4S)-4-fluoroproline (4S-fpr), (2S, 4R)-4-fluoroproline (4R-Fpr), and 4,4-difluoroproline (F2Pro). (D) Stereoelectronic effects associated with the C4 substituent influence both the proline ring conformation as well as the cis-trans amide bond equilibrium (17). An electron-withdrawing group at the 4R position stabilizes the exo ring pucker via a gauche effect. This conformation preorganizes main chain torsion angles, resulting in stabilization of the trans peptide bond isomer via an n→π* orbital interaction. The equilibrium cis-trans propensities for proline and its fluorinated analogs follow a different trend than the corresponding trans-cis isomerization rates (16).

In this paper, we report a robust chemical synthesis of the 99-residue human β2m protein that allows us to probe the role of Pro32 cis-to-trans isomerization in amyloid formation with mechanistically more informative substitutions than can be achieved using conventional site-directed mutagenesis. Specifically, we generated three β2m variants in which Pro32 was replaced with (2S, 4S)-4-fluoroproline (4S-fpr), (2S, 4R)-4-fluoroproline (4R-Fpr), and 4,4-difluoroproline (F2Pro), respectively (16) (Fig. 1C). Fluorinated proline analogs have been used extensively to probe the structure and stability of many proteins (17⇓⇓–20). They are nearly isosteric to the natural amino acid and preserve the unique conformational properties of the five-membered pyrrolidine ring but exhibit altered cis-trans amide isomer ratios. For example, the configuration at C4 in 4R-Fpr stabilizes the exo ring pucker via a gauche effect, which, in turn, leads to higher stability of the trans-prolyl conformer [6.7:1 versus 4.6:1 for proline in the model compound Ac-Xaa-OMe (20)] (Fig. 1D). The opposite is true for 4S-fpr, which favors the endo ring pucker, reducing the preference for the trans-prolyl conformer to a factor of 2.5:1 (20). F2Pro has similar cis-trans peptide bond propensities as native proline but a significantly lower activation barrier for isomerization (16). Biochemical characterization of WT β2m and the three analogs provides unique insight into the mechanistic consequences of perturbing the Pro32 peptide bond on protein (un)folding and subsequent aggregation to give oligomers and amyloid fibrils.

Results

Chemical Synthesis of Human β2m and Analogs by Native Chemical Ligation.

WT β2m was assembled by “one pot” native chemical ligation of three unprotected peptide fragments (Fig. 2 and SI Appendix, Figs. S1 and S2) (21). The peptide segments were synthesized using an “in situ neutralization” protocol for Boc-based solid phase peptide synthesis (22) and subsequently purified to homogeneity by reverse-phase HPLC (RP-HPLC). Native chemical ligation of the C terminal and middle fragments was followed by thiazolidine (Thz) deprotection, ligation with the N-terminal peptide, and formyl-deprotection of tryptophans. After RP-HPLC purification, formation of the native disulfide bond by air-oxidation, and further RP-HPLC purification, the protein was folded by dialysis against 25 mM sodium phosphate buffer (pH 7.5). High purity of synthetic β2m was verified by analytical RP-HPLC and high-resolution mass spectrometry (Fig. 2B). The protein adopts a native fold as judged by CD spectroscopy (Fig. 3A) and 1H-NOESY NMR spectroscopy (SI Appendix, Fig. S8).

Fig. 2.
  • Download figure
  • Open in new tab
  • Download powerpoint
Fig. 2.

Chemical synthesis of WT β2m and its analogs. (A) The synthetic strategy involved one pot native chemical ligation of three polypeptide fragments. (B) Reverse-phase HPLC analysis and electrospray ionization (ESI) mass spectrum of chemically synthesized β2m (1–99) (SI Appendix, Fig. S2). Analogous analytical characterization of the fluoroproline-containing β2m variants is presented in SI Appendix, Figs. S3–S6.

Fig. 3.
  • Download figure
  • Open in new tab
  • Download powerpoint
Fig. 3.

Properties of chemically synthesized WT β2m and three chemical analogs. (A) Far-UV CD spectra at 37 °C. (B) Thermal denaturation monitored by CD. (C) Chemical denaturation with GdmCl. A two-state folding mechanism was assumed to compare the data for different analogs. The data are summarized in Table 1. (D) Fluorescence of ANS dye [concentration (c) 10 μM] upon addition of WT β2m and three analogs (c protein 10 μM). (E) Size-exclusion HPLC of samples (c 40 μM) incubated for 2 wk at pH 7.5, 37 °C. (F) Kinetics of oligomerization of the variants (c 40 μM) in the presence of Cu2+ (c 80 μM) at pH 6.8. (G) The pH-dependent solubility of the synthetic proteins upon incubation for 2 wk (initial c 40 μM in 0.2 M Na.phosphate, 0.1 M Na.citrate buffers). For each plot, data for WT β2m are shown in black, data for [4S-fpr32]β2m are shown in green, data for [4R-Fpr32]β2m are shown in blue, and data for [F2Pro32]β2m are shown in red.

The three β2m analogs containing the noncanonical amino acids 4R-Fpr, 4S-fpr, and F2Pro at position 32 were synthesized analogously using N-Boc-protected 4-fluoroprolines instead of proline (see SI Appendix and Figs. S3–S6 for analytical data). Near-UV CD and 1H-NOESY spectra (SI Appendix, Figs. S7–S10) show that the variants adopt the same overall tertiary structure as WT β2m. As expected for such close structural analogs, only small differences localized near residue 32 are observed in the 1H-NOESY spectra (SI Appendix, Figs. S8–S10).

Fluoroprolines Modulate β2m Stability.

Far-UV CD spectra and thermal denaturation curves for WT β2m and the three analogs are depicted in Fig. 3 A and B, respectively. Chemically synthesized β2m has a CD spectrum that is virtually identical to the published spectra of recombinantly expressed protein containing an additional N-terminal methionine (9, 23). Chief features are a strong maximum at 203 nm and a weak minimum at 221 nm. The CD spectra of [4S-fpr32]β2m and [4R-Fpr32]β2m are similar (Fig. 3A). The relative stability of these three proteins was assessed by thermal denaturation. Their apparent melting temperatures (Tm) correlate with the cis-amide conformational preferences of the respective proline derivative at position 32 (Fig. 1D), decreasing in the order [4S-fpr32]β2m > WT β2m > [4R-Fpr32]β2m (Table 1).

View this table:
  • View inline
  • View popup
Table 1.

Thermodynamic stability and unfolding kinetics

The [F2Pro32]β2m protein deviates from these trends. Its far-UV CD spectrum shows a diminished maximum at 203 nm and a slightly blue-shifted minimum at 218 nm (Fig. 3A). Interestingly, similar features were previously observed for β2m variants containing trans-configured Pro32, including the P5G point mutant and ΔN6, which lacks six N-terminal amino acids (23). Like these proteins, [F2Pro32]β2m exhibits reduced thermal stability (Tm = 56 °C) and less cooperative thermal denaturation compared with WT β2m (Fig. 3B).

Chemical denaturation (Fig. 3C and SI Appendix, Fig. S13) and folding–unfolding kinetics (SI Appendix, Fig. S14) of the Pro32 variants further illuminate the interplay between cis-trans equilibrium propensities, the free-energy barrier for amide bond isomerization, and protein stability. As shown in Table 1, cis-trans equilibrium preferences correlate with the chemical denaturation midpoint ([D]50%]) and unfolding rate (ku), whereas the isomerization barrier is reflected in the cooperativity of denaturation (m value) and refolding kinetics. The Gibbs free energy for unfolding (ΔGu) is lowest for [F2Pro32]β2m and highest for [4S-fpr32]β2m.

Because F2Pro and unmodified proline have comparable cis-trans equilibrium ratios, the destabilizing effects due to incorporation of residue F2Pro32 are counterintuitive. Although the 1H-NOESY NMR spectra (SI Appendix, Fig. S10) provide strong evidence that [F2Pro32]β2m and WT β2m adopt similar folds, the far-UV CD spectra and folding data suggest that F2Pro32 is more structurally disruptive than the monofluorinated prolines. To gain additional insight into the effects of this substitution, we prepared a derivative of [F2Pro32]β2m in which seven residues in the hydrophobic core and an adjacent loop (Val9, Ile7, Val27, Gly29, Gly18, Ala15, Phe22) plus two residues flanking F2Pro32 (Phe30 and Ile35) were 15N-labeled (SI Appendix, Fig. S11A). The 1H-15N resonances in the 1H-15N heteronuclear single quantum coherence (HSQC) spectrum of this variant exhibit average chemical shifts that differ by less than 0.1 ppm from the corresponding resonances in WT β2m (SI Appendix, Fig. S11). In addition, 19F NMR studies showed that the F2Pro32 amide adopts a cis conformation in the folded protein but a trans conformation when denatured (SI Appendix, Fig. S12). Together, these observations strengthen the conclusion that the average structures of [F2Pro32]β2m and WT β2m are quite similar.

The destabilization of [F2Pro32]β2m can be rationalized by additionally considering the energy barrier for cis-trans isomerization (16). Inserting fluorine atoms into the proline five-membered ring reduces the double-bond character of the prolyl amide and lowers the free-energy barrier for cis-trans isomerization. For the model system Ac-Xaa-OMe, the energy barrier decreases in the order Pro > 4R-Fpr ∼ 4S-fpr > F2Pro (Fig. 1D) (16). Introducing F2Pro, the most conformationally labile residue in the series, at position 32 of β2m may destabilize the protein by “melting” the BC loop that maintains Pro32 in the cis conformation (Fig. 1B). Consistent with a more dynamic structure and increased exposure of hydrophobic surfaces, [F2Pro32]β2m binds the environmentally sensitive dye 8-anilino-1-naphthalenesulfonic acid (ANS) to a greater extent than the other variants, giving rise to enhanced fluorescence and a characteristic blue shift of the emission maximum (Fig. 3D). As for the previously published P32G variant (13), minor structural perturbations in the vicinity of residue 32, enhanced by the rapid isomerization kinetics of the F2Pro32 amide, may be propagated to other parts of the molecule without altering the core structure of the protein. The lower stability of [F2Pro32]β2m can then be attributed either to this slightly altered, less compact structure and/or to changes in the denatured state.

Aggregation Properties of Chemically Synthesized β2m Analogs.

The oligomerization state of β2m variants was analyzed by size-exclusion chromatography. Upon incubation at pH 7.5 for a period of 2 wk, WT β2m and the monofluoro-proline analogs (40 μM each) behaved as stable monomeric proteins, whereas [F2Pro32]β2m afforded a population of oligomers (Fig. 3E). Although previous studies on amyloidogenic proteins have shown that partially folded species serve as precursors to aggregates and, ultimately, amyloids (24⇓⇓–27), analysis of these samples by transmission electron microscopy (TEM) showed no evidence for amyloid fibrils even for the least stable [F2Pro32]β2m protein.

Oligomerization of β2m at neutral pH is known to be promoted by Cu2+ ions (6, 7). Although this process is characterized by a several hour lag phase in the case of WT β2m, the P32A mutant produces protein oligomers instantaneously in the presence of metal ions (6). The trans amide conformation adopted by Ala32 apparently preorganizes the protein for tighter Cu2+ binding, triggering formation of oligomers. The synthetic proteins also undergo Cu2+-induced oligomerization at rates that correlate with their copper affinity and preference for the trans amide conformation (Fig. 3F). Thus, the [4R-Fpr32]β2m analog, which favors the trans-amide and has the highest affinity for Cu2+ (Kd = 7.31 ± 0.53 μM), shows the most pronounced fluorescence response in a ThT-binding assay, followed by WT β2m (Kd = 10.3 ± 0.49 μM) and [4S-fpr32]β2m (Kd = 20.2 ± 0.84 μM). Interestingly, no oligomerization was detected with [F2Pro32]β2m, which has the lowest affinity for Cu2+ (Kd = 70.8 ± 4.59 μM) (SI Appendix, Fig. S15). Enhanced flexibility of the BC loop in this variant, as discussed above, may hinder formation of an appropriately preorganized environment for Cu2+ coordination even though F2Pro and unsubstituted proline have similar conformational preferences.

Protein solubility was further examined as a function of pH. The synthetic proteins (40 μM) were incubated at 37 °C without stirring for a period of 2 wk. The concentration of soluble protein in filtered (0.2 μm) aliquots was quantified by absorbance at 280 nm. The resultant solubility curves are shown in Fig. 3G. The most stable analog, [4S-fpr32]β2m, only precipitated below pH 3.5, whereas the other proteins began to form insoluble aggregates at higher pH values. The least stable protein, [F2Pro32]β2m, displayed the least cooperative behavior. However, in no sample were amyloid fibrils detected above pH 4, either by ThT fluorescence or TEM.

Polymorphism of β2m Amyloids Regulated by cis-trans Pro32 Isomerization.

Incubation at acidic pH (<3) is the most reliable way to produce β2m amyloids (9). Indeed, all synthetic variants spontaneously form amyloid fibrils at pH 2.5 within a few days to a week (Fig. 4A). Monitoring amyloid growth at this low pH value showed that WT β2m, [4S-fpr32]β2m, and [4R-Fpr32]β2m form fibrils at qualitatively similar rates, whereas [F2Pro32]β2m aggregates more slowly. Nevertheless, intrinsic tryptophan fluorescence suggests that the different Pro32 analogs give rise to distinct fibril polymorphs (SI Appendix, Fig. S16). For example, WT and [4S-fpr32]β2m amyloids afford similar fluorescence spectra, but the fluorescence emission of [4R-Fpr32]β2m amyloids is red-shifted and strongly diminished; amyloids derived from [F2Pro32]β2m show intermediate values. These results suggest that Trp60, which is buried in WT amyloids (28), is located in very different environments in the other polymorphs as a consequence of altered self-assembly of the respective cross-β structures.

Fig. 4.
  • Download figure
  • Open in new tab
  • Download powerpoint
Fig. 4.

Different polymorphic forms of the amyloids formed by WT β2m and its variants. (A) Substitution of Pro32 by 4S-fpr, 4R-Fpr, or F2Pro led to formation of fibrils of different morphology as seen in the TEM images. Growth conditions: 50 μM protein in 50 mM citrate buffer, 100 mM NaCl, pH 2.5 for 2 wk at 37 °C with shaking (250 rpm). Relative growth rates: WT β2m ∼ [4S-fpr32]β2m ∼ [4R-Fpr32]β2m > [F2Pro32]β2m. (Scale bar, 200 nm.) (B) Kinetics of fibril growth of WT β2m and the chemical analogs upon seeding with preformed WT, [4S-fpr32]β2m, [4R-Fpr32]β2m, or [F2Pro32]β2m amyloid seeds monitored by the increase in ThT fluorescence at pH 2.5. Differences in maximum ThT fluorescence levels for the different protein analogs are presumably due to structurally distinctive fibril surfaces that bind ThT dye to different extents. Control absorption measurements (at 280 nm) of protein remaining in solution confirmed that most of the sample (>90%) had precipitated at the end of each experiment.

We performed seeding experiments to characterize the kinetics of amyloid growth and structural differences in amyloids derived from the different protein variants. The preformed amyloid seeds serve as growth templates, effectively eliminating the nucleation requirement and hence the kinetic lag phase (29). This templating effect is structure-dependent: amyloid seeds of a particular protein polymorph are generally most efficient at inducing fibrils of the same protein polymorph (30, 31). Fibril formation was monitored by a standard ThT fluorescence assay using 40 μM protein in the presence of seeds derived from preformed amyloids of all four variants (16 combinations in total) (Fig. 4B).

The results obtained using cognate seeds, which likely resemble the nucleating species that arise spontaneously during unseeded aggregation, show the same qualitative trend for aggregation rates observed in the absence of seeds, namely WT β2m ∼ [4S-fpr32]β2m ∼ [4R-Fpr32]β2m > [F2Pro32]β2m. Because [4R-Fpr32]β2m does not aggregate faster than other variants, this trend does not support the simple hypothesis that aggregation is driven by the cis-trans amide bond ratio of residue 32. Other factors, including isomerization kinetics and the altered conformational preferences of the proline ring itself, probably play a role.

In cross-seeding experiments, both growth onset and assembly rate depend on the origin of the seeds. In Fig. 4B, the kinetics of fibril growth of the four studied proteins is quantitatively similar in the case of WT and [F2Pro32]β2m seeds, whereas [4S-fpr32]β2m and [4R-Fpr32]β2m seeds produced very different growth patterns. Furthermore, analysis of the samples upon completion of the reaction by TEM and intrinsic fluorescence indicates that seeding with WT or [F2Pro32]β2m amyloid seeds resulted in fibrils with well-defined morphology in all cases (SI Appendix, Fig. S17); they also exhibit a strong intrinsic tryptophan emission signal (SI Appendix, Fig. S18). In contrast, experiments with [4S-fpr32]β2m and [4R-Fpr32]β2m seeds yielded well-defined fibrils for [4S-fpr32]β2m and [4R-Fpr32]β2m, respectively, but gave mixtures of long, straight amyloids and large amounts of amorphous precipitate for the WT β2m and [F2Pro32]β2m proteins (SI Appendix, Fig. S17).

Although [4S-fpr32]β2m and [4R-Fpr32]β2m seeds have, individually, distinctive templating properties, hybrid seeds derived from amyloid fibrils grown from a mixture of [4S-fpr32]β2m and [4R-Fpr32]β2m proteins recapitulate the properties of WT β2m seeds (SI Appendix, Figs. S19 and S20). Thus, growth kinetics for the four β2m variants seeded with WT and hybrid seeds was equivalent (compare Fig. 4B, Top Left, and SI Appendix, Fig. S20A). Intrinsic tryptophan fluorescence of hybrid amyloids generated from a 1:1 mixture of [4S-fpr32]β2m and [4R-Fpr32]β2m proteins (SI Appendix, Fig. S20B) also suggests that their structural properties approximate those of WT fibrils. At pH 2.5, β2m exists as a mixture of partially structured and extensively unfolded conformers (32). Because the BC loop that maintains Pro32 in the cis conformation will be (partially) denatured under these conditions, the average trans-cis amide bond ratio for an equimolar mixture of 4R-Fpr (6.7:1) and 4S-fpr (2.5:1) effectively mimics that of unmodified proline (4.6:1) (20). Because hybrid fibrils formed by two β2m analogs containing isomeric 4-fluoroprolines behave like WT β2m amyloids, it is unlikely that replacement of hydrogen with a larger fluorine atom modulates the structure and stability of the amyloid fibrils for steric reasons. Instead, polymorphism would appear to have a conformational origin.

Discussion

Fluorinated proline derivatives are useful probes of structure–function relationships in proteins (16⇓⇓⇓–20). In addition to altering the equilibrium population of cis and trans isomers via stereoelectronic effects, fluorine substitution lowers the barrier for isomerization by reducing the double-bond character of the prolyl amide (16). Fig. 1D shows schematically the relative order of the two effects. The values for trans-cis equilibrium ratios for proline and its three 4-fluoroproline variants follow a different trend than the corresponding trans-cis isomerization rates. These counterbalanced properties are useful for studying folding, misfolding, and amyloid formation in β2m. A robust chemical synthesis of β2m allows site-specific replacement of one out of the five prolines with unnatural 4-fluoroproline analogs to dissect how the conformational properties of Pro32 influence folding and misfolding mechanisms.

Previous studies with WT β2m showed that trans→cis prolyl bond isomerization of Pro32 is the rate-limiting step in the protein-folding mechanism (33, 34). As a consequence, the apparent folding rate of WT β2m is independent of denaturant concentration (SI Appendix, Fig. S14A). In the case of [4S-fpr32]β2m and [4R-Fpr32]β2m, a weak dependence on denaturant concentration is evident on their folding kinetics (SI Appendix, Fig. S14 B and C), in agreement with lower cis-trans isomerization barriers for both 4S-fpr and 4R-Fpr residues. In contrast, folding of [F2Pro32]β2m showed a strong dependence on denaturant concentration (SI Appendix, Fig. S14D) and proceeded much faster than folding of WT β2m at low denaturant concentrations. In essence, the trans→cis Pro32 isomerization step is no longer rate-limiting when Pro32 is replaced by F2Pro, the residue with the lowest barrier for cis-trans isomerization.

Thermal and chemical denaturation of the folded β2m variants shows that their stability decreases in the order [4S-fpr32]β2m > WT β2m > [4R-Fpr32]β2m > [F2Pro32]β2m (Table 1). Thus, monomeric β2m is stabilized relative to WT when Pro32 is replaced with 4S-fpr, which favors the native-like cis-Pro32 conformer, and destabilized when proline is replaced with 4R-Fpr, which favors the nonnative trans-Pro32 conformer. In energetic terms, these effects are relatively modest, however. The 0.66 kcal/mol destabilization observed for [4R-Fpr32]β2m was not sufficient to induce protein oligomerization at neutral pH. Unexpectedly, [F2Pro32]β2m proved to be the least stable variant. Even though F2Pro and proline have similar cis-trans-amide ratios in model compounds, the Pro32 to F2Pro substitution destabilized β2m by 1.3 kcal/mol. This seemingly innocuous substitution also increased the dynamic properties of the protein, judging from enhanced ANS binding and facile oligomerization, probably by introducing structural flexibility in the BC loop where residue 32 is located. Although the NMR evidence shows that the [F2Pro32]β2m tertiary structures are, on average, very similar to that of WT β2m (SI Appendix, Figs. S10 and S11), enhanced flexibility of the F2Pro32 amide bond likely increases the solvent exposure of normally buried hydrophobic groups.

Much experimental evidence suggests that fibril formation occurs via metastable, partially unfolded protein conformers (26, 27). In the case of β2m, such species are obtained when Pro32, normally in a cis conformation, slowly isomerizes to trans (13, 33, 35). Cis-to-trans isomerization of this residue exerts a destabilizing effect on protein structure (13, 23), and ensuing conformational changes lead to exposure of hydrophobic residues and intermolecular aggregation via the D strand of the β-sandwich structure (6). The destabilized [F2Pro32]β2m variant possesses the characteristics of a partially unfolded conformer, leading to protein aggregation over a wide pH range. However, above pH 4, the products of [F2Pro32]β2m aggregation are either soluble oligomers or insoluble amorphous aggregates. Well-ordered amyloid fibrils were not observed for [F2Pro32]β2m, even after incubating the samples at neutral pH for a period of several weeks. Longer incubation may be necessary to convert such amorphous aggregates into “crystalline” amyloid fibrils.

At pH 2.5, all four synthetic proteins spontaneously formed typical amyloid structures, albeit with unique morphologies and spectroscopic signatures. Amyloid polymorphism is well known: a single sequence, under different conditions, can generate a range of amyloid structures (36). For example, WT β2m forms amyloids of different morphology and stability from solvent mixtures containing varying amounts of trifluoroethanol (37). Coprecipitation of WT β2m and the truncated ΔN6 variant was also shown to produce a distinct polymorph (38). Here, substitution of one or two hydrogens in Pro32 by fluorine elicits a similar effect.

The polymorphic nature of the fibrils is best illustrated in Fig. 4B. The four β2m variants show distinct amyloid growth kinetics, which vary depending on what kind of “seeds” were used to initiate growth. Amyloids of two proteins, WT β2m and [F2Pro32]β2m, served as efficient seeds for all four proteins, whereas seeds derived from [4S-fpr32]β2m and [4R-Fpr32]β2m selectively initiated amyloid growth of the cognate proteins only. Thus, the onset and kinetics of amyloid growth essentially depend on i) structural information encoded in amyloid seeds and ii) the competence of a particular β2m analog to perpetuate this structural information.

Although favoring aggregation, population of a trans-Pro32 amide conformation does not fully explain amyloid formation in WT β2m. The [4R-Fpr32]β2m variant, which favors trans-Pro32 to a greater extent than native β2m, formed amyloids that are spectroscopically distinct from WT fibrils (SI Appendix, Fig. S16). The structural and seeding properties of WT amyloids were better mimicked by fibrils grown from mixtures of this variant and [4S-fpr32]β2m, which should favor cis-Pro32 to a greater extent than native β2m. These observations suggest that cis-Pro32 conformers may be important for productive assembly of WT-like fibrils, presumably via transient hybrid oligomers (Fig. 5). Subsequent isomerization of the prolyl bond during fibril maturation would then exclusively afford the trans isomer observed by solid-state NMR (14, 15). Furthermore, growth kinetics for [F2Pro32]β2m amyloids shows that the activation energy for isomerization of amide32 also influences fibril assembly. The [F2Pro32]β2m protein, which has the most flexible His31-Xaa32 peptide bond and forms oligomers and amorphous precipitate more readily than the other β2m variants, showed the slowest kinetics of amyloid growth, even with “homologous” seeds (Fig. 4B).

Fig. 5.
  • Download figure
  • Open in new tab
  • Download powerpoint
Fig. 5.

Schematic representation of how cis-trans Pro32 isomerization might influence the mechanism of β2m amyloid assembly. The trans-Pro32 conformer dictates self-assembly into a polymorph that places Trp60 in an environment where fluorescence emission is quenched, whereas coassociation of cis-Pro32 and trans-Pro32 conformers leads to some of the Trp60 shielded in the core of cross-β structure and hence stronger fluorescence. Subsequent Pro32 cis→trans isomerization leads to rearrangement of the β-strands to give the distinctive cross-β structure of WT β2m amyloids.

In conclusion, this study links conformational isomerization of a single residue in the β2m sequence (Pro32) with global characteristics of the resultant amyloids such as aggregation and propagation properties. Extending this type of conformational analysis to multiple protein sites can be expected to delineate subtle molecular details and functional aspects of the polymorphic states of these and many other amyloids. Infectious prions in which conformational polymorphism gives rise to multiple distinct heritable strains are a pertinent example (39).

Materials and Methods

Details concerning the synthesis and characterization of the protein samples can be found in SI Appendix, Materials and Methods and SI Appendix, Figs. S1–S6. Biophysical characterization of the proteins is described in SI Appendix, Biophysical Characterization and SI Appendix, Figs. S7–S15. Protein oligomerization studies are detailed in SI Appendix, Protein Aggregation and SI Appendix, Figs. S16–S20.

Acknowledgments

We are grateful to Dr. Marc-Olivier Ebert for help with NMR measurements and Dr. Cindy Schulenburg for advice on protein folding. Electron microscopy measurements were performed at the Electron Microscopy Center of the ETH Zurich (EMEZ). This work was generously supported by the ETH Zurich and an ETH Fellowship to V.Y.T.

Footnotes

  • ↵1To whom correspondence should be addressed. E-mail: hilvert{at}org.chem.ethz.ch.
  • Author contributions: V.Y.T. and D.H. designed research; V.Y.T. performed research; V.Y.T. and D.H. analyzed data; and V.Y.T. and D.H. wrote the paper.

  • The authors declare no conflict of interest.

  • This article is a PNAS Direct Submission. R.T.R. is a guest editor invited by the Editorial Board.

  • This article contains supporting information online at www.pnas.org/lookup/suppl/doi:10.1073/pnas.1310414110/-/DCSupplemental.

References

  1. ↵
    1. Adams EJ,
    2. Luoma AM
    (2013) The adaptable major histocompatibility complex (MHC) fold: Structure and function of nonclassical and MHC class I-like molecules. Annu Rev Immunol 31:529–561.
    OpenUrlCrossRefPubMed
  2. ↵
    1. Smith DP,
    2. Ashcroft AE,
    3. Radford SE
    (2010) Hemodialysis-related amyloidosis. Protein Misfolding Diseases: Current and Emerging Principles and Therapies, eds Ramirez-Alvarado M, Kelly JW, Dobson CM (John Wiley and Sons, Hoboken, NJ).
  3. ↵
    1. Myers SL,
    2. et al.
    (2006) A systematic study of the effect of physiological factors on β2-microglobulin amyloid formation at neutral pH. Biochemistry 45(7):2311–2321.
    OpenUrlCrossRefPubMed
  4. ↵
    1. Yamamoto S,
    2. et al.
    (2004) Glycosaminoglycans enhance the trifluoroethanol-induced extension of β 2-microglobulin-related amyloid fibrils at a neutral pH. J Am Soc Nephrol 15(1):126–133.
    OpenUrlAbstract/FREE Full Text
  5. ↵
    1. Rennella E,
    2. et al.
    (2010) Folding and fibrillogenesis: Clues from β2-microglobulin. J Mol Biol 401(2):286–297.
    OpenUrlCrossRefPubMed
  6. ↵
    1. Eakin CM,
    2. Berman AJ,
    3. Miranker AD
    (2006) A native to amyloidogenic transition regulated by a backbone trigger. Nat Struct Mol Biol 13(3):202–208.
    OpenUrlCrossRefPubMed
  7. ↵
    1. Calabrese MF,
    2. Eakin CM,
    3. Wang JM,
    4. Miranker AD
    (2008) A regulatable switch mediates self-association in an immunoglobulin fold. Nat Struct Mol Biol 15(9):965–971.
    OpenUrlCrossRefPubMed
  8. ↵
    1. Yamamoto S,
    2. et al.
    (2004) Low concentrations of sodium dodecyl sulfate induce the extension of β 2-microglobulin-related amyloid fibrils at a neutral pH. Biochemistry 43(34):11075–11082.
    OpenUrlCrossRefPubMed
  9. ↵
    1. McParland VJ,
    2. et al.
    (2000) Partially unfolded states of β(2)-microglobulin and amyloid formation in vitro. Biochemistry 39(30):8735–8746.
    OpenUrlCrossRefPubMed
  10. ↵
    1. Kihara M,
    2. et al.
    (2005) Seeding-dependent maturation of β2-microglobulin amyloid fibrils at neutral pH. J Biol Chem 280(12):12012–12018.
    OpenUrlAbstract/FREE Full Text
  11. ↵
    1. Eichner T,
    2. Radford SE
    (2011) Understanding the complex mechanisms of β2-microglobulin amyloid assembly. FEBS J 278(20):3868–3883.
    OpenUrlCrossRefPubMed
  12. ↵
    1. Trinh CH,
    2. Smith DP,
    3. Kalverda AP,
    4. Phillips SEV,
    5. Radford SE
    (2002) Crystal structure of monomeric human β-2-microglobulin reveals clues to its amyloidogenic properties. Proc Natl Acad Sci USA 99(15):9771–9776.
    OpenUrlAbstract/FREE Full Text
  13. ↵
    1. Jahn TR,
    2. Parker MJ,
    3. Homans SW,
    4. Radford SE
    (2006) Amyloid formation under physiological conditions proceeds via a native-like folding intermediate. Nat Struct Mol Biol 13(3):195–201.
    OpenUrlCrossRefPubMed
  14. ↵
    1. Barbet-Massin E,
    2. et al.
    (2010) Fibrillar vs crystalline full-length β-2-microglobulin studied by high-resolution solid-state NMR spectroscopy. J Am Chem Soc 132(16):5556–5557.
    OpenUrlCrossRefPubMed
  15. ↵
    1. Debelouchina GT,
    2. Platt GW,
    3. Bayro MJ,
    4. Radford SE,
    5. Griffin RG
    (2010) Magic angle spinning NMR analysis of β2-microglobulin amyloid fibrils in two distinct morphologies. J Am Chem Soc 132(30):10414–10423.
    OpenUrlCrossRefPubMed
  16. ↵
    Renner C el al. (2001) Fluoroprolines as tools for protein design and engineering. Angew Chem Int Ed 40(5):923–925.
  17. ↵
    1. Shoulders MD,
    2. Raines RT
    (2009) Collagen structure and stability. Annu Rev Biochem 78:929–958.
    OpenUrlCrossRefPubMed
  18. ↵
    1. Shoulders MD,
    2. Kamer KJ,
    3. Raines RT
    (2009) Origin of the stability conferred upon collagen by fluorination. Bioorg Med Chem Lett 19(14):3859–3862.
    OpenUrlCrossRefPubMed
  19. ↵
    1. Merkel L,
    2. Budisa N
    (2012) Organic fluorine as a polypeptide building element: In vivo expression of fluorinated peptides, proteins and proteomes. Org Biomol Chem 10(36):7241–7261.
    OpenUrlCrossRefPubMed
  20. ↵
    1. Bretscher LE,
    2. Jenkins CL,
    3. Taylor KM,
    4. DeRider ML,
    5. Raines RT
    (2001) Conformational stability of collagen relies on a stereoelectronic effect. J Am Chem Soc 123(4):777–778.
    OpenUrlCrossRefPubMed
  21. ↵
    1. Bang D,
    2. Kent SBH
    (2004) A one-pot total synthesis of crambin. Angew Chem Int Ed Engl 43(19):2534–2538.
    OpenUrlCrossRefPubMed
  22. ↵
    1. Schnölzer M,
    2. Alewood P,
    3. Jones A,
    4. Alewood D,
    5. Kent SBH
    (2007) In situ neutralization in Boc-chemistry solid-phase peptide synthesis – rapid, high yield assembly of difficult sequences. Int J Pept Res Ther 13(1–2):31–44.
    OpenUrlCrossRef
  23. ↵
    1. Eichner T,
    2. Radford SE
    (2009) A generic mechanism of β2-microglobulin amyloid assembly at neutral pH involving a specific proline switch. J Mol Biol 386(5):1312–1326.
    OpenUrlCrossRefPubMed
  24. ↵
    1. Colon W,
    2. Kelly JW
    (1992) Partial denaturation of transthyretin is sufficient for amyloid fibril formation in vitro. Biochemistry 31(36):8654–8660.
    OpenUrlCrossRefPubMed
  25. ↵
    1. Pan KM,
    2. et al.
    (1993) Conversion of α-helices into β-sheets features in the formation of the scrapie prion proteins. Proc Natl Acad Sci USA 90(23):10962–10966.
    OpenUrlAbstract/FREE Full Text
  26. ↵
    1. Kelly JW
    (1996) Alternative conformations of amyloidogenic proteins govern their behavior. Curr Opin Struct Biol 6(1):11–17.
    OpenUrlCrossRefPubMed
  27. ↵
    1. Uversky VN,
    2. Fink AL
    (2004) Conformational constraints for amyloid fibrillation: The importance of being unfolded. Biochim Biophys Acta 1698(2):131–153.
    OpenUrlCrossRefPubMed
  28. ↵
    1. Chatani E,
    2. et al.
    (2010) Pre-steady-state kinetic analysis of the elongation of amyloid fibrils of β(2)-microglobulin with tryptophan mutagenesis. J Mol Biol 400(5):1057–1066.
    OpenUrlCrossRefPubMed
  29. ↵
    1. Harper JD,
    2. Lansbury PT Jr.
    (1997) Models of amyloid seeding in Alzheimer’s disease and scrapie: Mechanistic truths and physiological consequences of the time-dependent solubility of amyloid proteins. Annu Rev Biochem 66:385–407.
    OpenUrlCrossRefPubMed
  30. ↵
    1. Chien P,
    2. Weissman JS,
    3. DePace AH
    (2004) Emerging principles of conformation-based prion inheritance. Annu Rev Biochem 73:617–656.
    OpenUrlCrossRefPubMed
  31. ↵
    1. Petkova AT,
    2. et al.
    (2005) Self-propagating, molecular-level polymorphism in Alzheimer’s β-amyloid fibrils. Science 307(5707):262–265.
    OpenUrlAbstract/FREE Full Text
  32. ↵
    1. Yanagi K,
    2. et al.
    (2012) The monomer-seed interaction mechanism in the formation of the β2-microglobulin amyloid fibril clarified by solution NMR techniques. J Mol Biol 422(3):390–402.
    OpenUrlCrossRefPubMed
  33. ↵
    1. Kameda A,
    2. et al.
    (2005) Nuclear magnetic resonance characterization of the refolding intermediate of β2-microglobulin trapped by non-native prolyl peptide bond. J Mol Biol 348(2):383–397.
    OpenUrlCrossRefPubMed
  34. ↵
    1. Sakata M,
    2. et al.
    (2008) Kinetic coupling of folding and prolyl isomerization of β2-microglobulin studied by mutational analysis. J Mol Biol 382(5):1242–1255.
    OpenUrlCrossRefPubMed
  35. ↵
    1. Chiti F,
    2. et al.
    (2001) A partially structured species of β 2-microglobulin is significantly populated under physiological conditions and involved in fibrillogenesis. J Biol Chem 276(50):46714–46721.
    OpenUrlAbstract/FREE Full Text
  36. ↵
    1. Tycko R,
    2. Wickner RB
    (2013) Molecular structures of amyloid and prion fibrils: Consensus versus controversy. Acc Chem Res 46(7):1487–1496.
    OpenUrlCrossRefPubMed
  37. ↵
    1. Chatani E,
    2. Yagi H,
    3. Naiki H,
    4. Goto Y
    (2012) Polymorphism of β2-microglobulin amyloid fibrils manifested by ultrasonication-enhanced fibril formation in trifluoroethanol. J Biol Chem 287(27):22827–22837.
    OpenUrlAbstract/FREE Full Text
  38. ↵
    1. Sarell CJ,
    2. et al.
    (2013) Expanding the repertoire of amyloid polymorphs by co-polymerization of related protein precursors. J Biol Chem 288(10):7327–7337.
    OpenUrlAbstract/FREE Full Text
  39. ↵
    1. Halfmann R,
    2. Lindquist S
    (2010) Epigenetics in the extreme: prions and the inheritance of environmentally acquired traits. Science 330(6004):629–632.
    OpenUrlAbstract/FREE Full Text
View Abstract
PreviousNext
Back to top
Article Alerts
Email Article

Thank you for your interest in spreading the word on PNAS.

NOTE: We only request your email address so that the person you are recommending the page to knows that you wanted them to see it, and that it is not junk mail. We do not capture any email address.

Enter multiple addresses on separate lines or separate them with commas.
Both the cis-trans equilibrium and isomerization dynamics of a single proline amide modulate β2-microglobulin amyloid assembly
(Your Name) has sent you a message from PNAS
(Your Name) thought you would like to see the PNAS web site.
Citation Tools
Misfolding and aggregation of β2-microglobulin
Vladimir Yu. Torbeev, Donald Hilvert
Proceedings of the National Academy of Sciences Dec 2013, 110 (50) 20051-20056; DOI: 10.1073/pnas.1310414110

Citation Manager Formats

  • BibTeX
  • Bookends
  • EasyBib
  • EndNote (tagged)
  • EndNote 8 (xml)
  • Medlars
  • Mendeley
  • Papers
  • RefWorks Tagged
  • Ref Manager
  • RIS
  • Zotero
Request Permissions
Share
Misfolding and aggregation of β2-microglobulin
Vladimir Yu. Torbeev, Donald Hilvert
Proceedings of the National Academy of Sciences Dec 2013, 110 (50) 20051-20056; DOI: 10.1073/pnas.1310414110
del.icio.us logo Digg logo Reddit logo Twitter logo CiteULike logo Facebook logo Google logo Mendeley logo
  • Tweet Widget
  • Facebook Like
  • Mendeley logo Mendeley
Proceedings of the National Academy of Sciences: 116 (7)
Current Issue

Submit

Sign up for Article Alerts

Jump to section

  • Article
    • Abstract
    • Results
    • Discussion
    • Materials and Methods
    • Acknowledgments
    • Footnotes
    • References
  • Figures & SI
  • Info & Metrics
  • PDF

You May Also be Interested in

Several aspects of the proposal, which aims to expand open access, require serious discussion and, in some cases, a rethink.
Opinion: “Plan S” falls short for society publishers—and for the researchers they serve
Several aspects of the proposal, which aims to expand open access, require serious discussion and, in some cases, a rethink.
Image credit: Dave Cutler (artist).
Several large or long-lived animals seem strangely resistant to developing cancer. Elucidating the reasons why could lead to promising cancer-fighting strategies in humans.
Core Concept: Solving Peto’s Paradox to better understand cancer
Several large or long-lived animals seem strangely resistant to developing cancer. Elucidating the reasons why could lead to promising cancer-fighting strategies in humans.
Image credit: Shutterstock.com/ronnybas frimages.
Featured Profile
PNAS Profile of NAS member and biochemist Hao Wu
 Nonmonogamous strawberry poison frog (Oophaga pumilio).  Image courtesy of Yusan Yang (University of Pittsburgh, Pittsburgh).
Putative signature of monogamy
A study suggests a putative gene-expression hallmark common to monogamous male vertebrates of some species, namely cichlid fishes, dendrobatid frogs, passeroid songbirds, common voles, and deer mice, and identifies 24 candidate genes potentially associated with monogamy.
Image courtesy of Yusan Yang (University of Pittsburgh, Pittsburgh).
Active lifestyles. Image courtesy of Pixabay/MabelAmber.
Meaningful life tied to healthy aging
Physical and social well-being in old age are linked to self-assessments of life worth, and a spectrum of behavioral, economic, health, and social variables may influence whether aging individuals believe they are leading meaningful lives.
Image courtesy of Pixabay/MabelAmber.

More Articles of This Classification

Biological Sciences

  • Structural basis for activity of TRIC counter-ion channels in calcium release
  • PGC1A regulates the IRS1:IRS2 ratio during fasting to influence hepatic metabolism downstream of insulin
  • Altered neural odometry in the vertical dimension
Show more

Biochemistry

  • Opinion: “Plan S” falls short for society publishers—and for the researchers they serve
  • Peptidic degron for IMiD-induced degradation of heterologous proteins
  • Structure-based development of new RAS-effector inhibitors from a combination of active and inactive RAS-binding compounds
Show more

Physical Sciences

  • Deep elastic strain engineering of bandgap through machine learning
  • Single-molecule excitation–emission spectroscopy
  • Microscopic description of acid–base equilibrium
Show more

Chemistry

  • Deep elastic strain engineering of bandgap through machine learning
  • Single-molecule excitation–emission spectroscopy
  • Microscopic description of acid–base equilibrium
Show more

Related Content

  • No related articles found.
  • Scopus
  • PubMed
  • Google Scholar

Cited by...

  • C-terminal sequence of amyloid-resistant type F apolipoprotein A-II inhibits amyloid fibril formation of apolipoprotein A-II in mice
  • Scopus (24)
  • Google Scholar

Similar Articles

Site Logo
Powered by HighWire
  • Submit Manuscript
  • Twitter
  • Facebook
  • RSS Feeds
  • Email Alerts

Articles

  • Current Issue
  • Latest Articles
  • Archive

PNAS Portals

  • Classics
  • Front Matter
  • Teaching Resources
  • Anthropology
  • Chemistry
  • Physics
  • Sustainability Science

Information

  • Authors
  • Editorial Board
  • Reviewers
  • Press
  • Site Map

Feedback    Privacy/Legal

Copyright © 2019 National Academy of Sciences. Online ISSN 1091-6490