Your browser doesn't support javascript.
loading
Show: 20 | 50 | 100
Results 1 - 20 de 22
Filter
Add more filters










Publication year range
1.
Chembiochem ; 25(7): e202300827, 2024 Apr 02.
Article in English | MEDLINE | ID: mdl-38349283

ABSTRACT

We describe six compounds as early hits for the development of direct inhibitors of KRAS, an important anticancer drug target. We show that these compounds bind to KRAS with affinities in the low micromolar range and exert different effects on its interactions with binding partners. Some of the compounds exhibit selective binding to the activated form of KRAS and inhibit signal transduction through both the MAPK or the phosphatidylinositide 3-kinase PI3K-protein kinase B (AKT) pathway in cells expressing mutant KRAS. Most inhibit intrinsic and/or SOS-mediated KRAS activation while others inhibit RAS-effector interaction. We propose these compounds as starting points for the development of non-covalent allosteric KRAS inhibitors.


Subject(s)
Antineoplastic Agents , Proto-Oncogene Proteins p21(ras) , Proto-Oncogene Proteins p21(ras)/genetics , Mutation , Cell Line, Tumor , Signal Transduction , Antineoplastic Agents/pharmacology
2.
ACS Omega ; 8(34): 31419-31426, 2023 Aug 29.
Article in English | MEDLINE | ID: mdl-37663463

ABSTRACT

Mutations in KRAS account for about 20% of human cancers. Despite the major progress in recent years toward the development of KRAS inhibitors, including the discovery of covalent inhibitors of the G12C KRAS variant for the treatment of non-small-cell lung cancer, much work remains to be done to discover broad-acting inhibitors to treat many other KRAS-driven cancers. In a previous report, we showed that a 308.4 Da small-molecule ligand [(2R)-2-(N'-(1H-indole-3-carbonyl)hydrazino)-2-phenyl-acetamide] binds to KRAS with low micro-molar affinity [Chem. Biol. Drug Des.2019; 94(2):1441-1456]. Binding of this ligand, which we call ACA22, to the p1 pocket of KRAS and its interactions with residues at beta-strand 1 and the switch loops have been supported by data from nuclear magnetic resonance spectroscopy and microscale thermophoresis experiments. However, the inhibitory potential of the compound was not demonstrated. Here, we show that ACA22 inhibits KRAS-mediated signal transduction in cells expressing wild type (WT) and G12D mutant KRAS and reduces levels of guanosine triphosphate-loaded WT KRAS more effectively than G12D KRAS. We ruled out the direct effect on nucleotide exchange or effector binding as possible mechanisms of inhibition using a variety of biophysical assays. Combining these observations with binding data that show comparable affinities of the compound for the active and inactive forms of the mutant but not the WT, we propose conformational selection as a possible mechanism of action of ACA22.

3.
ACS Bio Med Chem Au ; 2(6): 617-626, 2022 Dec 21.
Article in English | MEDLINE | ID: mdl-37101428

ABSTRACT

We describe a small molecule ligand ACA-14 (2-hydroxy-5-{[(2-phenylcyclopropyl) carbonyl] amino} benzoic acid) as an initial lead for the development of direct inhibitors of KRAS, a notoriously difficult anticancer drug target. We show that the compound binds to KRAS near the switch regions with affinities in the low micromolar range and exerts different effects on KRAS interactions with binding partners. Specifically, ACA-14 impedes the interaction of KRAS with its effector Raf and reduces both intrinsic and SOS-mediated nucleotide exchange rates. Likely as a result of these effects, ACA-14 inhibits signal transduction through the MAPK pathway in cells expressing mutant KRAS and inhibits the growth of pancreatic and colon cancer cells harboring mutant KRAS. We thus propose compound ACA-14 as a useful initial lead for the development of broad-acting inhibitors that target multiple KRAS mutants and simultaneously deplete the fraction of GTP-loaded KRAS while abrogating the effector-binding ability of the already GTP-loaded fraction.

4.
Sci Rep ; 9(1): 17303, 2019 11 21.
Article in English | MEDLINE | ID: mdl-31754129

ABSTRACT

Histatin-5 (Hst-5) is an antimicrobial, salivary protein that is involved in the host defense system. Hst-5 has been proposed to bind functionally relevant zinc and copper but presents challenges in structural studies due to its disordered conformation in aqueous solution. Here, we used circular dichroism (CD) and UV resonance Raman (UVRR) spectroscopy to define metallo-Hst-5 interactions in aqueous solution. A zinc-containing Hst-5 sample exhibits shifted Raman bands, relative to bands observed in the absence of zinc. Based on comparison to model compounds and to a family of designed, zinc-binding beta hairpins, the alterations in the Hst-5 UVRR spectrum are attributed to zinc coordination by imidazole side chains. Zinc addition also shifted a tyrosine aromatic ring UVRR band through an electrostatic interaction. Copper addition did not have these effects. A sequence variant, H18A/H19A, was employed; this mutant has less potent antifungal activity, when compared to Hst-5. Zinc addition had only a small effect on the thermal stability of this mutant. Interestingly, both zinc and copper addition shifted histidine UVRR bands in a manner diagnostic for metal coordination. Results obtained with a K13E/R22G mutant were similar to those obtained with wildtype. These experiments show that H18 and H19 contribute to a zinc binding site. In the H18A/H19A mutant the specificity of the copper/zinc binding sites is lost. The experiments implicate specific zinc binding to be important in the antimicrobial activity of Hst-5.


Subject(s)
Anti-Infective Agents/pharmacology , Histatins/pharmacology , Intrinsically Disordered Proteins/pharmacology , Anti-Infective Agents/chemistry , Anti-Infective Agents/metabolism , Binding Sites/genetics , Circular Dichroism , Copper/metabolism , Histatins/chemistry , Histatins/genetics , Histatins/metabolism , Intrinsically Disordered Proteins/chemistry , Intrinsically Disordered Proteins/genetics , Intrinsically Disordered Proteins/metabolism , Mutation , Protein Binding/genetics , Spectrum Analysis, Raman , Zinc/metabolism
5.
Chem Commun (Camb) ; 55(63): 9399-9402, 2019 Aug 14.
Article in English | MEDLINE | ID: mdl-31322154

ABSTRACT

Tyrosine residues act as intermediates in proton coupled electron transfer reactions (PCET) in proteins. For example, in ribonucleotide reductase (RNR), a tyrosyl radical oxidizes an active site cysteine via a 35 Å pathway that contains multiple aromatic groups. When singlet tyrosine is oxidized, the radical becomes a strong acid, and proton transfer reactions, which are coupled with the redox reaction, may be used to control reaction rate. Here, we characterize a tyrosine-containing beta hairpin, Peptide O, which has a cross-strand, noncovalent interaction between its single tyrosine, Y5, and a cysteine (C14). Circular dichroism provides evidence for a thermostable beta-turn. EPR spectroscopy shows that Peptide O forms a neutral tyrosyl radical after UV photolysis at 160 K. Molecular dynamics simulations support a phenolic/SH interaction in the tyrosine singlet and radical states. Differential pulse voltammetry exhibits pH dependence consistent with the formation of a neutral tyrosyl radical and a pKa change in two other residues. A redox-coupled decrease in cysteine pKa from 9 (singlet) to 6.9 (radical) is assigned. At pD 11, picosecond transient absorption spectroscopy after UV photolysis monitors tyrosyl radical recombination via electron transfer (ET). The ET rate in Peptide O is indistinguishable from the ET rates observed in peptides containing a histidine and a cyclohexylalanine (Cha) at position 14. However, at pD 9, the tyrosyl radical decays via PCET, and the decay rate is slowed, when compared to the histidine 14 variant. Notably, the decay rate is accelerated, when compared to the Cha 14 variant. We conclude that redox coupling between tyrosine and cysteine can act as a PCET control mechanism in proteins.

6.
ACS Omega ; 4(2): 2921-2930, 2019 Feb 28.
Article in English | MEDLINE | ID: mdl-30842983

ABSTRACT

Approximately 15% of all human tumors harbor mutant KRAS, a membrane-associated small GTPase and notorious oncogene. Mutations that render KRAS constitutively active will lead to uncontrolled cell growth and cancer. However, despite aggressive efforts in recent years, there are no drugs on the market that directly target KRAS and inhibit its aberrant functions. In the current work, we combined structure-based design with a battery of cell and biophysical assays to discover a novel pyrazolopyrimidine-based allosteric KRAS inhibitor that binds to activated KRAS with sub-micromolar affinity and disrupts effector binding, thereby inhibiting KRAS signaling and cancer cell growth. These results show that pyrazolopyrimidine-based compounds may represent a first-in-class allosteric noncovalent inhibitors of KRAS. Moreover, by studying two of its analogues, we identified key chemical features of the compound that interact with a set of specific residues at the switch regions of KRAS and play critical roles for its high-affinity binding and unique mode of action, thus providing a blueprint for future optimization efforts.

7.
J Phys Chem B ; 123(13): 2780-2791, 2019 04 04.
Article in English | MEDLINE | ID: mdl-30888824

ABSTRACT

Tyrosine-tryptophan (YW) dyads are ubiquitous structural motifs in enzymes and play roles in proton-coupled electron transfer (PCET) and, possibly, protection from oxidative stress. Here, we describe the function of YW dyads in de novo designed 18-mer, ß hairpins. In Peptide M, a YW dyad is formed between W14 and Y5. A UV hypochromic effect and an excitonic Cotton signal are observed, in addition to singlet, excited state (W*) and fluorescence emission spectral shifts. In a second Peptide, Peptide MW, a Y5-W13 dyad is formed diagonally across the strand and distorts the backbone. On a picosecond timescale, the W* excited-state decay kinetics are similar in all peptides but are accelerated relative to amino acids in solution. In Peptide MW, the W* spectrum is consistent with increased conformational flexibility. In Peptide M and MW, the electron paramagnetic resonance spectra obtained after UV photolysis are characteristic of tyrosine and tryptophan radicals at 160 K. Notably, at pH 9, the radical photolysis yield is decreased in Peptide M and MW, compared to that in a tyrosine and tryptophan mixture. This protective effect is not observed at pH 11 and is not observed in peptides containing a tryptophan-histidine dyad or tryptophan alone. The YW dyad protective effect is attributed to an increase in the radical recombination rate. This increase in rate can be facilitated by hydrogen-bonding interactions, which lower the barrier for the PCET reaction at pH 9. These results suggest that the YW dyad structural motif promotes radical quenching under conditions of reactive oxygen stress.


Subject(s)
Biomimetic Materials , Tryptophan , Tyrosine , Biomimetic Materials/chemistry , Biomimetic Materials/metabolism , Hydrogen-Ion Concentration , Protein Conformation , Tryptophan/chemistry , Tryptophan/metabolism , Tyrosine/chemistry , Tyrosine/metabolism
8.
Chem Biol Drug Des ; 94(2): 1441-1456, 2019 08.
Article in English | MEDLINE | ID: mdl-30903639

ABSTRACT

RAS mutations account for >15% of all human tumors, and of these ~85% are due to mutations in a particular RAS gene: KRAS. Recent studies revealed that KRAS harbors four druggable allosteric sites. Here, we have (a) used molecular simulations to generate ensembles of wild type and four major oncogenic KRAS mutants (G12V, G12D, G13D, and Q61H); (b) characterized the druggability of each allosteric pocket in each protein; (c) conducted extensive ensemble-based virtual screening using pocket-tailored ligand libraries; (d) prioritized hits through hierarchical postdocking analysis; and (e) validated predicted hits with NMR. Of the 785 diverse potential hits identified by our in silico analysis, we tested 90 for their ability to bind KRAS using NMR and found that nine cause backbone amide chemical shift perturbations of residues near the functionally responsive switch loops, suggesting potential binding. We conducted detailed biophysical analyses on a novel indole-based compound to demonstrate the potential of our workflow to yield lead compounds. We believe the detailed information documented in this work regarding the druggability profile of each allosteric site and the chemical fingerprints of compounds that target them will serve as vital resources for future structure-based drug design efforts against KRAS, a high-value target for cancer therapy.


Subject(s)
Antineoplastic Agents/chemistry , Enzyme Inhibitors/chemistry , Mutation, Missense , Proto-Oncogene Proteins p21(ras)/antagonists & inhibitors , Proto-Oncogene Proteins p21(ras)/chemistry , Amino Acid Substitution , Antineoplastic Agents/therapeutic use , Drug Screening Assays, Antitumor , Enzyme Inhibitors/therapeutic use , Humans , Neoplasms/drug therapy , Neoplasms/enzymology , Neoplasms/genetics , Nuclear Magnetic Resonance, Biomolecular , Proto-Oncogene Proteins p21(ras)/genetics , Proto-Oncogene Proteins p21(ras)/metabolism
9.
J Phys Chem B ; 121(15): 3536-3545, 2017 04 20.
Article in English | MEDLINE | ID: mdl-28145121

ABSTRACT

Tyrosine-based radical transfer plays an important role in photosynthesis, respiration, and DNA synthesis. Radical transfer can occur either by electron transfer (ET) or proton coupled electron transfer (PCET), depending on the pH. Reversible conformational changes in the surrounding protein matrix may control reactivity of radical intermediates. De novo designed Peptide A is a synthetic 18 amino-acid ß-hairpin, which contains a single tyrosine (Y5) and carries out a kinetically significant PCET reaction between Y5 and a cross-strand histidine (H14). In Peptide A, amide II' (CN) changes are observed in the UV resonance Raman (UVRR) spectrum, associated with tyrosine ET and PCET; these bands were attributed previously to a reversible change in secondary structure. Here, we use molecular dynamics simulations to define this conformational change in Peptide A and its H14-to-cyclohexylalanine variant, Peptide C. Three different Y5 charge states, tyrosine (YH), tyrosinate (Y-), and neutral tyrosyl radical (Y·), are considered. The simulations show that Peptide A-YH and A-Y- retain secondary structure and noncovalent interactions, whereas A-Y· is unstable. In contrast, both Peptide C-Y- and Peptide C-Y· are unstable, due to the loss of the Y5-H14 π-π interaction. These simulations are consistent with previous UVRR experimental results on the two ß-hairpins. Furthermore, they demonstrate the ability of simulations using fixed-charge force fields to accurately capture redox-linked conformational dynamics in a ß-strand peptide.


Subject(s)
Molecular Dynamics Simulation , Photosystem II Protein Complex/chemistry , Electron Transport , Molecular Structure , Oxidation-Reduction , Photosystem II Protein Complex/metabolism , Protein Conformation , Spectrum Analysis
10.
J Phys Chem B ; 120(7): 1259-72, 2016 Feb 25.
Article in English | MEDLINE | ID: mdl-26886811

ABSTRACT

Photosystem II (PSII) and ribonucleotide reductase employ oxidation and reduction of the tyrosine aromatic ring in radical transport pathways. Tyrosine-based reactions involve either proton-coupled electron transfer (PCET) or electron transfer (ET) alone, depending on the pH and the pKa of tyrosine's phenolic oxygen. In PSII, a subset of the PCET reactions are mediated by a tyrosine-histidine redox-driven proton relay, YD-His189. Peptide A is a PSII-inspired ß-hairpin, which contains a single tyrosine (Y5) and histidine (H14). Previous electrochemical characterization indicated that Peptide A conducts a net PCET reaction between Y5 and H14, which have a cross-strand π-π interaction. The kinetic impact of H14 has not yet been explored. Here, we address this question through time-resolved absorption spectroscopy and 280-nm photolysis, which generates a neutral tyrosyl radical. The formation and decay of the neutral tyrosyl radical at 410 nm were monitored in Peptide A and its variant, Peptide C, in which H14 is replaced by cyclohexylalanine (Cha14). Significantly, both electron transfer (ET, pL 11, L = lyonium) and PCET (pL 9) were accelerated in Peptide A and C, compared to model tyrosinate or tyrosine at the same pL. Increased electronic coupling, mediated by the peptide backbone, can account for this rate acceleration. Deuterium exchange gave no significant solvent isotope effect in the peptides. At pL 9, but not at pL 11, the reaction rate decreased when H14 was mutated to Cha14. This decrease in rate is attributed to an increase in reorganization energy in the Cha14 mutant. The Y5-H14 mechanism in Peptide A is reminiscent of proton- and electron-transfer events involving YD-H189 in PSII. These results document a mechanism by which proton donors and acceptors can regulate the rate of PCET reactions.


Subject(s)
Histidine/metabolism , Peptides/metabolism , Photosystem II Protein Complex/metabolism , Tyrosine/metabolism , Amino Acid Sequence , Electron Transport , Electrons , Histidine/chemistry , Kinetics , Models, Molecular , Molecular Sequence Data , Peptides/chemistry , Photosystem II Protein Complex/chemistry , Protons , Tyrosine/analogs & derivatives , Tyrosine/chemistry
11.
Nat Commun ; 6: 10010, 2015 Dec 02.
Article in English | MEDLINE | ID: mdl-26627888

ABSTRACT

In class 1a ribonucleotide reductase (RNR), a substrate-based radical is generated in the α2 subunit by long-distance electron transfer involving an essential tyrosyl radical (Y122O·) in the ß2 subunit. The conserved W48 ß2 is ∼10 Å from Y122OH; mutations at W48 inactivate RNR. Here, we design a beta hairpin peptide, which contains such an interacting tyrosine-tryptophan dyad. The NMR structure of the peptide establishes that there is no direct hydrogen bond between the phenol and the indole rings. However, electronic coupling between the tyrosine and tryptophan occurs in the peptide. In addition, downshifted ultraviolet resonance Raman (UVRR) frequencies are observed for the radical state, reproducing spectral downshifts observed for ß2. The frequency downshifts of the ring and CO bands are consistent with charge transfer from YO· to W or another residue. Such a charge transfer mechanism implies a role for the ß2 Y-W dyad in electron transfer.


Subject(s)
Escherichia coli Proteins/chemistry , Escherichia coli/enzymology , Exoribonucleases/chemistry , Tryptophan/chemistry , Tyrosine/chemistry , Amino Acid Motifs , Electron Transport , Escherichia coli/chemistry , Escherichia coli Proteins/genetics , Escherichia coli Proteins/metabolism , Exoribonucleases/genetics , Exoribonucleases/metabolism , Hydrogen Bonding , Models, Molecular , Tryptophan/metabolism , Tyrosine/metabolism
12.
Biochim Biophys Acta ; 1847(6-7): 558-64, 2015.
Article in English | MEDLINE | ID: mdl-25791219

ABSTRACT

In photosynthesis, photosystem II (PSII) harvests sunlight with bound pigments to oxidize water and reduce quinone to quinol, which serves as electron and proton mediators for solar-to-chemical energy conversion. At least two types of quinone cofactors in PSII are redox-linked: QA, and QB. Here, we for the first time apply 257-nm ultraviolet resonance Raman (UVRR) spectroscopy to acquire the molecular vibrations of plastoquinone (PQ) in PSII membranes. Owing to the resonance enhancement effect, the vibrational signal of PQ in PSII membranes is prominent. A strong band at 1661 cm(-1) is assigned to ring CC/CO symmetric stretch mode (ν8a mode) of PQ, and a weak band at 469 cm(-1) to ring stretch mode. By using a pump-probe difference UVRR method and a sample jet technique, the signals of QA and QB can be distinguished. A frequency difference of 1.4 cm(-1) in ν8a vibrational mode between QA and QB is observed, corresponding to ~86 mV redox potential difference imposed by their protein environment. In addition, there are other PQs in the PSII membranes. A negligible anharmonicity effect on their combination band at 2130 cm(-1) suggests that the 'other PQs' are situated in a hydrophobic environment. The detection of the 'other PQs' might be consistent with the view that another functional PQ cofactor (not QA or QB) exists in PSII. This UVRR approach will be useful to the study of quinone molecules in photosynthesis or other biological systems.


Subject(s)
Cell Membrane/metabolism , Photosystem II Protein Complex/chemistry , Quinones/chemistry , Spectrophotometry, Ultraviolet , Spectrum Analysis, Raman/methods , Spinacia oleracea/metabolism , Chlorophyll/chemistry , Electron Transport , Oxidation-Reduction , Photosynthesis/physiology , Photosystem II Protein Complex/metabolism , Quinones/metabolism , Spinacia oleracea/chemistry , Vibration
13.
J Phys Chem B ; 119(6): 2726-36, 2015 Feb 12.
Article in English | MEDLINE | ID: mdl-25437178

ABSTRACT

In proteins, proton-coupled electron transfer (PCET) can involve the transient oxidation and reduction of the aromatic amino acid tyrosine. Due to the short life time of tyrosyl radical intermediates, transient absorption spectroscopy provides an important tool in deciphering electron-transfer mechanisms. In this report, the photoionization of solution tyrosine and tyrosinate was investigated using transient, picosecond absorption spectroscopy. The results were compared to data acquired from a tyrosine-containing ß-hairpin peptide. This maquette, peptide A, is an 18-mer that exhibits π-π interaction between tyrosine (Y5) and histidine (H14). Y5 and H14 carry out an orthogonal PCET reaction when Y5 is oxidized in the mid-pH range. Photolysis of all samples (280 nm, instrument response: 360 fs) generated a solvated electron signal within 3 ps. A signal from the S1 state and a 410 nm signal from the neutral tyrosyl radical were also formed in 3 ps. Fits to S1 and tyrosyl radical decay profiles revealed biphasic kinetics with time constants of 10-50 and 400-1300 ps. The PCET reaction at pH 9 was associated with a significant decrease in the rate of tyrosyl radical and S1 decay compared to electron transfer (ET) alone (pH 11). This pH dependence was observed both in solution and peptide samples. The pH 9 reaction may occur with a sequential electron-transfer, proton-transfer (ETPT) mechanism. Alternatively, the pH 9 reaction may occur by coupled proton and electron transfer (CPET). CPET would be associated with a reorganization energy larger than that of the pH 11 reaction. Significantly, the decay kinetics of S1 and the tyrosyl radical were accelerated in peptide A compared to solution samples at both pH values. These data suggest either an increase in electronic coupling or a specific, sequence-dependent interaction, which facilitates ET and PCET in the ß hairpin.


Subject(s)
Oncogene Protein pp60(v-src)/chemistry , Peptide Fragments/chemistry , Protons , Tyrosine/chemistry , Electron Transport , Kinetics , Models, Molecular , Photolysis , Protein Structure, Secondary
14.
J Phys Chem B ; 118(11): 2993-3004, 2014 Mar 20.
Article in English | MEDLINE | ID: mdl-24606240

ABSTRACT

Ribonucleotide reductase (RNR) catalyzes the production of deoxyribonucleotides in all cells. In E. coli class Ia RNR, a transient α2ß2 complex forms when a ribonucleotide substrate, such as CDP, binds to the α2 subunit. A tyrosyl radical (Y122O•)-diferric cofactor in ß2 initiates substrate reduction in α2 via a long-distance, proton-coupled electron transfer (PCET) process. Here, we use reaction-induced FT-IR spectroscopy to describe the α2ß2 structural landscapes, which are associated with dATP and hydroxyurea (HU) inhibition. Spectra were acquired after mixing E. coli α2 and ß2 with a substrate, CDP, and the allosteric effector, ATP. Isotopic chimeras, (13)Cα2ß2 and α2(13)Cß2, were used to define subunit-specific structural changes. Mixing of α2 and ß2 under turnover conditions yielded amide I (C═O) and II (CN/NH) bands, derived from each subunit. The addition of the inhibitor, dATP, resulted in a decreased contribution from amide I bands, attributable to ß strands and disordered structures. Significantly, HU-mediated reduction of Y122O• was associated with structural changes in α2, as well as ß2. To define the spectral contributions of Y122O•/Y122OH in the quaternary complex, (2)H4 labeling of ß2 tyrosines and HU editing were performed. The bands of Y122O•, Y122OH, and D84, a unidentate ligand to the diferric cluster, previously identified in isolated ß2, were observed in the α2ß2 complex. These spectra also provide evidence for a conformational rearrangement at an additional ß2 tyrosine(s), Yx, in the α2ß2/CDP/ATP complex. This study illustrates the utility of reaction-induced FT-IR spectroscopy in the study of complex enzymes.


Subject(s)
Escherichia coli/enzymology , Ribonucleotide Reductases/chemistry , Ribonucleotide Reductases/metabolism , Models, Molecular , Oxidation-Reduction , Protein Structure, Secondary , Spectroscopy, Fourier Transform Infrared
15.
ACS Chem Biol ; 9(4): 891-6, 2014 Apr 18.
Article in English | MEDLINE | ID: mdl-24437616

ABSTRACT

In photosystem II (PSII), water is oxidized at the oxygen-evolving complex. This process occurs through a light-induced cycle that produces oxygen and protons. While coupled proton and electron transfer reactions play an important role in PSII and other proteins, direct detection of internal proton transfer reactions is challenging. Here, we demonstrate that the unnatural amino acid, 7-azatryptophan (7AW), has unique, pH-sensitive vibrational frequencies, which are sensitive markers of proton transfer. The intrinsically disordered, PSII subunit, PsbO, which contains a single W residue (Trp241), was engineered to contain 7AW at position 241. Fluorescence shows that 7AW-241 is buried in a hydrophobic environment. Reconstitution of 7AW(241)PsbO to PSII had no significant impact on oxygen evolution activity or flash-dependent protein dynamics. We conclude that directed substitution of 7AW into other structural domains is likely to provide a nonperturbative spectroscopic probe, which can be used to define internal proton pathways in PsbO.


Subject(s)
Amino Acids/chemistry , Oxygen , Photosystem II Protein Complex/chemistry , Tryptophan/analogs & derivatives , Hydrogen Bonding , Hydrophobic and Hydrophilic Interactions , Molecular Structure , Spectroscopy, Fourier Transform Infrared , Tryptophan/chemistry , Water/chemistry
16.
J Phys Chem B ; 116(35): 10590-9, 2012 Sep 06.
Article in English | MEDLINE | ID: mdl-22860514

ABSTRACT

Long-distance electron transfer (ET) plays a critical role in solar energy conversion, DNA synthesis, and mitochondrial respiration. Tyrosine (Y) side chains can function as intermediates in these reactions. The oxidized form of tyrosine deprotonates to form a neutral tyrosyl radical, Y(•), a powerful oxidant. In photosystem II (PSII) and ribonucleotide reductase, redox-active tyrosines are involved in the proton-coupled electron transfer (PCET) reactions, which are key in catalysis. In these proteins, redox-linked structural dynamics may play a role in controlling the radical's extraordinary oxidizing power. To define these dynamics in a structurally tractable system, we have constructed biomimetic peptide maquettes, which are inspired by PSII. UV resonance Raman studies were conducted of ET and PCET reactions in these ß-hairpins, which contain a single tyrosine residue. At pH 11, UV photolysis induces ET from the deprotonated phenolate side chain to solvent. At pH 8.5, interstrand proton transfer to a π-stacked histidine accompanies the Y oxidation reaction. The UV resonance Raman difference spectrum, associated with Y oxidation, was obtained from the peptide maquettes in D(2)O buffers. The difference spectra exhibited bands at 1441 and 1472 cm(-1), which are assigned to the amide II' (CN) vibration of the ß-hairpin. This amide II' spectral change was attributed to substantial alterations in amide hydrogen bonding, which are coupled with the Y/Y(•) redox reaction and are reversible. These experiments show that ET and PCET reactions can create new minima in the protein conformational landscape. This work suggests that charge-coupled conformational changes can occur in complex proteins that contain redox-active tyrosines. These redox-linked dynamics could play an important role in control of PCET in biological oxygen evolution, respiration, and DNA synthesis.


Subject(s)
Biomimetic Materials/chemistry , Peptides/chemistry , Photosystem II Protein Complex/chemistry , Amino Acid Sequence , Biomimetic Materials/metabolism , Deuterium Oxide/chemistry , Electron Transport , Hydrogen-Ion Concentration , Molecular Sequence Data , Oxidation-Reduction , Peptides/metabolism , Photolysis , Photosystem II Protein Complex/metabolism , Protein Structure, Secondary , Protons , Solvents/chemistry , Spectrum Analysis, Raman , Tyrosine/chemistry
17.
Biochemistry ; 51(7): 1431-8, 2012 Feb 21.
Article in English | MEDLINE | ID: mdl-22250969

ABSTRACT

Metal substitution of heme proteins is widely applied in the study of biologically relevant electron transfer (ET) reactions. It has been shown that many modified proteins remain in their native conformation and can provide useful insights into the molecular mechanism of electron transfer between the native protein and its substrates. We investigated ET reactions between zinc-substituted cytochrome P450(cam) and small organic compounds such as quinones and ferrocene, which are capable of accessing the protein's hydrophobic channel and binding close to the active site, like its native substrate, camphor. Following the substitution method developed by Gunsalus and co-workers [Wagner, G. C., et al. (1981) J. Biol. Chem. 256, 6262-6265], we have identified two dominant forms of the zinc-substituted protein, F450 and F420, that exhibit different photophysical and photochemical properties. The ET behavior of F420 suggests that hydrophobic redox-active ligands are able to penetrate the hydrophobic channel and place themselves in the direct vicinity of the Zn-porphyrin. In contrast, the slower ET quenching rates observed in the case of F450 indicate that the association is weak and occurs outside of the protein channel. Therefore, we conclude that F420 corresponds to the open structure of the native cytochrome P450(cam) while F450 has a closed or partially closed channel that is characteristic of the camphor-containing cytochrome P450(cam). The existence of two distinct conformers of Zn-bound P450(cam) is consistent with the findings of Goodin and co-workers [Lee, Y.-T., et al. (2010) Biochemistry 49, 3412-3419] and has significant consequences for future electron transfer studies on this popular metalloenzyme.


Subject(s)
Camphor 5-Monooxygenase/chemistry , Cytochromes/chemistry , Riboflavin/analogs & derivatives , Zinc/chemistry , Biochemistry/methods , Catalytic Domain , Electron Transport , Escherichia coli/metabolism , Heme/chemistry , Hydrogen-Ion Concentration , Kinetics , Ligands , Photochemistry/methods , Protein Binding , Protein Conformation , Riboflavin/chemistry , Spectrometry, Fluorescence/methods , Spectrophotometry/methods
18.
J Vis Exp ; (51)2011 May 18.
Article in English | MEDLINE | ID: mdl-21633329

ABSTRACT

Raman spectroscopy is often plagued by a strong fluorescent background, particularly for biological samples. If a sample is excited with a train of ultrafast pulses, a system that can temporally separate spectrally overlapping signals on a picosecond timescale can isolate promptly arriving Raman scattered light from late-arriving fluorescence light. Here we discuss the construction and operation of a complex nonlinear optical system that uses all-optical switching in the form of a low-power optical Kerr gate to isolate Raman and fluorescence signals. A single 808 nm laser with 2.4 W of average power and 80 MHz repetition rate is split, with approximately 200 mW of 808 nm light being converted to < 5 mW of 404 nm light sent to the sample to excite Raman scattering. The remaining unconverted 808 nm light is then sent to a nonlinear medium where it acts as the pump for the all-optical shutter. The shutter opens and closes in 800 fs with a peak efficiency of approximately 5%. Using this system we are able to successfully separate Raman and fluorescence signals at an 80 MHz repetition rate using pulse energies and average powers that remain biologically safe. Because the system has no spare capacity in terms of optical power, we detail several design and alignment considerations that aid in maximizing the throughput of the system. We also discuss our protocol for obtaining the spatial and temporal overlap of the signal and pump beams within the Kerr medium, as well as a detailed protocol for spectral acquisition. Finally, we report a few representative results of Raman spectra obtained in the presence of strong fluorescence using our time-gating system.


Subject(s)
Microspectrophotometry/methods , Spectrum Analysis, Raman/methods , Fluorescence , Microspectrophotometry/instrumentation , Spectrum Analysis, Raman/instrumentation
19.
J Am Chem Soc ; 132(46): 16423-31, 2010 Nov 24.
Article in English | MEDLINE | ID: mdl-21038913

ABSTRACT

A water-soluble octacarboxyhemicarcerand was used as a shuttle to transport redox-active substrates across the aqueous medium and deliver them to the target protein. The results show that weak multivalent interactions and conformational flexibility can be exploited to reversibly bind complex supramolecular assemblies to biological molecules. Hydrophobic electron donors and acceptors were encapsulated within the hemicarcerand, and photoinduced electron transfer (ET) between the Zn-substituted cytochrome c (MW = 12.3 kD) and the host-guest complexes (MW = 2.2 kD) was used to probe the association between the negatively charged hemicarceplex and the positively charged protein. The behavior of the resulting ternary protein-hemicarcerand-guest assembly was investigated in two binding limits: (1) when K(encaps) ≫ K(assoc), the hemicarcerand transports the ligand to the protein while protecting it from the aqueous medium; and (2) when K(assoc) > K(encaps), the hemicarcerand-protein complex is formed first, and the hemicarcerand acts as an artificial receptor site that intercepts ligands from solution and positions them close to the active site of the metalloenzyme. In both cases, ET mediated by the protein-bound hemicarcerand is much faster than that due to diffusional encounters with the respective free donor or acceptor in solution. The measured ET rates suggest that the dominant binding region of the host-guest complex on the surface of the protein is consistent with the docking area of the native redox partner of cytochrome c. The strong association with the protein is attributed to the flexible conformation and adaptable charge distribution of the hemicarcerand, which allow for surface-matching with the cytochrome.


Subject(s)
Cytochromes c/chemistry , Light , Crystallography, X-Ray , Electron Transport , Hydrophobic and Hydrophilic Interactions , Models, Molecular , Solubility , Thermodynamics , Water/chemistry , Zinc/chemistry
20.
J Biomed Opt ; 15(4): 047006, 2010.
Article in English | MEDLINE | ID: mdl-20799837

ABSTRACT

In this study, we exploit the sensitivity offered by surface-enhanced Raman scattering (SERS) for the direct detection of thrombin using the thrombin-binding aptamer (TBA) as molecular receptor. The technique utilizes immobilized silver nanoparticles that are functionalized with thiolated thrombin-specific binding aptamer, a 15-mer (5'-GGTTGGTGTGGTTGG-3') quadruplex forming oligonucleotide. In addition to the Raman vibrational bands corresponding to the aptamer and blocking agent, new peaks (mainly at 1140, 1540, and 1635 cm(-1)) that are characteristic of the protein are observed upon binding of thrombin. These spectral changes are not observed when the aptamer-nanoparticle assembly is exposed to a nonbinding protein such as bovine serum albumin (BSA). This methodology could be further used for the development of label-free biosensors for direct detection of proteins and other molecules of interest for which aptamers are available.


Subject(s)
Aptamers, Nucleotide/blood , Aptamers, Nucleotide/chemistry , Biosensing Techniques/methods , Protein Interaction Mapping/methods , Spectrum Analysis, Raman/methods , Thrombin/analysis , Thrombin/chemistry , Aptamers, Nucleotide/immunology , Immunoassay/methods , Molecular Probe Techniques , Protein Binding , Thrombin/immunology
SELECTION OF CITATIONS
SEARCH DETAIL
...