Next Article in Journal
2-[N-(O-ethylaminophenyl) phenyl imidoyl methylylidene]-3,5-diphenyl-1,3,4-thiadiazole
Previous Article in Journal
[1-(4-Chlorophenyl)cyclopropyl](piperazin-1-yl)methanone
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Short Note

1,9-Bis(2-pyridyl)-1,2,8,9-tetrathia-5-oxanonane

by
Jeet Kalia
1 and
Ronald T. Raines
1,2,*
1
Department of Biochemistry, University of Wisconsin–Madison, 433 Babcock Drive, Madison, WI 53706-1544, USA
2
Department of Chemistry, University of Wisconsin–Madison, 1101 University Avenue, Madison, WI 53706-1322, USA
*
Author to whom correspondence should be addressed.
Molbank 2009, 2009(4), M642; https://doi.org/10.3390/M642
Submission received: 21 October 2009 / Accepted: 16 November 2009 / Published: 17 November 2009

Abstract

:
Disulfide crosslinking of proteins is typically performed by treating proteins bearing cysteine residues with small-molecule disulfide reagents. The process results in the formation of a mixed disulfide intermediate, which then reacts with the cysteine residue of another protein molecule to form the crosslinked product. This second step requires the intimate association of two large reactants. The ensuing steric hindrance can result in poor crosslinking yields. Here, we introduce a bis(disulfide) reagent in which activated disulfides are separated by linkers that can alleviate steric hindrance and thereby potentially increase the efficiency of crosslinking.

Graphical Abstract

Disulfide bonds are the most prevalent natural covalent crosslinks in proteins. Formed between cysteine side chains, these linkages are crucial for the structural and functional integrity of proteins. For example, the four disulfides of ribonuclease A remain intact at extremely low pH and high temperatures, thereby contributing enormously to the legendary stability of this protein [1,2]. Disulfides also play a major role in the remarkable stability of thermophilic proteins [3]. In addition to their role as structural scaffolds, disulfides often serve as linkages between protein monomers that exist physiologically as oligomers [4].
Following Nature’s lead, biochemists and chemical biologists have inserted disulfide bonds in proteins and peptides to increase their stability, and to link them together to form multimers [5,6]. Disulfide crosslinking has also been used as a tool to elucidate the structure of proteins [7], define the conformational changes that underlie protein function [8], and modulate enzymatic activity [9,10].
Most disulfide crosslinking experiments are performed by a two-step procedure. In the first step, a cysteine-bearing protein is treated with a labile disulfide reagent, such as 2,2′-dithiopyridine disulfide [11] or 5,5′-dithiobis(2-nitrobenzoic acid) [12]. This reaction results in the formation of a labile mixed disulfide intermediate. In the second step, this intermediate reacts with another protein molecule bearing a reactive cysteine to form the desired crosslinked product. This second step can be problematic because it entails a reaction between two large reactants that can be impeded by steric hindrance, compromising crosslinking yields.
To mitigate steric hindrance, we designed 1,9-bis(2-pyridyl)-1,2,8,9-tetrathia-5-oxanonane (1), a reagent that contains two labile disulfides separated by a linker. Upon treatment with a protein bearing a reactive cysteine, one disulfide of the reagent reacts with the protein to form a stable mixed disulfide. In a second step, the remaining labile disulfide reacts with a cysteine residue of another protein to form the desired crosslinked product. This method is likely to increase crosslinking efficiency compared to the conventional method.
The synthesis of bis(disulfide) 1 was accomplished in a single step with commercial starting materials (Scheme 1). We used the same route to avail a longer congener, 1,12-bis(2-pyridyl)-1,2,11,12-tetrathia-5,8-dioxadodecane (2). We chose oligoethylene oxides to serve as the linker within both reagents because oligo- and polyethylene oxides have favorable properties such as high water solubility, chemical inertness, and low toxicity [13]. Molecules containing polyethylene oxide backbones with various chain lengths are available. Hence, this strategy can be employed to generate reagents possessing a wide range of linker lengths, enabling utility in a variety of applications.

Synthesis

Materials. 2,2′-Dithiopyridine disulfide, 3-oxa-1,5-disulfanylpentane, and 3,6-dioxa-1,8-disulfanyloctane were from Sigma–Aldrich (St. Louis, MO). Synthetic reactions were monitored by thin-layer chromatography with visualization by UV-light or staining with phosphomolybdic acid. Flash chromatography was performed with columns of silica gel 60, 230–400 mesh (Silicycle, Québec City, Québec, Canada).
Instrumentation. NMR spectra were acquired with a Bruker AC+ 300 spectrometer (1H NMR: 300 MHz, 13C NMR: 75 MHz) at the Magnetic Resonance Facility in the Department of Chemistry at Madison. 13C spectra were proton-decoupled. Mass spectra were obtained with electrospray ionization (ESI) techniques.
Synthesis of 1,9-bis(2-pyridyl)-1,2,8,9-tetrathia-5-oxanonane (1). 2,2′-Dithiopyridine disulfide (1.00 g, 4.44 mmol) was dissolved in 20 mL of dichloromethane, and to this solution was added 3-oxa-1,5-disulfanylpentane (0.32 g, 2.22 mmol). The color of the reaction mixture changed to yellow within 2 min after addition of the dithiol. The reaction mixture was stirred at room temperature for 6 h. Solvent was then removed by using a rotary evaporator at water aspirator pressure (<20 torr), and the residue was purified by flash chromatography (80% v/v hexane in ethyl acetate). Bis(disulfide) 1 was obtained as a translucent, yellow liquid in an unoptimized yield of 53%. 1H NMR (300 MHz, CDCl3) δ 8.48–8.44 (m, 2H), 7.78–7.73 (m, 2H), 7.68–7.61 (m, 2H), 7.11–7.05 (m, 2H), 3.67 (t, J = 6.3 Hz, 4H), 2.97 (t, J = 6.3 Hz, 4H) ppm; 13C NMR (75 MHz, CDCl3) δ 160.5, 149.5, 137.2, 120.8, 119.9, 68.9, 38.7 ppm. MS (ESI) m/z 357.2 (MH+ [C14H17N2OS4] = 357.55).
Synthesis of 1,12-bis(2-pyridyl)-1,2,11,12-tetrathia-5,8-dioxadodecane (2). 2,2′-Dithiopyridine disulfide (0.94 g, 4.18 mmol) was dissolved in 20 mL of dichloromethane, and to this solution was added 3,6-dioxa-1,8-disulfanyloctane (0.40 g, 2.09 mmol). The color of the reaction mixture changed to yellow instantaneously upon addition of the dithiol. The reaction mixture was stirred at room temperature for 6 h. Solvent was then removed by using a rotary evaporator at water aspirator pressure (<20 torr), and the residue was purified by flash chromatography (70% v/v hexane in ethyl acetate). Bis(disulfide) 2 was obtained as a translucent, yellow liquid in an unoptimized yield of 33%. 1H NMR (300 MHz, CDCl3) δ 8.47–8.43 (m, 2H), 7.79–7.75 (m, 2H), 7.68–7.62 (m, 2H), 7.11–7.05 (m, 2H), 3.73 (t, J = 6.5 Hz, 4H), 3.56 (m, 4 H), 2.99 (t, J = 6.4 Hz, 4H) ppm; 13C NMR (75 MHz, CDCl3) δ 160.6, 149.7, 137.2, 120.8, 119.8, 70.5, 69.3, 38.6 ppm; MS (ESI) m/z 401.2 (MH+ [C16H21N2O2S4] = 401.6).

Supplementary Materials

Supplementary File 1Supplementary File 2Supplementary File 3

Acknowledgment

This work was supported by grant R01 GM044783 (NIH).

References and Notes

  1. Klink, T.A.; Woycechowsky, K.J.; Taylor, K.M.; Raines, R.T. Contribution of disulfide bonds to the conformational stability and catalytic activity of ribonuclease A. Eur. J. Biochem. 2000, 267, 566–572. [Google Scholar] [CrossRef] [PubMed]
  2. Pecher, P.; Arnold, U. The effect of additional disulfide bonds on the stability and folding of ribonuclease A. Biophys. Chem. 2009, 141, 21–28. [Google Scholar] [CrossRef] [PubMed]
  3. Beeby, M.; O’Connor, B.D.; Ryttersgaard, C.; Boutz, D.R.; Perry, L.J.; Yeates, T.O. The genomics of disulfide bonding and protein stabilization in thermophiles. PLoS Biol. 2005, 3, 1549–1558. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  4. Kim, J.-S.; Raines, R.T. Dibromobimane as a fluorescent crosslinking reagent. Anal. Biochem. 1995, 225, 174–176. [Google Scholar] [CrossRef] [PubMed]
  5. Kotch, F.W.; Raines, R.T. Self-assembly of synthetic collagen triple helices. Proc. Natl. Acad. Sci. USA 2006, 103, 3028–3033. [Google Scholar] [CrossRef] [PubMed]
  6. Ottl, J.; Moroder, L. Disulfide-bridged heterotrimeric collagen peptides containing the collagenase cleavage site of collagen type I. Synthesis and conformational properties. J. Am. Chem. Soc. 1999, 121, 653–661. [Google Scholar] [CrossRef]
  7. Pakula, A.A.; Simon, M.I. Determination of transmembrane protein structure by disulfide cross-linking: The Escherichia coli Tar receptor. Proc. Natl. Acad. Sci. USA 1992, 89, 4144–4148. [Google Scholar] [CrossRef] [PubMed]
  8. Armstrong, N.; Jasti, J.; Beich-Frandsen, M.; Gouaux, E. Measurement of conformational changes accompanying desensitization in an ionotropic glutamate receptor. Cell 2006, 127, 85–97. [Google Scholar] [CrossRef] [PubMed]
  9. Messmore, J.M.; Holmgren, S.K.; Grilley, J.E.; Raines, R.T. Sulfur shuffle: Modulating enzymatic activity by thiol–disulfide interchange. Bioconjugate Chem. 2000, 11, 408–413. [Google Scholar] [CrossRef]
  10. Park, C.; Raines, R.T. Adjacent cysteine residues as a redox switch. Protein Eng. 2001, 14, 939–942. [Google Scholar] [CrossRef] [PubMed]
  11. Grassetti, D.R.; Murray, J.F., Jr. Determination of sulfhydryl groups with 2,2′- or 4,4′-dithiopyridine. Arch. Biochem. Biophys. 1967, 119, 41–49. [Google Scholar] [CrossRef]
  12. Riddles, P.W.; Blakeley, R.L.; Zerner, B. Ellman’s reagent: 5,5′-Dithiobis(2-nitrobenzoic acid)—a reexamination. Anal. Biochem. 1979, 94, 75–81. [Google Scholar] [CrossRef]
  13. Zalipsky, S. Functionalized poly(ethylene glycol) for preparation of biologically relevant conjugates. Bioconjugate Chem. 1995, 6, 150–165. [Google Scholar] [CrossRef]
Scheme 1. Route for the syntheses of bis(disulfide) 1 and 2.
Scheme 1. Route for the syntheses of bis(disulfide) 1 and 2.
Molbank 2009 m642 sch001

Share and Cite

MDPI and ACS Style

Kalia, J.; Raines, R.T. 1,9-Bis(2-pyridyl)-1,2,8,9-tetrathia-5-oxanonane. Molbank 2009, 2009, M642. https://doi.org/10.3390/M642

AMA Style

Kalia J, Raines RT. 1,9-Bis(2-pyridyl)-1,2,8,9-tetrathia-5-oxanonane. Molbank. 2009; 2009(4):M642. https://doi.org/10.3390/M642

Chicago/Turabian Style

Kalia, Jeet, and Ronald T. Raines. 2009. "1,9-Bis(2-pyridyl)-1,2,8,9-tetrathia-5-oxanonane" Molbank 2009, no. 4: M642. https://doi.org/10.3390/M642

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop