Dataset Viewer (First 5GB)
Auto-converted to Parquet Duplicate
protein_name
stringlengths
4
4
structure_title
stringlengths
3
322
main_text
stringlengths
0
302k
3M2H
Crystallographic and Single Crystal Spectral Analysis of the Peroxidase Ferryl Intermediate
Crystallographic and Single Crystal Spectral Analysis of the Peroxidase Ferryl Intermediate Yergalem T. Meharenna, Tzanko Doukovb, Huiying Lia, S. Michael Soltisb,*, and Thomas L. Poulosa,* aDepartments of Molecular Biology and Biochemistry, Pharmaceutical Sciences, and Chemistry, University of California, Irvine, California 92697-3900 bMacromolecular Crystallographic Group, The Stanford Synchrotron Radiation Lightsource, SLAC, Stanford University, Stanford, California 94025 Abstract The ferryl (Fe(IV)O) intermediate is important in many heme enzymes and thus the precise nature of the Fe(IV)-O bond is critical in understanding enzymatic mechanisms. The 1.40 Å crystal structure of cytochrome c peroxidase Compound I has been solved as a function of x-ray dose while monitoring the visible spectrum. The Fe-O bond increases linearly from 1.73 Å in the low x- ray dose structure to 1.90 Å in the high dose structure. The low dose structure correlates well with a Fe(IV)=O bond while we postulate that the high dose structure is the cryo-trapped Fe(III)-OH species previously thought to be Fe(IV)-OH. The ferryl, Fe(IV)O, species is a critically important intermediate in a number of metalloproteins and especially heme enzymes. The high redox potential enables Fe(IV)O to serve as a potent oxidant utilized by several heme enzymes including cytochromes P450, nitric oxide synthase (NOS), cytochrome oxidase, and peroxidases. Since the ferryl intermediate is quite stable in peroxidases, most of what we know about Fe(IV)O in heme enzymes derives from studies with peroxidases. In most heme peroxidases one H2O2 oxidizing equivalent is used to oxidize Fe(III) to Fe(IV)O and the second is used to oxidize an organic group to give Fe(IV)R.+ (1) and this activated intermediate is called Compound I. In most heme peroxidases such as horse radish peroxidase (HRP) R is the porphyrin (2) although in yeast cytochome c peroxidase (CCP) R is the active site Trp191 (3). A majority of studies find that the Fe(IV)-O bond is short, somewhat less than 1.7 Å, thus indicating a Fe(IV)=O bond as opposed to a Fe(IV)-OH bond (4). An empirical formula called Badger’s rule relates the calculated Fe-O bond with the calculated vibrational frequency (5) and the experimental frequencies and EXAFS bond distances fit very well to these plots (5) further supporting a Fe(IV)=O double bond. However, a majority of x-ray crystal structures are distinct outliers giving distances closer to 1.8-1.9 Å (4, 6) with one exception being the HRP Compound I structure (7). These differences are not trivial since the longer bond predicts that the ferryl species should be protonated to give Fe(IV)-OH, while the shorter bond gives Fe(IV)=O. The chemistry of each of these species is quite different (8) and knowing the correct structure is essential if we are to understand details of heme enzyme mechanisms. *To whom correspondences should be addressed. T.L.P.: [email protected]; phone (494) 824-7020; FAX, (949) 824-3280. SUPPORTING INFORMATION AVAILABLE Experimental details and Tables 1S and 2S . This material is available free of charge at http://pubs.acs.org. NIH Public Access Author Manuscript Biochemistry. Author manuscript; available in PMC 2011 October 27. Published in final edited form as: Biochemistry. 2010 April 13; 49(14): 2984–2986. doi:10.1021/bi100238r. NIH-PA Author Manuscript NIH-PA Author Manuscript NIH-PA Author Manuscript A serious problem encountered at high intensity synchrotron x-ray sources is rapid reduction of metal centers, particularly high potential metal centers such as Fe(IV). As a result great care must be taken to minimize reduction and the redox state should be verified during data collection (for example with UV/VIS spectroscopy). We recently found that crystals of the CCP N184R mutant diffract unusually well (9) and thus might provide an opportunity to obtain a low x-ray dose Compound I structure but at sufficiently high resolution to resolve the discrepancies between crystal structures and solution studies. Here we present single crystal spectroscopy together with a composite data collection strategy that has allowed the Fe-O bond distance to be measured as a function of x-ray dose. Fig. 1A shows the single crystal spectrum of CCP Compound I as a function of x-ray dose. Before data collection the spectrum in the 500-700 nm region is identical to the solution spectrum of Compound I. After extensive x-ray exposure (inset to Fig. 1A) the spectrum clearly is no longer that of Compound I nor is this similar to the Fe(III) high spin solution spectrum of CCP. The nature of this species will be discussed further on. Fig. 1B shows the estimated percentage of Compound I remaining in the crystal as a function of x-ray exposure as monitored by changes in the visible spectrum. Based on this plot ~90% of Compound I remains after receiving an estimated x-ray dose of 0.035 MGy (calculations were performed using RADDOSE (10)) or just ~0.1% of the theoretical radiation damage limit for protein crystals, ≈30 MGy (11). Therefore, a data collection strategy for obtaining predominantly Compound I was employed using multiple crystals, none of which received more than 0.035 MGy. With this maximum dose, we estimate that the resulting “integrated” structure is comprised of ~90% Compound I. Crystallographic data collection was carried out at 65 K on SSRL BL9-2 (~4×1011 photons/s at 13.0 KeV). Nearly 100 crystals were mounted and indexed in an automated fashion. Exposures used for indexing were attenuated by 99% and did not significantly contribute to reduction of Compound I. For each crystal, data collections were carried out in 15 separate runs. Run 1 consisted of 5° of data, representing the first 0.035 MGy of x-ray exposure. Then the same 5° of scanning angle were recollected 12 more times giving runs 2 through 13 with increased x-ray dose. In run 14 a full 120° of data were collected in order to fully reduce the crystal followed by run 15 which again repeated the same 5° representing the highest x-ray dose. The same 15-run data collection protocol was adopted for similarly sized crystals and the scanning angles were chosen to optimize the completeness of the data. Each composite data set was assembled by merging 5° of data with identical run numbers from 19 crystals. A total of 15 structures at 1.40 Å resolution were refined providing a picture of the structural changes associated with increasing x-ray dose (Table S1). In Fig. 2A we compare the structures of the low dose (set 1) and the ferric resting state 1.06 Å structure of the N184R mutant (3E2O) (9). In the ferric resting state a water molecule is positioned ≈ 2.0 Å from the heme iron while in the low dose data set the Fe-O oxygen distance is 1.73 Å. In both structures a water molecule is within H-bonding distance of the Fe-linked oxygen. In the ferric state the heme iron is displaced from the porphyrin plane by 0.18 Å toward the proximal His ligand while in Compound I the iron is displaced by 0.07 Å in the opposite direction toward the distal pocket. Thus the net movement of the iron is ≈ 0.25 Å relative to the porphyrin plane owing to the oxidation of the iron from Fe(III) to Fe(IV). Note that the water molecules in the distal pocket, including the one closest to the iron, are located in nearly the same position relative to the heme while the His-Fe bond increases from 2.07 Å to 2.12 Å upon oxidation to Fe(IV). Thus, the short Fe-O bond in the Compound I structure is due in large part to motion of the iron. As in our previous work on peroxide treated CCP (12) Arg48 in the distal pocket forms a 2.78 Å H-bond with the iron linked O atom. Meharenn et al. Page 2 Biochemistry. Author manuscript; available in PMC 2011 October 27. NIH-PA Author Manuscript NIH-PA Author Manuscript NIH-PA Author Manuscript We next compare the set 1 (low dose, Fig. 2C) and set 15 (high dose, Fig. 2D) structures. At the 4.0 σ contour level the electron density between the Fe and O atoms is not continuous in set 15 and the Fe-O bond length has increased from 1.73 Å to 1.90 Å. The local water structure remains largely unchanged. The changes owing to x-ray induced reduction are highlighted by examining a Fo(low dose)-Fo(high dose) electron density difference map contoured at ±5σ (Fig. 2B). This map clearly shows that the iron is positioned quite differently in each structure and is closer toward the distal pocket in the low dose structure. In addition the His-Fe bond decreases from 2.12 Å to 2.07 Å upon photo reduction again due to motion of the iron back into the porphyrin plane. The only other notable feature in the Fo(low dose)-Fo(high dose) difference map is around the carbonyl O atom of the heme ligand, His175. This group is slightly less than 0.1 Å closer to Trp191 in the low dose structure and may reflect a local tightening of the structure around the Trp191 cation radical that provides additional electrostatic stability. The various heme parameter distances are provided in Table S2. The structures of set 1 through set 13 next were used to assess how the Fe-O bond changes as a function of x-ray dose and the results are shown in Fig. 3. The fit to a simple straight line equation is remarkably good and extrapolates to zero dose at a Fe-O bond distance of 1.72 Å. Raman data (13) coupled with Badger’s rule (4) gives a Fe-O bond of 1.68 Å. Therefore, the low dose Compound I crystal structure agrees within 0.04 Å with the Raman data and the ferryl center in CCP Compound I can best be described as Fe(IV)=O and not Fe(IV)-OH. The nature of the ferryl center after extensive x-ray exposure is intriguing: the short Fe-O bond (1.90 Å) compared to the ≈ 2.0 - 2.3 Å observed in Fe(III) high spin peroxidase structures and the total lack of similarity between the high dose spectrum (Fig. 1) and the solution spectrum of Fe(III) CCP shows that the high dose structure is not that of Fe(III) high spin CCP. The spectrum is similar to that of HRP Fe(II) in both the crystal and solution except in HRP there is no ligand coordinated to the iron (7). Since we clearly see a ligand coordinated to the iron in the high dose structure we very likely have trapped either Fe(II)- OH or Fe(III)-OH. Unfortunately we cannot compare single crystal and solution spectra since formation of Fe(III)-OH, and presumably Fe(II)-OH, requires an increase in pH and CCP is not stable above pH 8.0. Our first goal in this study was to further develop the necessary methods and protocols required to obtain x-ray structures of high potential intermediates in metalloproteins. This requires isomorphous crystals that diffract well in order to have sufficient resolution to obtain the level of accuracy required for estimating subtle bond parameter differences (7). Coupling data collection with on-line single crystal spectroscopy to monitor the redox state is also essential. Our second goal was to obtain a very low dose x-ray structure of CCP Compound I at high resolution in order to reconcile the long standing differences observed in the Fe(IV)-O bond distance between most available x-ray structures and other biophysical techniques. The low dose CCP Compound I structure agrees within 0.04 Å of previous experimental estimates indicating that the ferryl species in Compound I is Fe(IV)=O and not Fe(IV)-OH. It should be noted that from the perspective of the heme, CCP Compound I is equivalent to HRP Compound II since both contain Fe(IV) with no porphyrin radical. Thus it is likely that other crystal structures where the Fe(IV)-O bond in Compound II was estimated to be 1.8 Å (7, 14) or longer may also have a significant amount of a reduced iron species. Supplementary Material Refer to Web version on PubMed Central for supplementary material. Meharenn et al. Page 3 Biochemistry. Author manuscript; available in PMC 2011 October 27. NIH-PA Author Manuscript NIH-PA Author Manuscript NIH-PA Author Manuscript Acknowledgments We thank Aina Cohen, John Kovarick and Michael Hollenbeck for their contribution to the design and implementation of the single-crystal microspectrophotometer. Portions of this research were carried out at the Stanford Synchrotron Radiation Lightsource, a national user facility operated by Stanford University on behalf of the U.S. Department of Energy, Office of Basic Energy Sciences. The SSRL Structural Molecular Biology Program is supported by the Department of Energy, Office of Biological and Environmental Research, and by the National Institutes of Health, National Center for Research Resources, Biomedical Technology Program, and the National Institute of General Medical Sciences. Work at UCI was supported by NIH grant GM42614 (TLP). References 1. Poulos TL, Kraut J. J Biol Chem. 1980; 255:8199–8205. [PubMed: 6251047] 2. Dolphin D, Forman A, Borg DC, Fajer J, Felton RH. Proc Natl Acad Sci USA. 1971; 68:614–618. [PubMed: 5276770] 3. Sivaraja M, Goodin DB, Smith M, Hoffman BM. Science. 1989; 245:738–740. [PubMed: 2549632] 4. Behan RK, Green MT. J Inorg Biochem. 2006; 100:448–459. [PubMed: 16500711] 5. Green MT. J Am Chem Soc. 2006; 128:1902–1906. [PubMed: 16464091] 6. Hersleth HP, Hsiao YW, Ryde U, Gorbitz CH, Andersson KK. Chem Biodivers. 2008; 5:2067– 2089. [PubMed: 18972498] 7. Berglund GI, Carlsson GH, Smith AT, Szoke H, Henriksen A, Hajdu J. Nature. 2002; 417:463–468. [PubMed: 12024218] 8. Green MT, Dawson JH, Gray HB. Science. 2004; 304:1653–1656. [PubMed: 15192224] 9. Meharenna YT, Oertel P, Bhaskar B, Poulos TL. Biochemistry. 2008; 47:10324–10332. [PubMed: 18771292] 10. Paithankar KS, Owen RL, Garman EF. J Synchr Radiat. 2009; 16:152–162. 11. Owen RL, Rudino-Pinera E, Garman EF. Proc Natl Acad Sci U S A. 2006; 103:4912–4917. [PubMed: 16549763] 12. Bonagura CA, Bhaskar B, Shimizu H, Li H, Sundaramoorthy M, McRee D, Goodin DB, Poulos TL. Biochemistry. 2003; 42:5600–5608. [PubMed: 12741816] 13. Reczek CM, Sitter AJ, Terner J. J Molec Struc. 1989; 214:27–41. 14. Hersleth HP, Dalhus B, H GC, A KK. J Biol Inorg Chem. 2002; 7:299–304. [PubMed: 11935353] Meharenn et al. Page 4 Biochemistry. Author manuscript; available in PMC 2011 October 27. NIH-PA Author Manuscript NIH-PA Author Manuscript NIH-PA Author Manuscript Figure 1. Single crystal spectra of CCP Compound I as a function of x-ray dose. Prior to x-ray exposure the spectrum is identical to the solution spectrum of Compound I. The estimated percentage of Compound I remaining in the crystal as a function of x-ray dose in panel B was based on the decrease in the absorbance peak at 634 nm. Meharenn et al. Page 5 Biochemistry. Author manuscript; available in PMC 2011 October 27. NIH-PA Author Manuscript NIH-PA Author Manuscript NIH-PA Author Manuscript Figure 2. A) Superposition of the low dose structure (red) on the Fe(III) structure (cyan). Note that the iron is displaced below the plane of the heme in the Fe(III) structure and above the plane of the heme in the low dose structure; B) Fo(low dose)-Fo(high dose) electron density difference map using phases obtained from the low dose structure. The map is contoured at -5.0σ (green) and +5.0σ (blue); C and D) 2Fo-Fc electron density maps contoured at 4.0σ for the dose data set 1 (panel C) and high dose data set 15 (panel D). Oxygen and water molecules are represented by the small spheres. Meharenn et al. Page 6 Biochemistry. Author manuscript; available in PMC 2011 October 27. NIH-PA Author Manuscript NIH-PA Author Manuscript NIH-PA Author Manuscript Figure 3. Plot of the Fe-O distance as a function of x-ray dose. Each of the 13 structures was refined exactly the same way using the same starting structure and two different protocols. In the first the distances between the Fe and N atoms (4 pyrrole and 1 His closed circles) were restrained while in the second protocol no restraints were applied (open circles). At no time were restraints imposed on the Fe-O distance. The estimated error in the Fe-O bond distance is ≈0.017Å (see Supporting Information). Meharenn et al. Page 7 Biochemistry. Author manuscript; available in PMC 2011 October 27. NIH-PA Author Manuscript NIH-PA Author Manuscript NIH-PA Author Manuscript
3M2I
Crystallographic and Single Crystal Spectral Analysis of the Peroxidase Ferryl Intermediate
Crystallographic and Single Crystal Spectral Analysis of the Peroxidase Ferryl Intermediate Yergalem T. Meharenna, Tzanko Doukovb, Huiying Lia, S. Michael Soltisb,*, and Thomas L. Poulosa,* aDepartments of Molecular Biology and Biochemistry, Pharmaceutical Sciences, and Chemistry, University of California, Irvine, California 92697-3900 bMacromolecular Crystallographic Group, The Stanford Synchrotron Radiation Lightsource, SLAC, Stanford University, Stanford, California 94025 Abstract The ferryl (Fe(IV)O) intermediate is important in many heme enzymes and thus the precise nature of the Fe(IV)-O bond is critical in understanding enzymatic mechanisms. The 1.40 Å crystal structure of cytochrome c peroxidase Compound I has been solved as a function of x-ray dose while monitoring the visible spectrum. The Fe-O bond increases linearly from 1.73 Å in the low x- ray dose structure to 1.90 Å in the high dose structure. The low dose structure correlates well with a Fe(IV)=O bond while we postulate that the high dose structure is the cryo-trapped Fe(III)-OH species previously thought to be Fe(IV)-OH. The ferryl, Fe(IV)O, species is a critically important intermediate in a number of metalloproteins and especially heme enzymes. The high redox potential enables Fe(IV)O to serve as a potent oxidant utilized by several heme enzymes including cytochromes P450, nitric oxide synthase (NOS), cytochrome oxidase, and peroxidases. Since the ferryl intermediate is quite stable in peroxidases, most of what we know about Fe(IV)O in heme enzymes derives from studies with peroxidases. In most heme peroxidases one H2O2 oxidizing equivalent is used to oxidize Fe(III) to Fe(IV)O and the second is used to oxidize an organic group to give Fe(IV)R.+ (1) and this activated intermediate is called Compound I. In most heme peroxidases such as horse radish peroxidase (HRP) R is the porphyrin (2) although in yeast cytochome c peroxidase (CCP) R is the active site Trp191 (3). A majority of studies find that the Fe(IV)-O bond is short, somewhat less than 1.7 Å, thus indicating a Fe(IV)=O bond as opposed to a Fe(IV)-OH bond (4). An empirical formula called Badger’s rule relates the calculated Fe-O bond with the calculated vibrational frequency (5) and the experimental frequencies and EXAFS bond distances fit very well to these plots (5) further supporting a Fe(IV)=O double bond. However, a majority of x-ray crystal structures are distinct outliers giving distances closer to 1.8-1.9 Å (4, 6) with one exception being the HRP Compound I structure (7). These differences are not trivial since the longer bond predicts that the ferryl species should be protonated to give Fe(IV)-OH, while the shorter bond gives Fe(IV)=O. The chemistry of each of these species is quite different (8) and knowing the correct structure is essential if we are to understand details of heme enzyme mechanisms. *To whom correspondences should be addressed. T.L.P.: [email protected]; phone (494) 824-7020; FAX, (949) 824-3280. SUPPORTING INFORMATION AVAILABLE Experimental details and Tables 1S and 2S . This material is available free of charge at http://pubs.acs.org. NIH Public Access Author Manuscript Biochemistry. Author manuscript; available in PMC 2011 October 27. Published in final edited form as: Biochemistry. 2010 April 13; 49(14): 2984–2986. doi:10.1021/bi100238r. NIH-PA Author Manuscript NIH-PA Author Manuscript NIH-PA Author Manuscript A serious problem encountered at high intensity synchrotron x-ray sources is rapid reduction of metal centers, particularly high potential metal centers such as Fe(IV). As a result great care must be taken to minimize reduction and the redox state should be verified during data collection (for example with UV/VIS spectroscopy). We recently found that crystals of the CCP N184R mutant diffract unusually well (9) and thus might provide an opportunity to obtain a low x-ray dose Compound I structure but at sufficiently high resolution to resolve the discrepancies between crystal structures and solution studies. Here we present single crystal spectroscopy together with a composite data collection strategy that has allowed the Fe-O bond distance to be measured as a function of x-ray dose. Fig. 1A shows the single crystal spectrum of CCP Compound I as a function of x-ray dose. Before data collection the spectrum in the 500-700 nm region is identical to the solution spectrum of Compound I. After extensive x-ray exposure (inset to Fig. 1A) the spectrum clearly is no longer that of Compound I nor is this similar to the Fe(III) high spin solution spectrum of CCP. The nature of this species will be discussed further on. Fig. 1B shows the estimated percentage of Compound I remaining in the crystal as a function of x-ray exposure as monitored by changes in the visible spectrum. Based on this plot ~90% of Compound I remains after receiving an estimated x-ray dose of 0.035 MGy (calculations were performed using RADDOSE (10)) or just ~0.1% of the theoretical radiation damage limit for protein crystals, ≈30 MGy (11). Therefore, a data collection strategy for obtaining predominantly Compound I was employed using multiple crystals, none of which received more than 0.035 MGy. With this maximum dose, we estimate that the resulting “integrated” structure is comprised of ~90% Compound I. Crystallographic data collection was carried out at 65 K on SSRL BL9-2 (~4×1011 photons/s at 13.0 KeV). Nearly 100 crystals were mounted and indexed in an automated fashion. Exposures used for indexing were attenuated by 99% and did not significantly contribute to reduction of Compound I. For each crystal, data collections were carried out in 15 separate runs. Run 1 consisted of 5° of data, representing the first 0.035 MGy of x-ray exposure. Then the same 5° of scanning angle were recollected 12 more times giving runs 2 through 13 with increased x-ray dose. In run 14 a full 120° of data were collected in order to fully reduce the crystal followed by run 15 which again repeated the same 5° representing the highest x-ray dose. The same 15-run data collection protocol was adopted for similarly sized crystals and the scanning angles were chosen to optimize the completeness of the data. Each composite data set was assembled by merging 5° of data with identical run numbers from 19 crystals. A total of 15 structures at 1.40 Å resolution were refined providing a picture of the structural changes associated with increasing x-ray dose (Table S1). In Fig. 2A we compare the structures of the low dose (set 1) and the ferric resting state 1.06 Å structure of the N184R mutant (3E2O) (9). In the ferric resting state a water molecule is positioned ≈ 2.0 Å from the heme iron while in the low dose data set the Fe-O oxygen distance is 1.73 Å. In both structures a water molecule is within H-bonding distance of the Fe-linked oxygen. In the ferric state the heme iron is displaced from the porphyrin plane by 0.18 Å toward the proximal His ligand while in Compound I the iron is displaced by 0.07 Å in the opposite direction toward the distal pocket. Thus the net movement of the iron is ≈ 0.25 Å relative to the porphyrin plane owing to the oxidation of the iron from Fe(III) to Fe(IV). Note that the water molecules in the distal pocket, including the one closest to the iron, are located in nearly the same position relative to the heme while the His-Fe bond increases from 2.07 Å to 2.12 Å upon oxidation to Fe(IV). Thus, the short Fe-O bond in the Compound I structure is due in large part to motion of the iron. As in our previous work on peroxide treated CCP (12) Arg48 in the distal pocket forms a 2.78 Å H-bond with the iron linked O atom. Meharenn et al. Page 2 Biochemistry. Author manuscript; available in PMC 2011 October 27. NIH-PA Author Manuscript NIH-PA Author Manuscript NIH-PA Author Manuscript We next compare the set 1 (low dose, Fig. 2C) and set 15 (high dose, Fig. 2D) structures. At the 4.0 σ contour level the electron density between the Fe and O atoms is not continuous in set 15 and the Fe-O bond length has increased from 1.73 Å to 1.90 Å. The local water structure remains largely unchanged. The changes owing to x-ray induced reduction are highlighted by examining a Fo(low dose)-Fo(high dose) electron density difference map contoured at ±5σ (Fig. 2B). This map clearly shows that the iron is positioned quite differently in each structure and is closer toward the distal pocket in the low dose structure. In addition the His-Fe bond decreases from 2.12 Å to 2.07 Å upon photo reduction again due to motion of the iron back into the porphyrin plane. The only other notable feature in the Fo(low dose)-Fo(high dose) difference map is around the carbonyl O atom of the heme ligand, His175. This group is slightly less than 0.1 Å closer to Trp191 in the low dose structure and may reflect a local tightening of the structure around the Trp191 cation radical that provides additional electrostatic stability. The various heme parameter distances are provided in Table S2. The structures of set 1 through set 13 next were used to assess how the Fe-O bond changes as a function of x-ray dose and the results are shown in Fig. 3. The fit to a simple straight line equation is remarkably good and extrapolates to zero dose at a Fe-O bond distance of 1.72 Å. Raman data (13) coupled with Badger’s rule (4) gives a Fe-O bond of 1.68 Å. Therefore, the low dose Compound I crystal structure agrees within 0.04 Å with the Raman data and the ferryl center in CCP Compound I can best be described as Fe(IV)=O and not Fe(IV)-OH. The nature of the ferryl center after extensive x-ray exposure is intriguing: the short Fe-O bond (1.90 Å) compared to the ≈ 2.0 - 2.3 Å observed in Fe(III) high spin peroxidase structures and the total lack of similarity between the high dose spectrum (Fig. 1) and the solution spectrum of Fe(III) CCP shows that the high dose structure is not that of Fe(III) high spin CCP. The spectrum is similar to that of HRP Fe(II) in both the crystal and solution except in HRP there is no ligand coordinated to the iron (7). Since we clearly see a ligand coordinated to the iron in the high dose structure we very likely have trapped either Fe(II)- OH or Fe(III)-OH. Unfortunately we cannot compare single crystal and solution spectra since formation of Fe(III)-OH, and presumably Fe(II)-OH, requires an increase in pH and CCP is not stable above pH 8.0. Our first goal in this study was to further develop the necessary methods and protocols required to obtain x-ray structures of high potential intermediates in metalloproteins. This requires isomorphous crystals that diffract well in order to have sufficient resolution to obtain the level of accuracy required for estimating subtle bond parameter differences (7). Coupling data collection with on-line single crystal spectroscopy to monitor the redox state is also essential. Our second goal was to obtain a very low dose x-ray structure of CCP Compound I at high resolution in order to reconcile the long standing differences observed in the Fe(IV)-O bond distance between most available x-ray structures and other biophysical techniques. The low dose CCP Compound I structure agrees within 0.04 Å of previous experimental estimates indicating that the ferryl species in Compound I is Fe(IV)=O and not Fe(IV)-OH. It should be noted that from the perspective of the heme, CCP Compound I is equivalent to HRP Compound II since both contain Fe(IV) with no porphyrin radical. Thus it is likely that other crystal structures where the Fe(IV)-O bond in Compound II was estimated to be 1.8 Å (7, 14) or longer may also have a significant amount of a reduced iron species. Supplementary Material Refer to Web version on PubMed Central for supplementary material. Meharenn et al. Page 3 Biochemistry. Author manuscript; available in PMC 2011 October 27. NIH-PA Author Manuscript NIH-PA Author Manuscript NIH-PA Author Manuscript Acknowledgments We thank Aina Cohen, John Kovarick and Michael Hollenbeck for their contribution to the design and implementation of the single-crystal microspectrophotometer. Portions of this research were carried out at the Stanford Synchrotron Radiation Lightsource, a national user facility operated by Stanford University on behalf of the U.S. Department of Energy, Office of Basic Energy Sciences. The SSRL Structural Molecular Biology Program is supported by the Department of Energy, Office of Biological and Environmental Research, and by the National Institutes of Health, National Center for Research Resources, Biomedical Technology Program, and the National Institute of General Medical Sciences. Work at UCI was supported by NIH grant GM42614 (TLP). References 1. Poulos TL, Kraut J. J Biol Chem. 1980; 255:8199–8205. [PubMed: 6251047] 2. Dolphin D, Forman A, Borg DC, Fajer J, Felton RH. Proc Natl Acad Sci USA. 1971; 68:614–618. [PubMed: 5276770] 3. Sivaraja M, Goodin DB, Smith M, Hoffman BM. Science. 1989; 245:738–740. [PubMed: 2549632] 4. Behan RK, Green MT. J Inorg Biochem. 2006; 100:448–459. [PubMed: 16500711] 5. Green MT. J Am Chem Soc. 2006; 128:1902–1906. [PubMed: 16464091] 6. Hersleth HP, Hsiao YW, Ryde U, Gorbitz CH, Andersson KK. Chem Biodivers. 2008; 5:2067– 2089. [PubMed: 18972498] 7. Berglund GI, Carlsson GH, Smith AT, Szoke H, Henriksen A, Hajdu J. Nature. 2002; 417:463–468. [PubMed: 12024218] 8. Green MT, Dawson JH, Gray HB. Science. 2004; 304:1653–1656. [PubMed: 15192224] 9. Meharenna YT, Oertel P, Bhaskar B, Poulos TL. Biochemistry. 2008; 47:10324–10332. [PubMed: 18771292] 10. Paithankar KS, Owen RL, Garman EF. J Synchr Radiat. 2009; 16:152–162. 11. Owen RL, Rudino-Pinera E, Garman EF. Proc Natl Acad Sci U S A. 2006; 103:4912–4917. [PubMed: 16549763] 12. Bonagura CA, Bhaskar B, Shimizu H, Li H, Sundaramoorthy M, McRee D, Goodin DB, Poulos TL. Biochemistry. 2003; 42:5600–5608. [PubMed: 12741816] 13. Reczek CM, Sitter AJ, Terner J. J Molec Struc. 1989; 214:27–41. 14. Hersleth HP, Dalhus B, H GC, A KK. J Biol Inorg Chem. 2002; 7:299–304. [PubMed: 11935353] Meharenn et al. Page 4 Biochemistry. Author manuscript; available in PMC 2011 October 27. NIH-PA Author Manuscript NIH-PA Author Manuscript NIH-PA Author Manuscript Figure 1. Single crystal spectra of CCP Compound I as a function of x-ray dose. Prior to x-ray exposure the spectrum is identical to the solution spectrum of Compound I. The estimated percentage of Compound I remaining in the crystal as a function of x-ray dose in panel B was based on the decrease in the absorbance peak at 634 nm. Meharenn et al. Page 5 Biochemistry. Author manuscript; available in PMC 2011 October 27. NIH-PA Author Manuscript NIH-PA Author Manuscript NIH-PA Author Manuscript Figure 2. A) Superposition of the low dose structure (red) on the Fe(III) structure (cyan). Note that the iron is displaced below the plane of the heme in the Fe(III) structure and above the plane of the heme in the low dose structure; B) Fo(low dose)-Fo(high dose) electron density difference map using phases obtained from the low dose structure. The map is contoured at -5.0σ (green) and +5.0σ (blue); C and D) 2Fo-Fc electron density maps contoured at 4.0σ for the dose data set 1 (panel C) and high dose data set 15 (panel D). Oxygen and water molecules are represented by the small spheres. Meharenn et al. Page 6 Biochemistry. Author manuscript; available in PMC 2011 October 27. NIH-PA Author Manuscript NIH-PA Author Manuscript NIH-PA Author Manuscript Figure 3. Plot of the Fe-O distance as a function of x-ray dose. Each of the 13 structures was refined exactly the same way using the same starting structure and two different protocols. In the first the distances between the Fe and N atoms (4 pyrrole and 1 His closed circles) were restrained while in the second protocol no restraints were applied (open circles). At no time were restraints imposed on the Fe-O distance. The estimated error in the Fe-O bond distance is ≈0.017Å (see Supporting Information). Meharenn et al. Page 7 Biochemistry. Author manuscript; available in PMC 2011 October 27. NIH-PA Author Manuscript NIH-PA Author Manuscript NIH-PA Author Manuscript
3M2K
Crystal Structure of fluorescein-labeled Class A -beta lactamase PenP in complex with cefotaxime
RESEARCH ARTICLE Open Access Structural studies of the mechanism for biosensing antibiotics in a fluorescein- labeled β-lactamase Wai-Ting Wong, Ho-Wah Au, Hong-Kin Yap, Yun-Chung Leung*, Kwok-Yin Wong* and Yanxiang Zhao* Abstract Background: β-lactamase conjugated with environment-sensitive fluorescein molecule to residue 166 on the Ω-loop near its catalytic site is a highly effective biosensor for β-lactam antibiotics. Yet the molecular mechanism of such fluorescence-based biosensing is not well understood. Results: Here we report the crystal structure of a Class A β-lactamase PenP from Bacillus licheniformis 749/C with fluorescein conjugated at residue 166 after E166C mutation, both in apo form (PenP-E166Cf) and in covalent complex form with cefotaxime (PenP-E166Cf-cefotaxime), to illustrate its biosensing mechanism. In the apo structure the fluorescein molecule partially occupies the antibiotic binding site and is highly dynamic. In the PenP- E166Cf-cefatoxime complex structure the binding and subsequent acylation of cefotaxime to PenP displaces fluorescein from its original location to avoid steric clash. Such displacement causes the well-folded Ω-loop to become fully flexible and the conjugated fluorescein molecule to relocate to a more solvent exposed environment, hence enhancing its fluorescence emission. Furthermore, the fully flexible Ω-loop enables the narrow-spectrum PenP enzyme to bind cefotaxime in a mode that resembles the extended-spectrum β-lactamase. Conclusions: Our structural studies indicate the biosensing mechanism of a fluorescein-labelled β-lactamase. Such findings confirm our previous proposal based on molecular modelling and provide useful information for the rational design of β-lactamase-based biosensor to detect the wide spectrum of β-lactam antibiotics. The observation of increased Ω-loop flexibility upon conjugation of fluorophore may have the potential to serve as a screening tool for novel β-lactamase inhibitors that target the Ω-loop and not the active site. Background β-Lactamase is one of the major mechanisms of antibio- tic resistance in bacteria. Enzymes of this family deacti- vate β-lactam antibiotics by hydrolyzing the conserved β-lactam moiety in the antibiotics and rendering them ineffective to bind to their target proteins, the penicillin- binding proteins (PBPs), which are essential for bacterial cell wall synthesis and survival [1,2]. Detailed mechanis- tic studies of these enzymes over the past decades have revealed a conserved mechanism of β-lactam hydrolysis that consists of two steps, the acylation step in which the β-lactam ring is “opened” and acylated to the side chain hydroxyl group of Ser70 through nucleophilic attack to form the enzyme-substrate acyl adduct ES*; followed by the deacylation step in which the ES* inter- mediate is hydrolyzed and released as E + P facilitated by Glu166 (residue numbering according to the most conserved Class A β-lactamases) [3]. The substrate profile of a β-lactamase in hydrolyzing diverse β-lactam antibiotics is strongly influenced by a structural element termed Ω-loop, a short stretch of residues on the surface of the β-lactamase structure that forms part of the outer part of the antibiotic binding site [4-9]. For narrow-spectrum β-lactamases such as the PenP used in this study and the clinically significant TEM-1 or SHV-1 enzymes, Ω-loop is tightly packed onto the enzyme active site through hydrophobic and electrostatic interactions with residues lining the * Correspondence: [email protected]; [email protected]; [email protected] Department of Applied Biology and Chemical Technology, Central Laboratory of the Institute of Molecular Technology for Drug Discovery and Synthesis, The Hong Kong Polytechnic University, Hung Hom, Hong Hong, China Wong et al. BMC Structural Biology 2011, 11:15 http://www.biomedcentral.com/1472-6807/11/15 © 2011 Wong et al; licensee BioMed Central Ltd. This is an Open Access article distributed under the terms of the Creative Commons Attribution License (http://creativecommons.org/licenses/by/2.0), which permits unrestricted use, distribution, and reproduction in any medium, provided the original work is properly cited. catalytic site, posing as steric hindrance for binding of second- or third-generation antibiotics with bulky side chains attached onto the β-lactam nucleus. Many mutant strains of TEM- and SHV-like β-lactamases overcome this inefficiency and broaden their hydrolytic profile by acquiring mutations in the Ω-loop region to render this region more flexible to accommodate large- sized antibiotics [4-10]. Many extended-spectrum β-lactamases have significantly extended Ω-loop, result- ing in an enlarged active site that readily binds to and hydrolyzes almost all antibiotics [11-13]. Exploiting the proximity of Ω-loop to the antibiotic binding site and its structural flexibility, we have suc- cessfully converted a β-lactamase PenPC from Bacillus cereus 569/H into a biosensor for β-lactam antibiotics by mutating the catalytically critical residue Glu166 on the Ω-loop to cysteine and conjugating an environment- sensitive fluorescein molecule to its reactive side chain thiol group to form PenPC-E166Cf as reported in pre- vious studies [14-16]. Fluorescein is an environment- sensitive fluorophore with suppressed fluorescence in a hydrophobic environment but fluoresces strongly in a polar aqueous environment [17]. The mutation of Glu166 to cysteine severely reduces the efficiency of the deacylation step of β-lactamase catalysis, rendering the enzyme to stall at the acylation step and form a stable ES* acyl adduct that enhances the fluorescence emission of the conjugated fluorescein [16]. We have speculated that the fluorescein molecule is positioned near the cat- alytic site so that the binding and subsequent acylation of β-lactam antibiotics would displace it to a more polar environment, enhancing its fluorescence intensity [16]. Here we report structural studies of fluorescein- conjugated PenP β-lactamase from Bacillus licheniformis 749/C to validate our proposed biosensing mechanism. The structural findings suggest an important role of Ω- loop in the biosensing process, which will help the rational design of improved biosensors for β-lactam detection as well as for novel antibiotics discovery. Results and Discussion The biosensing profile of PenP-E166Cf The biosensing profile of fluorescein conjugated PenP (PenP-E166Cf) for detecting β-lactam antibiotics have never been reported before. In our previous study, a highly similar enzyme, PenPC from Bacillus cereus 569/ H with 58% amino acid sequence identity to PenP, was successfully engineered into a biosensor using the same design scheme (PenPC-E166Cf) [15,16]. We chose to work with PenP in this study for the advantage of its easy propensity for crystallization, which would enable structural studies to understand its biosensing mechan- ism at atomic resolution. PenPC, on the other hand, has poor thermal stability and is difficult to crystallize. Because of the high sequence similarity between these two proteins, as well as the general sequence conserva- tion among all Class A β-lactamase enzymes we expect that PenP can serve as a good model system to under- stand the biosensing mechanism of fluorescein-based biosensing. Indeed the biosensing profile of PenP-E166Cf is highly similar to that of PenPC-E166Cf. The conjugation of fluorescein to the mutated Cys166 residue through thiol linkage is highly efficient for PenP. The ESI-MS profile confirmed that over 90% of PenP was labelled by the fluorophore and converted to PenP-E166Cf, with little unlabelled PenP remaining (Figure 1a). The fluorescence scanning spectrum of PenP-E166Cf shows an increase of ~25% in emitted intensity when the antibiotic cefotaxime is present at 10 μM concentration (Figure 1b). A variety of β-lactam antibiotics, including the first-, second- and third-generation compounds with diverse chemical struc- tures in addition to the conserved β-lactam core, induce significant fluorescence enhancement in PenP-E166Cf at concentration as low as 1 μM (Figure 1c). Lastly the time-dependent spectra of PenP-E166Cf in the presence of cefotaxime at different concentrations ranging from 0.01 μM to 10 μM shows that PenP-E166Cf can detect cefotaxime at concentration as low as 0.01 μM and the fluorescence response is saturated at 1 μM (Figure 1d). The structure of PenP-E166Cf in apo form PenP-E166Cf readily crystallized in the form of clustered needles. These crystals were tinted in bright yellow col- our, indicating the presence of fluorescein (data not shown YZ). To confirm that fluorescein remaining con- jugated to the protein in the crystal form we harvested and thoroughly washed these yellow-coloured crystals and analyzed the dissolved crystals on SDS-PAGE gel under both visible and UV light. A band corresponding to PenP (~30.5 kDa) is clearly visible under both condi- tions, confirming that the crystals are indeed of PenP- E166Cf (data not shown YZ). The structure of PenP-E166Cf was solved by molecu- lar replacement using the known structure of PenP (PDB ID 4BLM) as search model. Two molecules of PenP-E166Cf are found in each asymmetric unit. Struc- ture rebuilding and refinement were done in CCP4 pro- gram [18]. The overall structure of PenP-E166Cf is largely identical to that of the wild-type unlabeled PenP. The RMSD of all 4011 protein atoms between the labeled and wild-type structures is just ~1.5 Å. For main chain atoms, the RMSD is only 0.8 Å. Key residues lin- ing the catalytic site, including Ser70 and mutated Cys166 are virtually identical between the labeled and wild-type structures (Figure 2a). In summary the conju- gation of fluorescein to PenP does not alter its overall structural folding. Wong et al. BMC Structural Biology 2011, 11:15 http://www.biomedcentral.com/1472-6807/11/15 Page 2 of 8 The fluorescein molecule was modeled onto the PenP structure after careful inspection of the fo-fc and 2fo-fc electron density map. These maps are not of high qual- ity at regions around the fluorescein conjugation site, with only pieces of discontinuous density visible at 2.0 s contour level in the fo-fc map (Figure 2a). We tried our best to fit fluorescein into these pieces of electron den- sity, particularly matching the melaimide group to a piece of electron density near the thiol side chain of Cys166, as well as matching the xanthene group at the end of the fluorescein molecule to a large piece of elec- tron density near the catalytic site (Figure 2a). This modeled structure is stable after rounds of structural refinement, showing good electron density for the Ω-loop residues and the fluorescein molecule at 1.0 s contour level in the 2fo-fc map, suggesting that our fitting is reasonable (Figure 2b). However, no electron density was visible for the benzoic group in the 30053 Da (a) (d) PenP(29608Da)+fluorescein(427Da)+water(18Da)=30053Da (c) 0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1 505 525 545 565 585 Wavelength (nm) Relative Fluorescence Intensity E166Cf only 10-5M cefotaxime (b) 10-5 10-6 10-7 5x10-8 10-8 E166Cf only Figure 1 Biosensing of b-lactam antibiotics by fluorescein-labelled PenP. (a) De-convoluted ESI mass spectrum of PenP-E166Cf. The add-up at the bottom confirms the correct mass of the labelled protein. (b) Fluorescence scanning spectra of PenP-E166Cf in the presence of 10-5M cefotaxime in 50 mM phosphate buffer (pH 7.0). (c) Change in fluorescence emission of PenP-E166Cf after incubation with different antibiotics (cefotaxime, ceftriaxone, ceftazidime, cephaloridine, cephalothin, cefoxitin, cefuroxime, penicillin G and ampicillin) at 10-6 M for 100 s. (d) Time- dependent fluorescence spectra in the presence of different concentrations (1 × 10-8 M - 1 × 10-5 M) of cefotaxime monitored at 515 nm. Wong et al. BMC Structural Biology 2011, 11:15 http://www.biomedcentral.com/1472-6807/11/15 Page 3 of 8 mid-region of the fluorophore molecule, indicating that this region is more disordered as compared to other parts of the fluorophore molecule. In our PenP-E166Cf structure the fluorescein molecule partially occupies the outer edge of the antibiotic binding region and is in close contact with several residues at the catalytic site. The maleimide moiety near the thiol link- age site is inserted into the catalytic core, located within 2.5 Å from the side chain of Ser70 on one side and 3.5 Å away from Ω-loop on the other side. The xanthene group near the other end of the fluorescein molecule extends toward the solvent (Figure 2b), loosely packed against β-strand B3 that forms part of the extended substrate binding area involved in coordinating antibiotics as shown in the extended-spectrum class A β-lactamase, Toho-1, in complex with cefotaxime, cephalothin, and benzylpenicillin [19]. No specific interactions were observed between fluorescein and the protein. Total sol- vent accessible area is 188 Å2, 33% of the total surface area, indicating that fluorescein is partially packed against the PenP molecule and not fully solvent exposed. The fluorescein molecule is highly dynamic, as reflected by the poor electron density map as well as high average temperature factor (~72.3). In contrast, the rest of the structure shows excellent electron density and low aver- age temperature factor (~23.5) that is typical of the 2.2 Å data set. The Ω-loop, on which the fluorescein molecule is conjugated, was little affected by the dynamic fluoro- phore and adopts the same conformation as that of the unlabelled PenP (Figure 2b). The structure of PenP-E166Cf in complex with cefotaxime We chose to determine the PenP-E166Cf-cefotaxime structure, using cefotaxime as a representative of the many β-lactam antibiotics because of its positive fluores- cence response induced in PenP-E166Cf as well as its chemical structure that contains functional groups typi- cal of both second- and third-generation antibiotics. Cefotaxime was soaked into the PenP-E166Cf crystals by incubating the crystals in the reservoir solution with 0.01 M cefotaxime added for 20 minutes. The PenP- E166C structure, without the conjugated fluorescein molecule, was used as the starting model for structure determination. After initial rounds of refinement both the fo-fc and 2fo-fc electron density maps were carefully inspected for evidence of cefotaxime and fluorescein, as well as for any structural changes on PenP. The cefotaxime was clearly visible in fo-fc electron density map as covalently bonded through its carbonyl carbon atom C7 to the Og atom of Ser70, which repre- sents the acylated ES* adduct (Figure 3a). But we could not identify any electron density in either fo-fc or 2fo-fc :-loop fluorescein Ser70 Cys166 2.5 Å (b) Ser70 Cys166 :-loop (a) Figure 2 Crystal structure of PenP-E166Cf. (a) The fo-fc omit map of fluorescein-5-maleimide contoured at 2.0 s. (b) The 2fo-fc map of Phe165 to Asn170 and fluorescein-5-maleimide contoured at 1.0 s. Side chains of Phe165 to Asn170 and Ser70 are shown in cpk cylinder model. Fluorescein is shown in green cylinder model. Wong et al. BMC Structural Biology 2011, 11:15 http://www.biomedcentral.com/1472-6807/11/15 Page 4 of 8 map that would be accountable for fluorescein molecule around the location seen in PenP-E166Cf or anywhere nearby. Furthermore, the fo-fc map showed strong nega- tive signal for a large segment of Ω-loop (residues 164 to 174) and the 2fo-fc map showed no electron density for this region at all, indicating this region became highly disordered upon acylation of cefotaxime (Figure 3b). Based on these observations we did not include fluorescein molecule or the disordered region of Ω-loop in our final refined structure of PenP-E166Cf- cefotaxime. The overall structure folding of fluorescein-labeled and cefotaxime-bound PenP is nearly identical to that of the wild-type unlabeled PenP and the fluorescein-labeled PenP-E166Cf. From the calculation result by the CCP4 program, it was found that the B factor of Glu163, (c) :-loop GC1 Toho-1 PenP-E166Cf (c) :-loop GC1 Toho-1 PenP-E166Cf Ser70 cefotaxime Ω-loop Cys166 Ser70 cefotaxime fluorescein :-loop Cys166 Ser70 cefotaxime fluorescein :-loop (a) (b) (c) Figure 3 Crystal structure of PenP-E166Cf-cefotaxime. (a) The fo-fc map of cefotaxime in PenP-E166Cf-cefotaxime complex contoured at 2.0 s. The light blue dash line represents the disordered Arg164 to Pro174 due to the poor electron density. (b) Comparison of PenP-E166Cf- cefotaxime complex with apo PenP-E166Cf structure. The two structures are superimposed by main chain atoms. Key residues including Cys166, Ser70 and cefotaxime are also shown in cpk cylinder model. (c) Comparison of binding mode of cefotaxime in PenP-E166Cf with that of Toho-1 and GC-1. PenP-E166Cf, Toho-1 and GC1 are superimposed by aligning on overall main chain atoms. Cefotaxime is in cylinder model colored in cpk (PenP-E166Cf-cefotaxime), golden (Toho-1) and red (GC1) respectively. Wong et al. BMC Structural Biology 2011, 11:15 http://www.biomedcentral.com/1472-6807/11/15 Page 5 of 8 Gly175, Glu176 on Ω-loop, which are next to the disor- dered region, is significantly higher (~65 Å2) than other parts of the protein (~20 Å2). The refinement statistic for this set of crystal structure has different values from that of the apo PenP-E166Cf structure due to the cefo- taxime and the difference in Ω-loop. To investigate why the binding and acylation of cefotax- ime causes the Ω-loop and the conjugated fluorescein molecule to become highly flexible and structural disor- dered, we superposed the PenP-E166Cf structure onto the PenP-E166Cf-cefotaxime complex structure. Fluorescein is seen as occupying a site that partially overlaps with the acylated cefotaxime; particularly the benzoic group of fluorescein molecule is in direct steric clash with the 7-amino substituent of cefotaxime (Figure 3b). Thus the binding and acylation of cefotaxime to PenP would dis- place fluorescein from its original position to avoid steric clash. It is likely that the Ω-loop, in order to accommodate such displacement, loses its well-folded structure and becomes highly flexible. As a consequence the fluorescein molecule conjugated to the flexible Ω-loop becomes fully exposed to the polar aqueous environment, leading to enhanced fluorescence. Thus our structural findings con- firmed our initial proposal of a biosensing mechanism based on displacement of fluorescein [15,16]. To understand the impact of conjugated fluorescein molecule on the substrate binding kinetics of PenP we compared the PenP-E166Cf-cefotaxime structure to two other β-lactamase structures in complex with cefotaxime, including the narrow-spectrum Toho-1 and the extended-spectrum GC1 [19,20]. In Toho-1 structure the methoxyimino side chain points away from the active site and is solvent-exposed (Figure 3c). Such an orientation packs the methoxyimino side chain tightly against the thiozolyl ring, leading to a distorted configuration of the cephem nucleus that is catalytically incompetent for dea- cylation [19]. In GC1 structure the transition analog of cefotaxime binds to GC1 in a fully extended conforma- tion, with oxyimino group inserted to active site and extended away from the thiozolyl ring (Figure 3c). This conformation is regarded as catalytically competent to facilitate deacylation because the distortion on the cephem nucleus is released [20]. Importantly, the binding mode of cefotaxime in our PenP-E166Cf-cefotaxime structure closely resembles that of GC1 (Figure 3c), sug- gesting that with its Ω-loop fully flexible the naturally narrow-spectrum PenP can accommodate cefotaxime in a manner that resembles the extended-spectrum GC1. Conclusions Our structural studies indicate the molecular mechan- ism how fluorescein-labeled β-lactamase detects β-lac- tam antibiotics. The conjugated fluorescein molecule is located near the catalytic site and partially occupies the antibiotic binding region. The binding and acylation of β-lactam antibiotics such as cefotaxime would expel the fluorescein molecule from its original position and leads to increased flexibility of the Ω-loop, to which the fluor- ophore is linked. As a result, the fluorophore is relo- cated from its original position with partial solvent exposure to become fully solvent exposed, leading to enhanced fluorescence emission. These findings confirm our previous proposal based on structural modeling. Furthermore the Ω-loop demonstrates the propensity of becoming highly flexible and unstructured if its tight packing against the catalytic site is disturbed. Such increased flexibility enables PenP to bind and acylate cefotaxime, a naturally poor substrate, in a manner that resembles the extended-spectrum cefotaxime-resistant β-lactamases. This finding could be valuable in the future design of novel antibiotics that resist the binding or hydrolysis by β-lactamases. Methods Protein expression and purification Two constructs of PenP protein were used for our experiments, the maltose binding protein (MBP)-fusion construct for time-dependent fluorescence measure- ments and the His6-tagged construct for crystallization and structural studies, as well as scanning fluorescence spectra. The MBP fusion has been shown not to inter- fere with fluorescence measurements in our previous studies (data not shown). The MBP-fusion construct was cloned into pMAL-c2X vector (NEB). The His6- tagged PenP enzyme was cloned into a modified pRset- A vector (Invitrogen) with a TEV protease cleavage site upstream of the PenP gene. The E166C mutation was constructed using QuikChange Site-Directed Mutagen- esis Kits (Strategene). The MBP-fusion construct was expressed in E. coli strain BL21 (DE3) at 37°C for overnight after induction by 300 μM IPTG when A600 reached 0.5-0.7. The har- vested cells were centrifuged and lysed by sonication. The supernatant after sonication was passed through amylose affinity chromatography. The eluted fractions were pooled and buffer exchanged to 20 mM ammo- nium bicarbonate. The protein was freeze-dried for sto- rage afterwards. The His6-tagged PenP protein was expressed in E. coli strain BL21 (DE3) at 37°C for overnight after induction by 200 μM IPTG when A600 reached 0.8-1.2. The har- vested cells were centrifuged and the supernatant was passed through Nickel affinity chromatography, followed by DEAE anion exchange chromatography. The frac- tions containing the target protein were pooled and con- centrated by Amicon® Ultra-15 Centrifugal Filter Devices (Millipore NMWL = 10,000). The His6-tag was cleaved by adding the TEV protease in 1:20 molar ratio Wong et al. BMC Structural Biology 2011, 11:15 http://www.biomedcentral.com/1472-6807/11/15 Page 6 of 8 to the concentrated PenP-E166C protein (2 mg/ml). The mixture was incubated at 30°C for 6 hours and was further purified by Nickel affinity chromatography to remove uncleaved protein. Fluorescein labeling of PenP-E166C to form PenP-E166Cf A ten-fold molar excess of fluorescein, with concentra- tion of 20 mM, was dissolved in DMF (Dimethyl forma- mide) and added to the concentrated PenP-E166C protein solution drop by drop. The labelling reaction was allowed to proceed in darkness with stirring for 1 hour, and then dialysed against 50 mM potassium phos- phate buffer (pH 7.0) at 4°C for several times in order to remove excess fluorescein. The labelled PenP-E166Cf pro- tein was concentrated to less than 1 ml and further puri- fied by Superdex™75 gel filtration column (GE Healthcare). The running buffer contains 20 mM Tris- HCl, 50 mM NaCl, pH 7.5. The target fractions were pooled and concentrated by Amicon Ultra to 25 mg/ml. The labelling efficiency was confirmed by ESI-MS. Fluorescence spectra of PenP-E166Cf for antibiotic detection Fluorescence profile of PenP-E166Cf alone, as well as in presence of various β-lactams were measured using Per- kin-Elmer LS50B spectrofluorimeter. Both scanning spectra and time-dependent spectra were measured. Dif- ferent β-lactam antibiotics, including cefotaxime, cef- triaxone, ceftazidime, cephaloridine, cephalothin, cefoxitin, cefuroxime, penicillin G, and ampicillin, were incubated with PenP-E166Cf for 100 s at 1 μM to allow sufficient acylation of the antibiotic to form ES* adduct. The product after acylation was subjected to fluores- cence measurement as previously described [15]. Crystallization, structure determination and refinement Crystals of PenP-E166Cf were grown by hanging-drop vapour diffusion method after mixing 1 μl of protein and 1 μl of reservoir solution containing 25% (w/v) PEG 4000, 0.1 M Hepes pH 7.2, 0.4 M NH4Acetate and 0.2 M K2HPO4. Small crystals in the form of clustered nee- dles appeared readily. For data collection, single crystals were obtained after separating them from the clustered needles. Crystals were harvested and cryoprotected in its reservoir solution supplemented with 20% ethylene gly- col for one minute prior to flash freeze and data collec- tion on the Rigaku MicroMax™-007HF x-ray machine. For PenP-E166Cf-cefotaxime data set, crystals were soaked in its growth solution added with 0.01 M of cefotaxime for 15 minutes and then mounted to the x-ray machine. Data were integrated and scaled by Crys- talClear™1.3.5 SP2 (Rigaku Inc.). The crystals belong to the monoclinic group P21 with cell parameter: a = 43.43 Å, b = 92.3 Å, c = 66.43 Å and β = 104°. The PenP-E166Cf crystals diffracted to 2.15 Å resolution, while the PenP-E166Cf-cefotaxime crystal diffracted to 2.8 Å. Both structures were determined by molecular replacement using PenP structure as the search model (PDB ID 4BLM) [21]. The program COOT was used for inspection of electron density maps and model building [22]. There are two molecules per asymmetric unit. The fluorescein and cefotaxime mole- cules were built by PRODRG [23] and appended to the PenP structure for refinement. Structure determination and refinement of PenP-E166Cf and PenP-E166Cf- cefotaxime were done using the CCP4 program suite [18]. A summary of the crystallographic data and refine- ment statistics are given in Table 1. The coordinates and structure factors from this study have been Table 1 X-ray data-collection and structure refinement statistics. E166Cf E166Cf+cefotaxime PDB code 3M2J 3M2K Data collection Space group P21 P21 Unit cell parameters (Å) a 43.3 43.5 b 92.3 91.4 c 66.3 66.1 b 104.82 104.52 Resolution range (Å) 52-2.15 (2.24-2.15) 45-2.80 (2.95-2.80) No. of total reflections 79750 40611 No. of unique reflections 29537 12412 I/s 7.1 (2.7) 6.3 (2.4) Completeness (%) 97.0 (99.5) 99.8 (99.9) Rmerge (%) 9.7 (27.1) 11.8 (32.0) Structure refinement Resolution (Å) 50.0-2.20 45.0-2.80 Rcryst/Rfree (%) 20.0/23.2 21.2/27.7 r.m.s.d. bonds (Å)/angles (°) 0.018/1.784 0.010/1.672 No. of reflections Working set 24217 11749 Test set 1291 647 No. of atoms Protein atoms 4011 3706 Water molecules 254 29 Average B-factor (Å2) Main chain 24.7 16.96 Ligand molecules 48.4 42.46 Water 32.7 10.7 Wong et al. BMC Structural Biology 2011, 11:15 http://www.biomedcentral.com/1472-6807/11/15 Page 7 of 8 deposited into Protein Data Bank (PDB) under accession codes 3M2J (PenP-E166Cf apo structure) and 3M2K (PenP-E166Cf-cefotaxime). Acknowledgements This work was supported by the Research Grants Council (PolyU 5463/05 M, PolyU 5017/06P, PolyU 5641/08 M, and PolyU 5639/09M), the Area of Excellence Fund of the University Grants Committee (AoE/P-10/01) and the Research Committee of the Hong Kong Polytechnic University. We thank Shanghai Synchrotron Radiation Facility (SSRF) for access to beam time. Mr. C.H. Cheng is acknowledged for technical assistance with in-house x-ray crystallography facility. Authors’ contributions WTW performed experiments, analyzed data and drafted manuscript. HWA and HKY assisted in experiments. YXZ, KYW and YCL designed project, analyzed data and drafted manuscript. All authors read and approved the final manuscript. Received: 21 September 2010 Accepted: 28 March 2011 Published: 28 March 2011 References 1. Fisher JF, Mobashery S: Three decades of the class A beta-lactamase acyl- enzyme. Curr Protein Pept Sci 2009, 10(5):401-407. 2. Neu HC: The crisis in antibiotic resistance. Science 1992, 257(5073):1064-1073. 3. Strynadka NC, Adachi H, Jensen SE, Johns K, Sielecki A, Betzel C, Sutoh K, James MN: Molecular structure of the acyl-enzyme intermediate in beta- lactam hydrolysis at 1.7 A resolution. Nature 1992, 359(6397):700-705. 4. Palzkill T, Botstein D: Identification of amino acid substitutions that alter the substrate specificity of TEM-1 beta-lactamase. J Bacteriol 1992, 174(16):5237-5243. 5. Palzkill T, Le QQ, Venkatachalam KV, LaRocco M, Ocera H: Evolution of antibiotic resistance: several different amino acid substitutions in an active site loop alter the substrate profile of beta-lactamase. Mol Microbiol 1994, 12(2):217-229. 6. Petrosino J, Cantu C, Palzkill T: beta-Lactamases: protein evolution in real time. Trends Microbiol 1998, 6(8):323-327. 7. Petrosino JF, Palzkill T: Systematic mutagenesis of the active site omega loop of TEM-1 beta-lactamase. J Bacteriol 1996, 178(7):1821-1828. 8. Trehan I, Beadle BM, Shoichet BK: Inhibition of AmpC beta-lactamase through a destabilizing interaction in the active site. Biochemistry 2001, 40(27):7992-7999. 9. Banerjee S, Pieper U, Kapadia G, Pannell LK, Herzberg O: Role of the omega-loop in the activity, substrate specificity, and structure of class A beta-lactamase. Biochemistry 1998, 37(10):3286-3296. 10. Knox JR: Extended-spectrum and inhibitor-resistant TEM-type beta- lactamases: mutations, specificity, and three-dimensional structure. Antimicrob Agents Chemother 1995, 39(12):2593-2601. 11. Powers RA, Caselli E, Focia PJ, Prati F, Shoichet BK: Structures of ceftazidime and its transition-state analogue in complex with AmpC beta-lactamase: implications for resistance mutations and inhibitor design. Biochemistry 2001, 40(31):9207-9214. 12. Nukaga M, Kumar S, Nukaga K, Pratt RF, Knox JR: Hydrolysis of third- generation cephalosporins by class C beta-lactamases. Structures of a transition state analog of cefotoxamine in wild-type and extended spectrum enzymes. J Biol Chem 2004, 279(10):9344-9352. 13. Crichlow GV, Kuzin AP, Nukaga M, Mayama K, Sawai T, Knox JR: Structure of the extended-spectrum class C beta-lactamase of Enterobacter cloacae GC1, a natural mutant with a tandem tripeptide insertion. Biochemistry 1999, 38(32):10256-10261. 14. Chan PH, Chan KC, Liu HB, Chung WH, Leung YC, Wong KY: Fluorescein- labeled beta-lactamase mutant for high-throughput screening of bacterial beta-lactamases against beta-lactam antibiotics. Anal Chem 2005, 77(16):5268-5276. 15. Chan PH, Liu HB, Chen YW, Chan KC, Tsang CW, Leung YC, Wong KY: Rational design of a novel fluorescent biosensor for beta-lactam antibiotics from a class A beta-lactamase. J Am Chem Soc 2004, 126(13):4074-4075. 16. Chan PH, So PK, Ma DL, Zhao Y, Lai TS, Chung WH, Chan KC, Yiu KF, Chan HW, Siu FM, et al: Fluorophore-labeled beta-lactamase as a biosensor for beta-lactam antibiotics: a study of the biosensing process. J Am Chem Soc 2008, 130(20):6351-6361. 17. Klonis N, Clayton AH, Voss EW Jr, Sawyer WH: Spectral properties of fluorescein in solvent-water mixtures: applications as a probe of hydrogen bonding environments in biological systems. Photochem Photobiol 1998, 67(5):500-510. 18. The CCP4 suite: programs for protein crystallography. Acta Crystallogr D Biol Crystallogr 1994, 50(Pt 5):760-763. 19. Shimamura T, Ibuka A, Fushinobu S, Wakagi T, Ishiguro M, Ishii Y, Matsuzawa H: Acyl-intermediate structures of the extended-spectrum class A beta-lactamase, Toho-1, in complex with cefotaxime, cephalothin, and benzylpenicillin. J Biol Chem 2002, 277(48):46601-46608. 20. Crichlow GV, Nukaga M, Doppalapudi VR, Buynak JD, Knox JR: Inhibition of class C beta-lactamases: structure of a reaction intermediate with a cephem sulfone. Biochemistry 2001, 40(21):6233-6239. 21. Knox JR, Moews PC: Beta-lactamase of Bacillus licheniformis 749/C. Refinement at 2 A resolution and analysis of hydration. J Mol Biol 1991, 220(2):435-455. 22. Emsley P, Cowtan K: Coot: model-building tools for molecular graphics. Acta Crystallogr D Biol Crystallogr 2004, 60(Pt 12 Pt 1):2126-2132. 23. Schuttelkopf AW, van Aalten DM: PRODRG: a tool for high-throughput crystallography of protein-ligand complexes. Acta Crystallogr D Biol Crystallogr 2004, 60(Pt 8):1355-1363. doi:10.1186/1472-6807-11-15 Cite this article as: Wong et al.: Structural studies of the mechanism for biosensing antibiotics in a fluorescein-labeled β-lactamase. BMC Structural Biology 2011 11:15. Submit your next manuscript to BioMed Central and take full advantage of: • Convenient online submission • Thorough peer review • No space constraints or color figure charges • Immediate publication on acceptance • Inclusion in PubMed, CAS, Scopus and Google Scholar • Research which is freely available for redistribution Submit your manuscript at www.biomedcentral.com/submit Wong et al. BMC Structural Biology 2011, 11:15 http://www.biomedcentral.com/1472-6807/11/15 Page 8 of 8
3M2L
Crystal structure of the M113F mutant of alpha-hemolysin
Molecular bases of cyclodextrin adapter interactions with engineered protein nanopores Arijit Banerjeea, Ellina Mikhailovaa, Stephen Cheleya, Li-Qun Gub,2, Michelle Montoyac,3, Yasuo Nagaokaa,4, Eric Gouauxd, and Hagan Bayleya,1 aDepartment of Chemistry, University of Oxford, Oxford, OX1 3TA, United Kingdom; bDepartment of Medical Biochemistry & Genetics, Texas A&M University System Health Science Center, College Station, TX 77843-1114; cDepartment of Biochemistry and Molecular Biophysics and Howard Hughes Medical Institute, Columbia University, New York, NY 10032; and dVollum Institute and Howard Hughes Medical Institute, Oregon Health and Science University, Portland, OR 97239 Edited by Gregory A. Petsko, Brandeis University, Waltham, MA, and approved March 3, 2010 (received for review December 15, 2009) Engineered protein pores have several potential applications in biotechnology: as sensor elements in stochastic detection and ultrarapid DNA sequencing, as nanoreactors to observe single- molecule chemistry, and in the construction of nano- and micro- devices. One important class of pores contains molecular adapters, which provide internal binding sites for small molecules. Mutants of the α-hemolysin (αHL) pore that bind the adapter β-cyclodextrin (βCD) ∼104 times more tightly than the wild type have been ob- tained. We now use single-channel electrical recording, protein en- gineering including unnatural amino acid mutagenesis, and high- resolution x-ray crystallography to provide definitive structural in- formation on these engineered protein nanopores in unparalleled detail. alpha-hemolysin ∣single molecule ∣stochastic sensing ∣structure ∣ unnatural amino acid M any research groups have used protein engineering to obtain enzymes and antibodies with new properties suited for specific tasks (1–6). Fewer groups have taken on the difficult problem of engineering membrane proteins (7). We have engi- neered the α-hemolysin protein pore, mindful of several potential applications in biotechnology, including its ability to act as a de- tector in stochastic sensing (8) and ultrarapid DNA sequencing (9), to serve as a nanoreactor for the observation of single- molecule chemistry (10) and to act as a component for the con- struction of nano- and microdevices (11). An important breakthrough in this area, which enabled the sto- chastic sensing of organic molecules including the detection of DNA bases in the form of nucleoside monophosphates (12, 13), was the discovery of internal molecular adapters, a form of non- covalent protein modification (14). Most useful have been cyclo- dextrin (CD) adapters, which have until now been used in the absence of detailed structural information about how they work. The present paper is a definitive investigation, which provides such information through the application of a wide variety of technical approaches: single-channel electrical recording, protein engineering including unnatural amino acid mutagenesis, and x-ray crystallography. The studies employing mutagenesis show that the striking interactions seen in the crystal structures also occur in individual pores in lipid bilayers. We reveal that the tight-binding αHL mutants (15) M113N7 and M113F7 bind βCD in different orientations within the hep- tameric pore. In the case of M113N7, the top (primary hydroxyls) of the CD ring faces the trans entrance of the pore. In the case of M113F7, the bottom (secondary hydroxyls) of the CD ring faces the trans entrance, while the top of the ring is bonded to the pore through remarkable CH-π interactions. Another tight-binding mutant, M113V7, can bind the CD in both orientations. These results illustrate the exquisite level of engineering that can be achieved with protein nanopores, which is, for example, far be- yond what is possible with solid-state pores. The work also pro- vides information valuable for the design of new binding sites within the lumen of the αHL pore or within other β-barrel pro- teins. Our results will be of interest to others exploring the inter- actions of CDs with the αHL pore (16, 17), including groups involved in computational studies (18, 19). In addition CDs bind to a variety of other pores, including porins (20, 21) and connex- ins (22), and are being tested in vivo as blockers of the anthrax protective antigen pore (23, 24). The CD adapter concept has also been incorporated into other formats, e.g., with glass nano- pores (25), and artificial pores based on CDs have been made by several groups (26–28). Our work is pertinent to these studies. Results Kinetics and Thermodynamics of the Interactions of βCD with αHL Pores Containing Met, Phe and Asn at Position 113. We showed earlier that position 113 in the αHL pore (Fig. 1A) is critical for the bind- ing of βCD (14). Subsequently, residue 113, which is Met in the WT protein, was changed to each of the remaining 19 naturally occurring amino acids by site-directed mutagenesis (15). We found that 11 of these mutants, expressed as homoheptamers, bound βCD with a similar affinity and with similar kinetics to the WT homoheptamer. Two mutants (P, W) bound βCD about 10 times more strongly than the WT homoheptamer, while six of them (V, H, Y, D, N, F) bound with high affinity, i.e., with a Kd value 103 to 104 times lower than the WT. Remarkably, the side chains of the latter six amino acids bear little resemblance to one another, and this issue is addressed in the present paper. We first examined the two amino acids with the most disparate side chains (Fand N) by making heteromeric pores containing WT (Met-113), M113F, and M113N subunits. Three series of heteroheptamers were produced: WT7−nM113Nn, WT7−nM113Fn, and M113F7−nM113Nn. The heteroheptamers were separated by SDS-polyacrylamide gel electrophoresis aided by an oligoaspartate (D8) tail on the first of the two types of sub- unit (Fig. 1B) (29). All 21 combinations of WT, M113F, and M113N subunits formed αHL pores that interacted with βCD as shown by single-channel current recordings, which revealed the extent of block by βCD (Fig. S1), the association and dissociation Author contributions: A.B., S.C., E.G., and H.B. designed research; A.B., E.M., S.C., L.-Q.G., M.M., and Y.N. performed research; A.B., E.G., and H.B. analyzed data; and A.B. and H.B. wrote the paper. The authors declare no conflict of interest. This article is a PNAS Direct Submission. 1To whom correspondence should be addressed. E-mail: [email protected]. 2Present address: Department of Biological Engineering and Dalton Cardiovascular Research Center, University of Missouri, Columbia, MO 65211. 3Present address: Nature Structural & Molecular Biology, 75 Varick Street, 9th Floor, New York NY 10013-1917. 4Present address: Department of Biotechnology, Faculty of Engineering, Kansai University, 3-3-35 Yamate-cho, Suita, Osaka 564-8680, Japan. This article contains supporting information online at www.pnas.org/cgi/content/full/ 0914229107/DCSupplemental. www.pnas.org/cgi/doi/10.1073/pnas.0914229107 PNAS ∣May 4, 2010 ∣vol. 107 ∣no. 18 ∣8165–8170 BIOCHEMISTRY rate constants for βCD (kon and koff), and (from the latter) the equilibrium dissociation constant for βCD (Kd ¼ koff∕kon) (15). The kon values for βCD for the 21 combinations of subunits were all similar at ∼5 × 105 M−1 s−1 (Fig. 1C, Upper). By contrast, the koff values differed widely, ranging from ∼5 × 10−2 s−1 to ∼103 s−1. For WT7−nM113Nn and WT7−nM113Fn, the koff values decreased as M113N or M113F subunits were added. In the case of M113N, there was a steep drop in the value of koff after the fifth subunit had been incorporated. In the case of M113F, the decrease in the value of koff occurred less precipitously as the M113F subunits were added (Fig. 1C, Lower). Intriguingly, with M113F7−nM113Nn, koff first increased as M113N subunits were added to M113F7 until n ¼ 4 (M113F3M113N4) and then de- creased for larger values of n (Fig. 1C, Lower). We recognize that there is more than one permutation of heteromers containing two to five mutant subunits (Fig. 1B), but we have ignored this fact here because no significant differences in the properties of indi- vidual heteromers were observed. For example, 42 recordings were made of WT5M113N2, which has three permutations. Because, kon showed little variation with subunit composition, the variation in Kd was similar to the variation in koff (Fig. 1C). While these studies were in progress, the crystal structures of βCD complexed to M113N7 (Fig. 2B) and M113F7 (Fig. 2C) were solved (Table S1) (30). High-resolution structures could be obtained because the CD and the αHL pore have the same C7 symmetry. In the case of M113N7, βCD is bound with the second- ary hydroxyl face “upward” (Fig. 2B). In each glucose unit of the βCD, the 2-hydroxyl is hydrogen bonded to the side-chain amide of an Asn-113 (the residue introduced by mutagenesis) and the 3-hydroxyl is hydrogen bonded to the ϵ-amino group of Lys-147. In the case of M113F7, two βCDs are bound to the αHL pore (Fig. 2C). It is the top βCD in the structure that concerns us, be- cause it is in contact with the Phe-113 residues introduced by mu- tagenesis. It is immediately apparent that the top βCD in M113F7 is in the opposite orientation to the βCD in M113N7 with each 6-hydroxyl group in a CH-π bonding interaction (31–35) with a Phe-113 side chain. The opposite orientations of the βCDs in M113N7 and M113F7 immediately explain why heteromers formed from similar numbers of M113N and M113F subunits (e.g., M113N4M113F3) bind βCD weakly (see also Discussion). Unnatural Amino Acid Mutagenesis. To further explore the range of noncovalent interactions that are available when βCD binds to the αHL pore, five unnatural amino acids (Fig. 3A and Fig. S2) were incorporated at position 113, by using the in vitro nonsense codon suppression method (36). In particular, we had noted that M113V7 containing the β-branched Val binds βCD tightly (15), and therefore we compared cyclopropylglycine (Cpg) and cyclo- propylalanine (Cpa). We also further examined the means by which M113F7 binds βCD tightly, by comparing the properties of 4-fluorophenylalanine (f1F), pentafluorophenylalanine (f5F), and cyclohexylalanine (Cha) at position 113. The five homomeric pores all produced single-channel cur- rents with unitary conductance values in the range expected for properly assembled heptamers (Fig. S3). All five bound βCD (Fig. 3B, Level 2), either tightly (f1F, Cpg) or weakly (f5F, Cha, Cpa) as described in detail below. During the long βCD binding events, additional current spikes were seen (Fig. 3B). Similar Fig. 1. Binding of βCD by heteromeric αHL pores formed by WT, M113F and M113N subunits. (A) Crystal structure of WT-αHL (61) showing residue 113 (Met, yellow). Left panel, side view and right panel, top view. (B) Separation of 35S-labeled αHL heteroheptamers by SDS-polyacrylamide electrophoresis. The separation of the M113F7-nM113Nn heteromers is shown as detected by autoradiography of a dried gel. The M113F subunits carried a D8 tail. Lane 1, M113N7; lane 2, M113F7−nM113Nn (the heteromers formed from several preparations made with differing ratios of M113F and M113N subunits were mixed to give roughly equal amounts of each subunit combination); lane 3, M113F7. A diagram of the eight different combinations of subunits and their permuta- tions is shown to the right of the autoradiogram. The various permutations are not separated by electrophoresis. (C) Kinetics of the interaction of βCD with single heteromeric αHL pores as determined by bilayer recording. Values of kon were calculated by using kon ¼ 1∕ðτon½βCDÞ, where τon is the mean interevent interval. Values of koff were determined by using koff ¼ 1∕τoff, where τoff is the mean dwell time of βCD in the pore. Values of Kd were calculated by using Kd ¼ koff∕kon. Each point represents the mean  s:d: for three or more determinations. Where they cannot be seen, the s.d. values lie within the symbol. Black squares, WT7−nM113Nn; gray squares, M113F7−nM113Nn; empty squares, M113F7−nWTn. 8166 ∣ www.pnas.org/cgi/doi/10.1073/pnas.0914229107 Banerjee et al. events had been observed previously with certain Met-113 repla- cement mutants and may represent movement of the βCD at its binding site (e.g., rotation about axes perpendicular to the C7 axis) (15). The additional current spikes were more prevalent for M113V7 and M113Cpg7, which may take part in more con- formationally labile interactions with βCD, compared with say M113F7 (Fig. S4). Interactions of βCD with Homoheptamers Bearing Aromatic Residues at Position 113. To further understand the nature of the binding of βCD to aromatic side chains, we examined the kinetics of βCD binding to the homoheptamers containing f1F or f5F at position 113, M113f1F7 and M113f5F7 (Fig. 3C). For both mutants, the value of kon was very similar to that of WT7, but the values of koff and therefore Kd for M113f1F7 differed dramatically from WT7 and were close to the values for the tight-binding mutant M113F7 (Table S2A). By contrast, koff and Kd for M113f5F7 were similar to the values for WT7 (Table S2A). To determine whether M113f1F7 binds βCD in the same orien- tation as M113F7 (Fig. 2C), we made heteromers of the M113f1F subunit with M113N or M113F and examined M113F4M113f1F3 and M113N4M113f1F3. M113F4M113f1F3 binds βCD as strongly as either M113F7 or M113f1F7, but M113N4M113f1F3 binds βCD weakly with a similar affinity to WT7 (Fig. 3D and Table S3). Therefore, it is reasonable to infer that M113F7 and M113f1F7 bind βCD in the same orientation with the 6- hydroxyl groups of the CD in proximity to the aromatic rings on the protein. Cyclohexylalanine (Cha) was used to replace the aromatic side chains with a roughly isosteric hydrophobic group. Again the va- lue of kon for βCD was little changed, but koff for M113Cha7 had an intermediate value of 42  6 s−1. Therefore, M113Cha7 binds βCD more weakly than M113F7 but distinctly more strongly than the WT7 pore (Table S2A and Fig. 3C). Interactions of βCD with Homoheptamers Bearing Hydrophobic Resi- dues at Position 113. M113V7 binds βCD very strongly, and there- fore we compared αHL pores with Cpg or Cpa at position 113. Cpg is roughly isosteric with Val, and like Val has a β-branched side chain. Gratifyingly, M113Cpg7 has a kon value similar to the other αHL pores, and koff and Kd values close to those of M113V7 (Table S2B and Fig. 3E). Cyclopropylalanine (Cpa), with an additional methylene group compared to Cpg, is roughly isosteric with Leu, a weak binder, and M113Cpa7 also binds βCD weakly with kon, koff and Kd values similar to those of WT7 (Table S2B and Fig. 3E). M113I7 and M113T7, which are β-branched, are also weak binders, but Ile and Thr are less closely related to Val than Cpg. To determine whether M113V7 binds βCD in the same orien- tation as M113F7 or M113N7 (Fig. 2), we made heteromers of M113V and the M113N or M113F subunits. M113V3M113F4, M113V4M113F3, M113V3M113N4, and M113V4M113N3 were examined in detail. All four heteroheptamers bound βCD more weakly than M113V7, M113F7 or M113N7 (Fig. 3F and Table S4), suggesting that Val at position 113 interacts with βCD strongly but in a different manner to either Phe or Asn. Each heteromer exhibited a range of Kd values, perhaps reflecting the various pos- sible permutations of the two different subunits around the cen- tral axis of the heptamer, although this heterogeneity was not seen for heteromers made from WT, M113F and M113N (Fig. 1). Discussion Soon after we discovered that βCD binds to the WT-αHL pore for around a millisecond, we found a mutant pore, M113N7, that re- leases βCD ∼104 times more slowly (14). This prompted us to examine all 19 mutants in which residue 113 is replaced by a nat- ural amino acid, with the surprising result that a collection of ami- no acids with structurally unrelated side chains (V, H, Y, D, N, F) are tight binders (15). Here, we have examined the nature of the binding interactions more closely by single-channel electrical re- cording, protein engineering including unnatural amino acid mu- tagenesis, and high-resolution x-ray crystallography, and we provide the first definitive structural information on an engi- neered protein nanopore. We find that βCD can bind tightly to the αHL pore in three different ways depending on the residue at 113, as exemplified by Asn, Phe, and Val. Because Asn and Phe have quite different side chains, we first compared the ability of M113N and M113F subunits to take part in binding the CD. The examination of het- eromeric proteins containing WT (Met-113), M113N and M113F subunits showed that the replacement of WT subunits in WT7 with M113N or M113F subunits led to increased affinity for βCD. The more M113N or M113F subunits that were added, the tighter binding became. By contrast, when subunits in M113N7 were replaced with M113F subunits, binding became weaker, reaching a minimum at three to four M113F subunits, and then increasing in strength with five M113F subunits or more (Fig. 1C). Parallel structural studies (30) revealed the basis of the “oppos- ing” effects of the M113N and M113F subunits. βCD binds to M113N7 in the opposite orientation to that in which it binds to M113F7. In M113N7, the secondary hydroxyls in the βCD ring are hydrogen bonded to Lys-147 and Asn-113 (Fig. 2A). By con- trast, βCD interacts with M113F7 through its primary hydroxyl face (Fig. 2B). It seemed likely that M113V7, bound βCD in yet another way, and this was examined by forming heteromers between M113V and M113N or M113F. The presence of three or four subunits of either M113N or M113F greatly decreases the affinity of the pore for βCD (Fig. 3F), with an average koff of 7.3 × 101 s−1, indicating that a third binding mode is indeed operating Fig. 2. X-ray structures of M113N and M113F homoheptamers with βCD bound. (A) Side view of heptameric αHL. βCD binds in the blue highlighted region. (B) βCD bound to M113N7 (dotted lines indicate hy- drogen bonding). The side chains of Lys-147 are in pale brown and the side chains of Asn- 113 in yellow. (C) βCD bound to M113F7 (dotted lines indicate CH-π bonding). The side chains of Phe-113 are in yellow. The sec- ond βCD in the M113F7 · ðβCDÞ2 structure is hydrogen bonded to the top βCD in a head- to-head arrangement and has no apparent interactions with the protein. For both (B) and (C), four β strands were omitted from the barrel to give a better view. Banerjee et al. PNAS ∣ May 4, 2010 ∣ vol. 107 ∣ no. 18 ∣ 8167 BIOCHEMISTRY (Table S4). In summary, the three groups of tight-binding mutants comprise αHL pores incorporating, at position 113: (i) the hydro- gen-bonding amino acids N, D (the latter would have to be largely in the protonated form), and possibly H; (ii) the aromatics F, Y, f1F, and possibly H, and more weakly W; (iii) the β-branched ami- no acids V, Cpg. There may be yet other means by which CDs can bind to the αHL pore. For example, we earlier found that hepta- 6-sulfato-βCD can bind tightly to αHL pores containing the N139Q mutation (37). Presumably, this CD is bound at a site low- er down in the β barrel in a fashion that includes hydrogen bond- ing to the Gln at position 139. While the various mutants exhibited widely different koff values, the value of kon was almost invariant and averaged ∼2.3 × 105 M−1 s−1 (Table S2) (15). Ap- parently, transport to the binding site is rate limiting, through a route unaffected by mutagenesis. koff increased precipitously with the addition of WTsubunits to M113N7 (Fig. 1C). Crystal structures of M113N7 show that resi- dues 111, 113, and 147 are reorganized by compari- son with WT7 and then undergo a more limited rearrangement when βCD binds (Fig. S5). For example, the side chain of Lys-147 shifts position to form a bifurcated hydrogen bond with a 3-hydroxyl group of βCD and the side chain carbonyl of an Asn- 113 (Fig. S6). Therefore, the side chains of residues 111, 113, and 147 might be in a variety of conformations in WT7−nM113Nn het- eromers and offer less well preorganized binding sites for βCD than they do in M113N7. Further, the intramolecular hydrogen bonds of the secondary hydroxyls in βCD (38) must be disrupted upon binding as both hydroxyls on each glucose ring form hydro- gen bonds to the mutant subunits (Fig. 2B). Because the hydrogen bonds that are broken in βCD are arranged in a circle, the break- age of bonds involving a single glucose (three bonds in all) will be relatively more disruptive than those involving adjoining glucose residues or the entire circle. The overall binding cooperativity in M113N7 could be attributed to enthalpic cooperativity outweigh- ing entropic penalties to binding (39). Positive cooperativity has been observed previously in fairly rigid model systems (40). By contrast with M113N7, there is little movement of side chains in ðM113FÞ7 by comparison with WT7 and little move- ment, including Phe-113, upon binding βCD (Fig. S7A). Further, the crystal structure of M113F7 · βCD suggests that each Phe re- sidue interacts independently with the βCD through what appear to be CH-π interactions (Fig. S7B). These interactions are ex- pected to be weak and not strongly directional and hence offer less enthalpic cooperativity, as supported by the B-factors (crys- tallographic temperatures factors) at the primary βCD binding site, which are between ∼40 and 50. Positive cooperativity is ob- served, but it is less pronounced than in the case of M113N7 (Table S5). In the case of M113N7, the B-factors of the residues that bind βCD are in the 20s implying that the βCD is more rigidly held than it is in M113F7. The binding of sugars to aromatic residues in proteins can in- clude CH-π bonding (41) or OH-π bonding or a finely balanced complement of both (42, 43). However, we have dismissed the possibility of an OH-π interaction between Phe-113 and the 6-hydroxyl groups of βCD as the distance between the center of the phenyl rings to the nearest hydroxyl oxygen is higher (5.2  0.65 Å, n ¼ 7) than that expected for a favorable OH-π interaction (33). While we propose that βCD binds to Phe-113 through a C-6 CH-π interaction (Fig. S7B), the distances between the center of the Phe-113 ring and the nearest C-6 of βCD ob- served in the M113F7 · βCD structure (4.66  0.24 Å, n ¼ 7) are in the upper range of the expected distance for a strong inter- action, which is ∼4.5 Å (33). The angle between the normal to the aromatic rings and the line connecting the C-6 atoms to the aro- matic midpoint is 8.0  5.6°, which is well within the expected range (44). The measurements with M113f5F7 argue against a hydrophobic interaction between Phe residues at position 113 and the βCD ring. In f5F, the hydrophobicity of the phenyl ring is significantly increased (45) yet M113f5F7 binds βCD weakly, like WT7 (Fig. 3C and Table S2A). By contrast with F, f1F, Y and N, homomeric αHL pores with f5F and W at position 113 bound βCD relatively weakly (Fig. 3C and Table S2A). In the case of f5F, the powerful electron with- drawing action of the five fluorine atoms leaves a highly increased positive charge at the center of the ring (46, 47), mitigating against a hydrogen-bonding interaction. The electron-rich Trp Fig. 3. Properties of pores containing natural and unnatural amino acid sub- stitutions at position 113. The data were recorded at þ40 mV in 1.0 M NaCl, 10 mM sodium phosphate, pH 7.5. (A) Unnatural amino acids used in this study: 4-fluorophenylalanine, f1F; pentafluorophenylalanine, f5F; cyclohex- ylalanine, Cha; cyclopropylglycine, Cpg; cyclopropylalanine, Cpa. (B) Repre- sentative current traces from single homoheptameric αHL pores, containing unnatural amino acids at position 113, in the presence of βCD. βCD (40 μM final) was added to the trans chamber. Level 1, open pore current; level 2, pore occupied by βCD. The broken line indicates zero current. (C) In- teraction of βCD with homomeric αHL pores containing aromatic amino acids at position 113. Kd values for the interaction between βCD and the αHL pore were calculated by using Kd ¼ koff∕kon. Each column represents the mean  s:d: for 10 or more determinations: dark gray, natural amino acids; light gray, unnatural amino acids. Data adapted from Gu and colleagues (15) are marked (*). (D) Representative current traces from single-channel recordings of βCD binding to M113F4M113f1F3 and M113N4M113f1F3. βCD (40 μM final) was added to the trans chamber. The broken line indicates zero current. (E) Interaction of βCD with homomeric αHL pores containing hydrophobic amino acids at position 113. Kd values for the interaction between βCD and the αHL pore were calculated by using Kd ¼ koff∕kon. Each column represents the mean  s:d: for ten or more determinations: dark gray, natural amino acids; light gray, unnatural amino acids. Data adapted from Gu and colleagues (15) are marked (*). (F) koff values for βCD from heteroheptamers formed with M113F and M113V subunits and with M113N and M113V subunits. βCD (40 μM final) was added to the trans chamber. The kon values for βCD for all these mutants are similar, at ∼3 × 105 M−1 s−1. Empty square: average koff values for the mutant (bar is s:d). Filled square: M113V3M113F4; filled circle: M113V4M113F3; filled upright triangle: M113V3M113N4; filled in- verted triangle: M113V4M113N3. 8168 ∣ www.pnas.org/cgi/doi/10.1073/pnas.0914229107 Banerjee et al. ring (44, 46, 47) should favor hydrogen bonding, but here we can- not make a direct comparison with the crystal structure of M113F7 as the indole ring is far larger than benzene. It is possible that it cannot become oriented in the same manner and that it is misaligned for hydrogen bonding. Our experiments suggest that M113V7 and M113Cpg7 bind βCD in a third way. In heteromers with M113V, both M113F and M113N reduce the affinity of the pore for βCD suggesting that neither the CH-π interaction with Phe-113 nor the hydrogen- bonding interactions with Asn-113 and Lys-147 are compatible with binding to Val. Close interactions of Val with glucose rings have been noted previously (48). Therefore, we propose that the Val side-chain interacts with the side of the glucose ring. This in- teraction might occur in one or both orientations of the CD ring (Fig. 4). Conclusion We provide structural information on engineered protein nano- pores and describe three distinct ways in which βCD can bind within the lumen of mutant αHL pores in atomic detail. Our re- sults will be useful in several areas of basic science and biotech- nology. By using host molecules lodged within the αHL pore, host-guest interactions can be investigated in fine detail at the single-molecule level (17, 49). The present work will now permit us to examine binding events at a known face of a CD. The work also provides information for designing new binding sites within the lumen of the αHL pore (37) or within other β barrel proteins (21, 50) and for using molecular design to devise ways in which to covalently attach CDs within pores (13, 51). These areas impact practical applications of nanopore technology including stochas- tic sensing (8), single-molecule DNA sequencing (9, 12, 13, 52), the use of nanoreactors for the observation of single-molecule chemistry (10), and the construction of nano- and microdevices (11, 53), as well as the design of CDs as therapeutic agents (23, 24). Methods Full details of the experimental procedures can be found in SI Appendix. Materials L-Amino acids were obtained as follows: 4-fluorophenylalanine (f1F) (Fluka); pentafluorophenylalanine (f5F) (PepTech Corp.); cyclopropylglycine (Cpg) (Ty- ger); cyclopropylalanine (Cpa) (Tyger). 4-N-benzoyl-5′-O-(4,4′-dimethoxytri- tyl)-2′-deoxycytidine-3′-[(2-cyanoethyl)-(N,N-diisopropyl)]-phosphoramidite and bis(2-cyanoethyl)-N,N-diisopropylphosphoramidite for the synthesis of pdCpA were purchased from Glen Research and Toronto Research Chemicals, respectively. Preparation of NVOC-Protected Aminoacyl-pdCpA. NVOC-protected aminoacyl-pdCpAs were prepared as reported previously by reacting the dinucleotide pdCpA with N-protected, carboxylic acid-activated, amino acids (54–56). Preparation of NVOC-Protected Aminoacyl-tRNA. NVOC-protected aminoacyl- pdCpAs were ligated enzymatically with a truncated tRNA, prepared by using methods described elsewhere (57, 58). Genetic Constructs and Mutagenesis. All new αHL constructs were verified by DNA sequencing. Details of each construct can be found in SI Appendix. Synthesis, Assembly, and Purification of Mutant αHL pores. αHL monomers (WT and mutants) were prepared in vitro by coupled transcription and translation (IVTT) and assembled into homoheptamers on rabbit red blood cell membranes followed by purification by SDS–PAGE as described earlier (59). Heteroheptamers were prepared by mixing the two required DNAs (one encoding an αHL with a D8 tail) before IVTT and then oligomerizing the mixed translation products on rabbit red blood cell membranes. Pores with the desired combinations of subunits were purified by SDS–PAGE (59). Synthesis, Assembly, and Purification of αHL Mutants Containing Unnatural Ami- no Acids. αHL polypeptides containing unnatural amino acids were synthe- sized by IVTT in the presence of rabbit red blood cell membranes. The plasmid with a stop codon (TAG) at position 113 was used. Deprotected ami- noacyl-tRNAs (SI Appendix) were added to the IVTT mixtures. For heterohep- tamers with subunits containing unnatural amino acids in combination with M113N or M113F, monomers were first made, which were then coassembled on rabbit red blood cell membranes and subsequently purified by SDS–PAGE. Single-Channel Current Recordings in Planar Lipid Bilayers. (15, 60) Recordings were made with 1.0 M NaCl, 10 mM sodium phosphate, pH 7.5, in both cham- bers, at an applied potential of þ40 mV. Data were recorded at 22  2°C. The bilayer was formed from 1,2-diphytanoyl-sn-glycero-phosphocholine (Avanti Polar Lipids). Proteins were added to the cis chamber, and βCD to the trans chamber. Single-channel currents were recorded with an Axopatch 200B patch-clamp amplifier (Axon Instruments) and filtered at 2 kHz with a built-in 4-pole low-pass Bessel Filter. The data were acquired at a sampling rate of 10 kHz. For mutants that bind βCD strongly, the data were acquired for at least 30 min and for weak-binding mutants for at least 10 min. Kinetic Data Analysis. Current amplitude and dwell-time histograms were made by using ClampFit 9.0. The mean dwell times, τoff, were determined by fitting the dwell-time histograms to single exponentials. Values of kon and koff were obtained by using the mean dwell times and mean interevent intervals, as described previously (15, 60). This analysis assumes a binary in- teraction, which was supported in all cases examined by the finding of only one major blockade level and a single exponential distribution of dwell times (τoff). Fig. 4. Molecular model showing the three classes of interaction between the αHL pore and βCD identified in this work. The model identifies the region of βCD responsible for each interaction (H atoms interacting with Phe-113 or Asn-113 and Lys-147: gray). The first class of interaction is with aromatic residues and involves the seven -CH2OH groups of the βCD. The second class is typified by the interactions with Asn at position 113, which involve hydro- gen-bonds to the secondary 2-hydroxyls of the βCD. Structural studies show that this interaction is supported by hydrogen bonding between Lys-147 and the secondary 3-hydroxyls of the βCD. Structural studies and experiments with heteromers suggest that the βCD in M113F7 is in the opposite orienta- tion to the βCD in M113N7, in support of the model shown here. As the inter- action with Val is hydrophobic, it is not directional and βCD may not bind at the same position inside the β barrel as it does in M113F7 or M113N7. Banerjee et al. PNAS ∣ May 4, 2010 ∣ vol. 107 ∣ no. 18 ∣ 8169 BIOCHEMISTRY Protein Crystallography. Details can be found in SI Appendix. Protein Data Bank: The coordinates and structure factors of the described structures have been deposited with accession codes 3M2L ðM113F7Þ, 3M3R ðM113F7 · βCDÞ, 3M4D ðM113N7Þ, 3M4E ðM113N7 · βCDÞ. ACKNOWLEDGMENTS. We thank Dennis Dougherty for the plasmid pTHG73. This work was funded by a Royal Society Wolfson Research Merit Award (to H.B.), the Medical Research Council (H.B.), the National Institutes of Health (H.B.), and the Howard Hughes Medical Institute (E.G.). 1. Lu Y, Yeung N, Sieracki N, Marshall NM (2009) Design of functional metalloproteins. Nature 460:855–862. 2. Gebauer M, Skerra A (2009) Engineered protein scaffolds as next-generation antibody therapeutics. Curr Opin Chem Biol 13:245–255. 3. Gronwall C, Stahl S (2009) Engineered affinity proteins—Generation and applications. J Biotechnol 140:254–269. 4. Arnold U (2009) Incorporation of non-natural modules into proteins: Structural features beyond the genetic code. Biotechnol Lett 31:1129–1139. 5. Tracewell CA, Arnold FH (2009) Directed enzyme evolution: Climbing fitness peaks one amino acid at a time. Curr Opin Chem Biol 13:3–9. 6. Fruk L, Kuo CH, Torres E, Niemeyer CM (2009) Apoenzyme reconstitution as a chemical tool for structural enzymology and biotechnology. Angew Chem Int Ed Engl 48:1550–1574. 7. Bayley H, Jayasinghe L (2004) Functional engineered channels and pores. Mol Membr Biol 21:209–220. 8. Bayley H, Cremer PS (2001) Stochastic sensors inspired by biology. Nature 413:226–230. 9. Branton D, et al. (2008) The potential and challenges of nanopore sequencing. Nature Biotechnol 26:1146–1153. 10. Bayley H, Luchian T, Shin S-H, Steffensen MB (2008) Single Molecules and Nanotech- nology, eds R Rigler and H Vogel (Springer, Heidelberg), pp 251–277. 11. Maglia G, et al. (2009) Droplet networks with incorporated protein diodes show collective properties. Nat Nanotechnol 4:437–440. 12. Astier Y, Braha O, Bayley H (2006) Toward single molecule DNA sequencing: Direct identification of ribonucleoside and deoxyribonucleoside 5'-monophosphates by using an engineered protein nanopore equipped with a molecular adapter. J Am Chem Soc 128:1705–1710. 13. Clarke J, et al. (2009) Continuous base identification for single-molecule nanopore DNA sequencing. Nature Nanotechnol 4:265–270. 14. Gu L-Q, Braha O, Conlan S, Cheley S, Bayley H (1999) Stochastic sensing of organic analytes by a pore-forming protein containing a molecular adapter. Nature 398:686–690. 15. Gu L-Q, Cheley S, Bayley H (2001) Prolonged residence time of a noncovalent molecular adapter, β-cyclodextrin, within the lumen of mutant α-hemolysin pores. J Gen Physiol 118:481–494. 16. Ervin EN, Kawano R, White RJ, White HS (2008) Simultaneous alternating and direct current readout of protein ion channel blocking events using glass nanopore membranes. Anal Chem 80:2069–2076. 17. Gurnev PA, Harries D, Parsegian VA, Bezrukov SM (2009) The dynamic side of the Hof- meister effect: A single-molecule nanopore study of specific complex formation. ChemPhysChem 10:1445–1449. 18. Mamonova T, Kurnikova M (2006) Structure and energetics of channel-forming protein-polysaccharide complexes inferred via computational statistical thermody- namics. J Phys Chem B 110:25091–25100. 19. Egwolf B, Luo Y, Walters DE, Roux B (2010) Ion selectivity of alpha-hemolysin with beta-cyclodextrin adapter. II. Multi-ion effects studied with grand canonical monte carlo/brownian dynamics simulations. J Phys Chem B 114:2901–2909. 20. Orlik F, et al. (2003) CymA of Klebsiella oxytoca outer membrane: Binding of cyclodex- trins and study of the current noise of the open membrane. Biophys J 85:876–885. 21. Chen M, Khalid S, Sansom MSP, Bayley H (2008) OmpG: Engineering a quiet pore for biosensing. Proc Natl Acad Sci USA 105:6272–6277. 22. Locke D, Koreen IV, Liu JY, Harris AL (2004) Reversible pore block of connexin channels by cyclodextrins. J Biol Chem 279:22883–22892. 23. Karginov VA, Nestorovich EM, Moayeri M, Leppla SH, Bezrukov SM (2005) Blocking anthrax lethal toxin at the protective antigen channel by using structure-inspired drug design. Proc Natl Acad Sci USA 102:15075–15080. 24. Moayeri M, Robinson TM, Leppla SH, Karginov VA (2008) In vivo efficacy of beta- cyclodextrin derivatives against anthrax lethal toxin. Antimicrob Agents Chemother 52:2239–2241. 25. Gao C, Ding S, Tan Q, Gu LQ (2009) Method of creating a nanopore-terminated probe for single-molecule enantiomer discrimination. Anal Chem 81:80–86. 26. Pregel MJ, Jullien L, Lehn J-M (1992) Towards artificial ion channels: Transport of alkali metal ions across liposomal membranes by “bouquet” molecules. Angew Chem Int Edit 31:1637–1639. 27. Bacri L, Benkhaled A, Guegan P, Auvray L (2005) Ionic channel behavior of modified cyclodextrins inserted in lipid membranes. Langmuir 21:5842–5846. 28. Jog PV, Gin MS (2008) A light-gated synthetic ion channel. Org Lett 10:3693–3696. 29. Howorka S, Cheley S, Bayley H (2001) Sequence-specific detection of individual DNA strands using engineered nanopores. Nat Biotechnol 19:636–639. 30. Montoya M (2004) Insights into Membrane Association and Bioengineering of a Pore- Forming Toxin: Structural Studies of Staphylococcal α-Hemolysin (Columbia University, New York). 31. Steiner T (2002) The hydrogen bond in the solid state. Angew Chem Int Ed 41:49–76. 32. Steiner T (2002) Hydrogen bonds from water molecules to aromatic acceptors in very high-resolution protein crystal structures. Biophys Chem 95:195–201. 33. Steiner T, Koellner G (2001) Hydrogen bonds with pi-acceptors in proteins: Frequencies and role in stabilizing local 3D structures. J Mol Biol 305:535–557. 34. Brandl M, Weiss MS, Jabs A, Sühnel J, Hilgenfeld R (2001) CH...π-interactions in proteins. J Mol Biol 307:357–377. 35. Weiss MS, Brandl M, Suhnel J, Pal D, Hilgenfeld R (2001) More hydrogen bonds for the (structural) biologist. Trends Biochem Sci 26:521–523. 36. Wang L, Xie J, Schultz PG (2006) Expanding the genetic code. Annu Rev Biophys Biomol Struct 35:225–249. 37. Gu L-Q, Cheley S, Bayley H (2001) Capture of a single molecule in a nanocavity. Science 291:636–640. 38. Saenger W, et al. (1998) Structures of the common cyclodextrins and their larger analogues—Beyond the doughnut. Chem Rev 98:1787–1802. 39. Hunter CA, Tomas S (2003) Cooperativity, partially bound states, and enthalpy-entropy compensation. Chem Biol 10:1023–1032. 40. Bisson AP, Hunter CA (1996) Cooperativity in the assembly of zipper complexes. Chem Commun 1723–1724. 41. del Carmen Fernandez-Alonso M, Canada FJ, Jimenez-Barbero J, Cuevas G (2005) Molecular recognition of saccharides by proteins. Insights on the origin of the carbohydrate-aromatic interactions. J Am Chem Soc 127:7379–7386. 42. Jimenez-Barbero J, Asensio JL, Canada FJ, Poveda A (1999) Free and protein-bound carbohydrate structures. Curr Opin Struct Biol 9:549–555. 43. Stanca-Kaposta EC, et al. (2007) Carbohydrate molecular recognition: A spectroscopic investigation of carbohydrate-aromatic interactions. Phys Chem Chem Phys 9:4444–4451. 44. Brandl M, Weiss MS, Jabs A, Suhnel J, Hilgenfeld R (2001) CH...π-interactions in proteins. J Mol Biol 307:357–377. 45. Woll MG, Hadley EB, Mecozzi S, Gellman SH (2006) Stabilizing and destabilizing effects of phenylalanine –>F5-phenylalanine mutations on the folding of a small protein. J Am Chem Soc 128:15932–15933. 46. Mecozzi S, West AP, Jr, Dougherty DA (1996) Cation-pi interactions in aromatics of biological and medicinal interest: Electrostatic potential surfaces as a useful qualitative guide. Proc Natl Acad Sci USA 93:10566–10571. 47. Dougherty DA (2008) Cys-loop neuroreceptors: Structure to the rescue?. Chem Rev 108:1642–1653. 48. Hondoh H, et al. (2008) Substrate recognition mechanism of alpha-1,6-glucosidic linkage hydrolyzing enzyme, dextran glucosidase from Streptococcus mutans. J Mol Biol 378:913–922. 49. Kang XF, Cheley S, Guan X, Bayley H (2006) Stochastic detection of enantiomers. J Am Chem Soc 128:10684–10685. 50. Chen M, Li QH, Bayley H (2008) Orientation of the monomeric porin OmpG in planar lipid bilayers. ChemBioChem 9:3029–3036. 51. Wu H-C, Astier Y, Maglia G, Mikhailova E, Bayley H (2007) Protein nanopores with covalently attached molecular adapters. J Am Chem Soc 129:16142–16148. 52. Bayley H (2006) Sequencing single molecules of DNA. Curr Opin Chem Biol 10:628–637. 53. Astier Y, Bayley H, Howorka S (2005) Protein components for nanodevices. Curr Opin Chem Biol 9:576–584. 54. Robertson SA, Noren CJ, Anthony-Cahill SJ, Griffith MC, Schultz PG (1989) The use of 5'-phospho-2 deoxyribocytidylylriboadenosine as a facile route to chemical aminoacy- lation of tRNA. Nucleic Acids Res 17:9649–9660. 55. Ellman J, Mendel D, Anthony-Cahill S, Noren CJ, Schultz PG (1991) Biosynthetic method for introducing unnatural amino acids site-specifically into proteins. Method Enzymol 202:301–336. 56. Kearney PC, et al. (1996) Dose-response relations for unnatural amino acids at the agonist binding site of the nicotinic acetylcholine receptor: Tests with novel side chains and with several agonists. Mol Pharmacol 50:1401–1412. 57. England TE, Bruce AG, Uhlenbeck OC (1980) Specific labeling of 3′ termini of RNA with T4 RNA ligase. Method Enzymol 65:65–74. 58. Nowak MW, et al. (1998) In vivo incorporation of unnatural amino acids into ion channels in xenopus oocyte expression system. Method Enzymol 293:504–529. 59. Cheley S, Braha O, Lu X, Conlan S, Bayley H (1999) A functional protein pore with a “retro” transmembrane domain. Protein Sci 8:1257–1267. 60. Gu L-Q, et al. (2000) Reversal of charge selectivity in transmembrane protein pores by using non-covalent molecular adapters. Proc Natl Acad Sci USA 97:3959–3964. 61. Song L, et al. (1996) Structure of staphylococcal α-hemolysin, a heptameric transmem- brane pore. Science 274:1859–1865. 8170 ∣ www.pnas.org/cgi/doi/10.1073/pnas.0914229107 Banerjee et al.
3M2R
Structural Insight into Methyl-Coenzyme M Reductase Chemistry using Coenzyme B Analogues
Structural Insight into Methyl-Coenzyme M Reductase Chemistry using Coenzyme B Analogues,†,‡ Peder E. Cedervall§,#, Mishtu Dey||,#,⊥, Arwen R. Pearson§,⊥, Stephen W. Ragsdale||, and Carrie M. Wilmot*,||,§ § Department of Biochemistry, Molecular Biology and Biophysics, University of Minnesota, Minneapolis, Minnesota 55455 || Department of Biological Chemistry, University of Michigan, Ann Arbor, Michigan 48109 Abstract Methyl-coenzyme M reductase (MCR) catalyzes the final and rate-limiting step in methane biogenesis; the reduction of methyl-coenzyme M (methyl-SCoM) by coenzyme B (CoBSH) to methane and a heterodisulfide (CoBS-SCoM). Crystallographic studies show that the active site is deeply buried within the enzyme, and contains a highly reduced nickel-tetrapyrrole, coenzyme F430. Methyl-SCoM must enter the active site prior to CoBSH, as species derived from analogues of methyl-SCoM are always observed bound to the F430 nickel in the deepest part of the 30 Å long substrate channel that leads from the protein surface to the active site. The seven-carbon mercaptoalkanoyl chain of CoBSH binds within a 16 Å predominantly hydrophobic part of the channel close to F430, with the CoBSH thiolate lying closest to the nickel at a distance of 8.8 Å. It has previously been suggested that binding of CoBSH initiates catalysis by inducing a conformational change that moves methyl-SCoM closer to the nickel promoting cleavage of the C- S bond of methyl-SCoM. In order to better understand the structural role of CoBSH early in the MCR mechanism, we have determined crystal structures of MCR in complex with four different CoBSH analogues; pentanoyl-, hexanoyl-, octanoyl- and nonanoyl- derivatives of CoBSH (CoB5SH, CoB6SH, CoB8SH and CoB9SH respectively). The data presented here reveal that the shorter CoB5SH mercaptoalkanoyl chain overlays with that of CoBSH, but terminates two units short of the CoBSH thiolate position. In contrast, the mercaptoalkanoyl chain of CoB6SH adopts a different conformation, such that its thiolate is coincident with the position of the CoBSH thiolate. †This work was supported by a Department of Energy grant DE-FG02-08ER15931 to S.W.R and supplement to C.M.W. and a Minnesota Partnership for Biotechnology and Medical Genomics grant SPAP-05-0013-P-FY06. ‡Co-ordinates and structure factors have been deposited in the Protein Data Bank as entries 3m1v (MCRHSCoM); 3m2r (MCRCoB5SH); 3m2u (MCRCoB6SH); 3m2v (MCRCoB8SH); 3m30 (MCRCoB9SH). *Address correspondence to: Carrie M. Wilmot, Tel: (612) 624-2406, [email protected]. ⊥Current addresses: Mishtu Dey, Department of Chemistry, Massachusetts Institute of Technology, 77 Massachusetts Ave., Cambridge, Massachusetts 02139; Arwen Pearson, Astbury Centre for Structural Molecular Biology, Institute for Molecular and Cellular Biology, Astbury Building. Leeds, LS2 9JT, U.K. #These authors contributed equally to this work. Supporting Informational material is available free of charge via the Internet at http://pubs.acs.org and contains the following: MATERIAL AND METHODS for single crystal UV-visible microspectrophotometry, X-ray photoreduction experiment and X-ray crystallography of the MCR-heterodisulfide product complex (MCRCoBSH + methyl-SCoM); RESULTS AND DISCUSSION for redox changes and MCRCoBSH + methyl-SCoM crystal structure; Table S1, X-ray Data Collection, Processing and Refinement Statistics for MCRCoBSH + methyl-SCoM; Figure S1, use of Fo-Fc electron density in modelling MCRCoB8SH; Figure S2, illustration of electron density quality of MCRCoB5SH; Figure S3, solution and single crystal UV-visible spectra; Figure S4, modeling of CoB6SH and CoBSH into the electron density of MCRCoB6SH; Figure S5, partially occupied HSCoM in MCRCoB6SH; Figure S6, the active site and substrate channel of MCRCoBSH + methyl-SCoM; Figure S7, alternative conformation of Valα482 in MCRHSCoM; Figure S8, the two conformations of the gly-rich loop in MCRHSCoM; Figure S9, propagation of conformational changes in MCRHSCoM; Figure S10, the two conformations of CoB9SH in MCRCoB9SH; Figure S11, EPR spectra of MCRred1 sample; Scheme S1, scheme of the characterized forms of MCR. NIH Public Access Author Manuscript Biochemistry. Author manuscript; available in PMC 2011 September 7. Published in final edited form as: Biochemistry. 2010 September 7; 49(35): 7683–7693. doi:10.1021/bi100458d. NIH-PA Author Manuscript NIH-PA Author Manuscript NIH-PA Author Manuscript This is consistent with the observation that CoB6SH is a slow substrate. A labile water in the substrate channel was found to be a sensitive indicator for the presence of CoBSH and HSCoM. The longer CoB8SH and CoB9SH analogues can be accommodated in the active site through exclusion of this water. These analogues react with Ni(III)-methyl; a proposed MCR catalytic intermediate of methanogenesis. The CoB8SH thiolate is 2.6 Å closer to the nickel than that of CoBSH, but the additional carbon of CoB9SH only decreases the nickel thiolate distance a further 0.3 Å. Although the analogues did not induce any structural changes in the substrate channel, the thiolates appeared to preferentially bind at two distinct positions in the channel; one being the previously observed CoBSH thiolate position, and the other being at a hydrophobic annulus of residues that lines the channel proximal to the nickel. INTRODUCTION Methanogenic archaea are organisms that under strict anaerobic conditions derive energy by reducing compounds such as carbon dioxide, methylamine, acetate and methanol, to methane (1, 2). The global production of methane by these organisms is estimated at one billion tons annually. Microbially produced methane is not only a potential source of renewable energy but also a potent greenhouse gas, and as such study of this process has environmental ramifications. In these microorganisms, methyl-coenzyme M reductase (MCR)1 is the enzyme that catalyzes the final step in methanogenesis, in which the substrates methyl-coenzyme M (methyl-SCoM, 2-(methylthio)ethanesulfonate) and coenzyme B (CoBSH, N-7-mercaptoheptanoylthreonine phosphate) are converted to methane and a heterodisulfide (CoBS-SCoM) (Scheme 1) (3). MCR is a 300 kDa protein with six subunits arranged in a α2β2γ2 oligomer (4). The known crystal structures show that MCR has two active sites approximately 50 Å apart that are deeply buried within the enzyme (5). The active site pocket is comprised of residues from subunits α, α′, β and γ, with a 30 Å long substrate channel leading to the enzyme surface (Figure 1). At the heart of the active site pocket is coenzyme F430, which is a highly reduced nickel-containing tetrapyrrole (6–8). Currently sixteen distinct enzymatic and complexed states of MCR have been spectroscopically characterized (Supporting Information, Scheme S1) (9–31). In the resting active state of the enzyme, denoted MCRred1, the redox-active nickel of F430 is present in the Ni(I) state (9, 16, 32). MCR is extremely oxygen sensitive and upon oxygen exposure the enzyme enters an inactive Ni(II) state, denoted MCRred1-silent (6). In this state it cannot be converted back to the active Ni(I) form by any known reducing agent making this a challenging system to study. Additional complications involve the tight association of coenzymes to purified MCR that are not easily displaced as demonstrated by X-ray crystallographic and kinetic studies (5, 33–35). Despite the fact that MCR has been studied for decades, no true catalytic intermediate has been observed, and the actual mechanism remains elusive. Currently three general mechanistic schemes for the enzymatic reaction have been proposed, each of which posit different chemistry to initiate catalysis. Mechanism I involves Ni(I) acting as a nucleophile in an SN2-type reaction that generates Ni(III)-methyl as an intermediate (Scheme 2A) (35– 38). Mechanism II starts with methyl-SCoM undergoing homolytic cleavage at the Ni(I) to generate a methyl radical and a Ni(II)-SCoM species (Scheme 2B) (39–41). A more recently proposed mechanism III suggests protonation of coenzyme F430 promotes reductive cleavage of the methyl-SCoM thioether bond (42). 1Abbreviations: MCR, Methyl-coenzyme M reductase; methyl-SCoM, methyl-coenzyme M; CoBSH, coenzyme B; HSCoM, coenzyme M; CoBS-SCoM, heterodisulfide of coenzyme B and coenzyme M; APS, Advanced Photon Source; ASU, asymmetric unit; BPS, bromopropanesulfonate. Cedervall et al. Page 2 Biochemistry. Author manuscript; available in PMC 2011 September 7. NIH-PA Author Manuscript NIH-PA Author Manuscript NIH-PA Author Manuscript Due to the stringent requirement to exclude O2, the available MCR crystal structures are all in the inactive Ni(II) state. All but one contain CoBSH and HSCoM (demethylated methyl- SCoM, an inhibitor and substrate analogue) in the active site (PDB codes 1hbn, 1hbo, 1hbu, 1e6y, 1e6v) (5, 33, 34). Another crystal structure has bound heterodisulfide product, CoBS- SCoM (MCRsilent, PDB code 1hbm, Scheme 1 and Supporting Information, Scheme S1) (5, 33). All these structures reveal that both substrates access the active site through the same channel (Figure 1). The binding site of HSCoM (and presumably methyl-SCoM) is more deeply buried within the enzyme, and so it must enter prior to CoBSH for productive chemistry to occur. As binding of CoBSH in the absence of co-substrate would be inhibitory, it was suggested that a conformational change upon methyl-SCoM binding might lower the Kd for CoBSH, and thus promote an ordered mechanism. Furthermore it has been suggested that CoBSH binding induces a conformational change that brings the methyl- SCoM substrate into closer proximity to the nickel, and this promotes C-S bond cleavage. To investigate the proposed structural role of CoBSH in initiating catalysis, we have solved the X-ray crystal structures of MCR in complex with four different CoBSH analogues. CoBSH has a heptanoyl moiety linked to the thiol group, and the analogues are pentanoyl-, hexanoyl-, octanoyl- or nonanoyl-containing derivatives of CoBSH (CoB5SH, CoB6SH, CoB8SH and CoB9SH respectively; Figure 2) (3, 35, 43–47). In addition, we present a structure in which the substrate channel predominantly lacks either CoBSH or heterodisulfide product. MATERIALS AND METHODS Materials The organism Methanothemobacter marburgensis (catalog OCM82) was obtained from the Oregon Collection of Methanogens (Portland, OR). All buffers and media reagents were obtained from Sigma-Aldrich (St. Louis, MO). The gases N2 (99.98%), H2/CO2 (80%/20%), and ultra high purity H2 (99.999%) were obtained from Cryogenic Gases (Grand Rapids, MI). Ti(III) citrate solutions were prepared from a stock solution of 246 mM Ti(III) citrate, which was synthesized by adding sodium citrate to Ti(III) trichloride (30 wt % solution in 2 N hydrochloric acid) (Acros Organics, Morris Plains, NJ) under anaerobic conditions and adjusting pH to 7.0 with sodium bicarbonate (48). The concentration of Ti(III) citrate was determined by titrating against a solution of methyl viologen. Synthesis of methyl-SCoM, CoB5SH, CoB6SH, CoBSH, CoB8SH, and CoB9SH Methyl-SCoM was prepared from HSCoM and methyl iodide (49). The homodisulfides, CoB5S-SCoB5, CoB6S-SCoB6, CoBS-SCoB, CoB8S-SCoB8, CoB9S-SCoB9 were prepared as described from 5-bromovaleric acid, 6-bromohexanoic acid (Sigma-Aldrich, St. Louis, MO), 7-bromoheptanoic acid (Karl Industries, Aurora, OH), 8-bromooctanoic acid, and 9- bromononanoic acid (Matrix Scientific, Columbia, SC), respectively (43, 46). The free thiol forms of CoB5SH, CoB6SH, CoBSH, CoB8SH, and CoB9SH were generated by the reduction of the homodisulfides as previously described (45). The purity of the CoBSH analogues was determined by 1H NMR spectroscopy. All compounds synthesized were stored in a Vacuum Atmospheres chamber maintained at an oxygen level below 1 ppm, as monitored continually with an oxygen analyzer (model 317, Teledyne Analytical Instruments, City of Industry, CA) until use. M. marburgensis Growth and MCRred1 Purification Buffer preparations and all manipulations were performed under strict anaerobic conditions in a Vacuum Atmospheres chamber maintained at an oxygen level below 1 ppm, as monitored continually with an oxygen analyzer (model 317, Teledyne Analytical Instruments, City of Industry, CA). MCRred1 was isolated from M. marburgensis cultured on Cedervall et al. Page 3 Biochemistry. Author manuscript; available in PMC 2011 September 7. NIH-PA Author Manuscript NIH-PA Author Manuscript NIH-PA Author Manuscript H2/CO2 (80%/20%) at 65 °C in a 14-L fermentor (New Brunswick Scientific Co., Inc. New Brunswick, NJ). Culture media were prepared as previously described (20, 24). MCRred1 was generated in vivo and purified as described previously (20). The purification procedure routinely generates 65–75% MCRred1 as determined by UV-visible and EPR spectroscopy. Spectroscopy of MCR UV-visible spectra of the Ni(I)-containing MCRred1 were recorded in the anaerobic chamber using a spectrophotometer (Model USB4000-UV-VIS, Ocean Optics, Dunedin, FL). EPR spectra were recorded on a Bruker EMX spectrometer (Bruker Biospin Corp., Billerica, MA), equipped with an Oxford ITC4 temperature controller, a Hewlett-Packard Model 5340 automatic frequency counter and Bruker gaussmeter. The EPR spectroscopic parameters included: temperature, 70 K; microwave power, 10 mW; microwave frequency, 9.43 GHz; receiver gain, 2 × 104; modulation amplitude, 10.0 G; modulation frequency, 100 kHz. Double integrations of the EPR spectra were performed and referenced to a 1 mM copper perchlorate standard. The NMR data were acquired at 298 K on a Bruker Avance DRX 500 MHz instrument equipped with a TXI cryoprobe. Preparation of MCRred1 for Crystallization All crystallization experiments were performed in the anaerobic chamber in which MCR was purified unless otherwise noted. MCRred1 was prepared in 50 mM Tris, pH 7.6 and excess HSCoM was removed by buffer exchange using an Amicon Ultra centrifuge filter with a 50 kDa cut-off membrane (Millipore). Typically, 2–3 ml of MCRred1 was exchanged with 10–15 ml of 50 mM Tris, pH 7.6. The enzyme was concentrated to 500–600 μl, and this process was repeated three times. The fraction of MCRred1 in the purified MCR sample was calculated from the UV-visible spectrum using extinction coefficients of 27.0 mM−1cm−1 at 385 nm for Ni(I)-MCRred1, and 9.15 mM−1cm−1 at 420 nm for Ni(II)- MCRred1-silent (20). The amount of MCRred1 in samples used for crystallization was determined to be ~80% and the concentration of total enzyme used was in the range of about 120–150 μM (~32–40 mg/ml). All crystallization experiments were performed anaerobically by incubating 2.0 μl of enzyme solution in 50 mM Tris, pH 7.5 and 2.0 μl of reservoir solution (100 mM Hepes-Na, pH 7.3/7.5/8.0; 150 mM magnesium acetate (Mg(CH3COO)2), and 20/22% (w/v) PEG 400) in a sitting drop over 1 ml reservoir solution at 9 ºC. Triangular and rectangular prismatic crystals with a bright yellowish-green color confirmed the presence of nickel coenzyme F430. The crystals grew to a size of approximately 100–200 μm in 4–5 days. CoBSH-depleted crystals were obtained by incubating 2 μl of a reaction mixture containing 139 μM MCRred1 and 13 mM HSCoM with 2 μl of reservoir solution (100 mM Hepes-Na pH 7.5, 150 mM magnesium acetate (Mg(CH3COO)2), 22% PEG 400). Crystals of MCR complexed with the CoBSH analogues were grown by co-crystallization. The CoB5SH co-crystals were obtained by incubating 2μl enzyme solution containing 124 μM MCRred1, 10 mM methyl-SCoM and 1 mM CoB5SH with 2μl of reservoir solution (100 mM Hepes-Na, pH 7.5, 150 mM (Mg(CH3COO)2), 22 % PEG 400). The crystals with bound CoB6SH and CoB9SH were obtained by co-crystallization of 1 mM of analogue with 142 μM MCRred1 and equilibrated with 2μl of reservoir solution (100 mM Hepes-Na, pH 7.5, 150 mM Mg(CH3COO)2, 20 % PEG 400 for CoB6SH and 100 mM Hepes-Na, pH 7.3, 150 mM Mg(CH3COO)2), 22 % PEG 400 for CoB9SH). Crystals were cryoprotected in reservoir solution containing 25 % (v/v) PEG 400 by soaking for 2–5 minutes before cryocooling in liquid nitrogen in the anaerobic chamber. Crystals of CoB8SH bound to MCR were obtained by incubating 2 μl of a mixture of 119 μM MCRred1 and 1 mM CoB8SH with 2 μl of reservoir solution (100 mM Hepes-Na, pH 7.3, 150 mM Mg(CH3COO)2), 20 % PEG 400). Before cryoprotection, the crystals were soaked for 5–10 minutes in a 100 mM solution of methyl iodide2. The methyl iodide solution used for soaking was prepared by adding a concentrated stock of methanolic solution of methyl iodide to the reservoir Cedervall et al. Page 4 Biochemistry. Author manuscript; available in PMC 2011 September 7. NIH-PA Author Manuscript NIH-PA Author Manuscript NIH-PA Author Manuscript solution. Soaked crystals were quickly cryoprotected as described above and cryocooled in the anaerobic chamber. X-ray Diffraction Data Collection, Processing and Refinement X-ray diffraction data were collected at 100 K on a ADSC Quantum-315 detector at the APS Beamline 14-BM-C (BioCARS). The wavelength of X-rays was 0.979 Å. Data were processed using HKL2000 (50). As in the previous X-ray crystallographic studies, the crystals belong to the monoclinic space group P21 (a = 82 Å, b = 118 Å, c = 122 Å, β = 92°), with one MCR molecule (two active sites) per asymmetric unit (5, 33). For refinement, REFMAC in the Collaborative Computational Project Number 4 (CCP4) program suite was used (51). A random sample of 5 % of the data across all resolution shells was chosen to check refinement progress through calculation of an Rfree. The same reflections were used to calculate Rfree for all structures, thus preventing bias due to high structural identity. The remaining reflections were used in refinement (Rwork). Model building was done using the Crystallographic Object-Oriented Toolkit (COOT) (52). The diffraction data and their models are designated as MCRCoBXSH, where X is the number of carbons in the alkanoyl portion of the analogue. Library files in CCP4 for F430 and CoBSH were incorrect, and these were modified in Monomer Library Sketcher in the CCP4 program suite by comparison with schematic drawings from Grabarse et al. (33). Co-ordinate and CCP4 library files for the different CoBSH analogues were created in Monomer Library Sketcher. The general model building and refinement strategy for all structures was as follows. It was clear from the electron density in the substrate channel and at the active site that mixtures of species were present in all datasets. These could be visualized with Fo-Fc and Fo-Fo difference electron density maps (Supporting Information, Figure S1). The known positions of CoBSH and HSCoM from the published Ni(II)-MCR crystal structures (PDB codes 1hbn, 1hbo, 1hbu (33)) were used as guides to determine which species could be present in each dataset, and these were then simultaneously modeled into the electron density. By alteration of their relative occupancies (in 10% increments) followed by refinement, the ratio of occupancy between different species was determined using the assumption that the average B-factors for all molecular species bound should be similar to that of F430 and adjacent well-ordered protein atoms within the active site and substrate channel. The combinations of modeled ligands were constantly reassessed throughout refinement based on the remaining difference electron density. This included test refinements of different ligand combinations during the latter stages, thus using the optimized phases to check whether a different combination of ligands could also explain the electron density. Sensible chemical structures and interactions, along with keeping the combined occupancies of sterically mutually exclusive species ≤ 100%, were maintained throughout refinement. The model was finally accepted when the difference electron density map was minimal and the B-factors for the models converged. In practice the first structure refined was that of MCRCoB5SH. Initial phases were generated by difference Fourier using a previously determined crystal structure (PDB code 1mro (5)) but with all non-bonded molecules, including water, removed from the model except F430. Initial rigid body refinement followed by restrained refinement of MCRCoB5SH reduced the Rwork to 26.5%. After model building and subsequent rounds of restrained refinement the Rwork was 14.3% (Rfree 16.6%). Of the five structures only the CoB5SH analogue is completely coincident with CoBSH, and so particular care had to be used in teasing apart the ratios of the two species in modeling the MCRCoB5SH electron density. This was done by 2Methyl iodide was added with the intention of creating a Ni(III)-methyl species, which CoB8SH stabilizes. This was not achieved, but the diffraction quality of this crystal was significantly better than crystals co-crystallized with CoB8SH alone, and so has been included in this study. Cedervall et al. Page 5 Biochemistry. Author manuscript; available in PMC 2011 September 7. NIH-PA Author Manuscript NIH-PA Author Manuscript NIH-PA Author Manuscript initially refining 100% CoBSH in the substrate channel. Positive Fo-Fc difference density located at the carbon where the shorter CoB5SH thiol might be expected to be indicated the presence of a more electron-rich species than carbon, which is consistent with the presence of the CoB5SH sulfur. The refinement converged at a model containing a 50:50 mix of CoBSH and the analogue. However, positive Fo-Fc difference density was still present at the position of the CoBSH thiol, therefore a water molecule was added to the CoB5SH model at 50% occupancy and upon refinement this accounted for the electron density. An illustration of the electron density quality from this structure is shown in Supporting Information, Figure S2. An HSCoM-, CoBSH- and CoBSH analogue-free version of the refined MCRCoB5SH structure was used as the starting model to generate initial phases for the four other structures. After the initial round of restrained refinement the Rwork for these structures were reduced to 14.5–15.6 %. RESULTS AND DISCUSSION Crystal Structures of MCR Five crystal structures were determined, four of which are in complex with CoBSH analogues differing in the number of carbons atoms in the alkanoyl portion of the molecule. CoBSH is an N-7-mercaptoheptanoyl-containing molecule, whereas the four CoBSH analogues contain N-5-mercaptopentanoyl-, N-6-mercaptohexanoyl-, N-8-mercaptooctanoyl- or N-9-mercaptononanoyl-moieties (Figure 2). The corresponding crystal structures are designated as MCRCoBXSH, where X is the number of carbons in the alkanoyl portion of the analogue. The other crystal structure is of MCRred1c-silent (MCR in the Ni(II) state in complex with HSCoM, designated here as MCRHSCoM) that is CoBSH-depleted. The datasets have resolutions in the range from 1.35 – 1.8 Å. Although the crystallizations were set up with the MCR solution predominantly in the Ni(I)-MCRred1 state, by the time X-ray diffraction data were collected they had been oxidized to the Ni(II)-MCRred1-silent state (Supporting Information). Following data collection there was no evidence for photoreduction of the Ni(II) back to Ni(I) in any of the crystals, as assessed by single crystal UV-visible microspectrophotometry (Supporting Information and Figure S3). Attempts to photoreduce the crystals using different wavelengths and temperatures were unsuccessful (Supporting Information). Overall, the resulting structures are very similar to each other and to the previously published structures of MCR, with differences mainly localized to the active site and substrate channel. The two active sites in the ASU were refined independently. Unless otherwise stated there was no difference between them. All five datasets contain a mixture of species bound to the enzyme. There is always a background of CoBSH and HSCoM, which co-purify with MCR and cannot be fully removed by extensive buffer exchange or by the addition of a CoBSH analogue. HSCoM is added during purification of MCR, as it stabilizes the resting active Ni(I) state (unpublished data), and this leads to HSCoM occupancies between 50–100% amongst the structures (Table 1). In contrast CoBSH, which is not added during purification, has occupancies ranging from 30–50%. As these confounding species have all been described at high occupancy in other crystallographic studies, the structural data of interest could be isolated (5, 33). In each case, the additional electron density could be explained by inclusion of the appropriate CoBXSH model used in that experiment at 50% or higher occupancy. The resulting models, along with 2Fo-Fc electron density, are shown in Figure 3. The Rwork for the final structures range from 13.0 to 15.0 % (Rfree 15.5 to 19.5 %). The X-ray data collection, processing, refinement and model building statistics are given in Table 1. Cedervall et al. Page 6 Biochemistry. Author manuscript; available in PMC 2011 September 7. NIH-PA Author Manuscript NIH-PA Author Manuscript NIH-PA Author Manuscript Analogues shorter than CoBSH; CoB5SH and CoB6SH CoB5SH is two methylene groups shorter than CoBSH, the MCR substrate. The MCRCoB5SH structure is to 1.35 Å resolution. As expected the pentanoyl chain follows the path of the CoBSH heptanoyl carbons down the substrate channel, and thus its thiol is positioned in the same place as the second carbon preceding the CoBSH thiol (Figure 3A and 4). There are no published MCR kinetic studies using CoB5SH, but as it binds in the substrate channel, it is likely to be an inhibitor. CoB6SH is one methylene shorter than CoBSH, and is a slow substrate of MCR. In this case the 1.4 Å resolution electron density of MCRCoB6SH indicates that the analogue unexpectedly binds in the substrate channel such that its thiol is virtually in the same position as the thiol of CoBSH (Figure 3B and 4). The hexanoyl chain is oriented so that it takes a shorter route down the substrate channel between carbons 2 and 5 (with the carbonyl carbon labeled as carbon 1) than CoBSH (Figure 4, Supporting Information, Figure S4). This short-cut is not seen in any of the other CoBXSH complex crystal structures, but presumably arises because this CoB6SH binding conformer is energetically more favorable, although it is not clear from the structure why this might be the case. CoB6SH binds very tightly to MCR, with an apparent Ki value of 0.1 μM (3). Water structure in the absence of HSCoM The electron density for the MCRCoB6SH crystal structure only supported the modeling of 50 % bound HSCoM. In the fraction of MCR molecules where HSCoM is absent, the HSCoM binding site is occupied by a network of four water molecules (Supporting Information, Figure S5). Two waters are positioned close to the absent sulfonate oxygen positions of HSCoM. Based on the presence of positive difference electron density, a third water was modeled ligated to the Ni, and refined to a distance of ~2 Å (2.0 Å and 2.1 Å in the two active sites of the ASU) with no distance restraint imposed between the Ni and water. This water is in a similar position as the Ni coordinating sulfonate oxygen of the heterodisulfide product in MCRsilent (Supporting Information, Figure S6 and PDB codes 3m32, 1hbm) (5, 33). The fourth water was in the vicinity of the expected position of a bridging water (W1) seen in other structures (Figure 1, 3A and 3C). Water structure in the absence of CoBSH The 1.45 Å resolution electron density obtained for MCRHSCoM indicates that the substrate channel contains only 30 % CoBSH. Nine ordered waters (W1–W9), along with an acetate ion from the crystallization solution occupy the channel, with the acetate positioned where the phosphothreonine linkage of CoBSH would be (Figure 3C). Presumably 1–2 further waters would replace the acetate under physiological conditions. Other than W3 and W7, the waters form hydrogen bonds with protein (Figure 5). One water (W2) occupies the same site as the CoBSH thiol. Presumably due to the loss of favorable interactions that exist when CoBSH is present, the hydrophobic side-chain of Valα482 adopts a second conformation modeled at 60 % occupancy (Supporting Information, Figure S7). Position of the “bridging” water, W1 The equivalent of W1 has previously been observed in MCRred1-silent and MCRox1-silent crystal structures where, in the presence of CoBSH and HSCoM, it is sited equidistant (3.2 Å) between the two coenzyme thiols (PDB codes: 1hbn, 1hbo, 1hbu) and thus been termed the “bridging water” (Figure 1) (5, 33). However, in the MCRHSCoM structure, due to the presence of the more polarized W2 water, W1 is displaced away from HSCoM to maximize the hydrogen bond interaction with W2 (2.9 Å to W2; 3.5 Å to HSCoM thiol, Figure 5). In the MCRCoB5SH structure that also contained W2, the electron density indicated that this Cedervall et al. Page 7 Biochemistry. Author manuscript; available in PMC 2011 September 7. NIH-PA Author Manuscript NIH-PA Author Manuscript NIH-PA Author Manuscript repositioning of W1 towards W2 also occurred. In contrast, the MCRCoB6SH structure contained 100% thiol at the CoBSH position, but a partial occupancy of HSCoM (50%). In this case the electron density for W1 indicated it had moved towards the nickel to form an optimal hydrogen bond with a Ni-ligating water that was only present in the absence of HSCoM (3.7 Å to CoB(6)SH thiol; 3.0 Å to Ni-ligating water, Supporting Information, Figure S5). In all structures reported here, W1 (if present) appears to be a sensitive indicator of the relative electronegativity of the Ni-ligated atom to that occupying the position of the CoBSH thiol, and was a useful check in the crystallographic modeling and refinement process. Flexibility in the substrate channel: Alternative protein conformers The binding site of HSCoM (and presumably methyl-SCoM) is more deeply buried within the enzyme, and so it must enter prior to CoBSH for productive chemistry to occur. As binding of CoBSH in the absence of co-substrate would be inhibitory, it was suggested that a conformational change upon methyl-SCoM binding might lower the Kd for CoBSH, and thus promote an ordered mechanism. Compared to the 1hbn MCRox1-silent and 1hbu MCRred1-silent crystal structures, which both have full occupancy HSCoM, the lower occupancy of HSCoM in the 1hbo MCRred1-silent structure was associated with significantly greater flexibility within the channel, and the ability to model a second conformation of a Gly-rich amino acid stretch that formed part of the CoBSH channel. This suggested that methyl-SCoM binding might cause the channel to become more ordered, increasing the affinity of MCR for CoBSH by conformational restriction rather than a switch mechanism where the structure reorganizes from one well-defined conformer to another (33). In the MCRHSCoM data containing 30 % CoBSH and 100% HSCoM, the Fo-Fc difference electron density map at one of the two independent active sites in the ASU contained positive peaks that suggested the presence of an alternate conformation also involving this part of the polypeptide (Supporting Information, Figure S8). Using this as a guide, a similar second conformation involving seven contiguous amino acid residues of the same Gly-rich amino acid stretch (β366–372) could be modeled and refined at 20 % occupancy leaving no residual difference density. Parts of the α′ subunit (α′111–129 and α′237–242) that are in close proximity to this stretch of amino acids also exhibit second conformations, with the main-chain carbonyl of α′243 in van der Waals contact with the B ring of F430 tetrapyrrole (Supporting Information, Figure S9). Modeling these at 20% occupancy accounted for the weak positive Fo-Fc difference electron density peaks observed in these areas. The evidence of alternate conformers in these areas lends support to the proposal that increased flexibility in the substrate channel propagates through the protein (33). The MCRCoB6SH crystal structure contains 50 % CoB6SH, 50% CoBSH and 50% HSCoM. In this case there is no evidence of an alternate loop conformation in either active site of the ASU. However, as CoBSH and CoB6SH combined are at 100% occupancy, it is not surprising their favorable interactions with the substrate channel would reduce conformational disorder, despite the partial occupancy of HSCoM. Analogues longer than CoBSH; CoB8SH and CoB9SH Both analogues could be accommodated in the MCR substrate channel (Figures 3D and 3E). The electron density supported final models containing 50% CoB8SH for MCRCoB8SH (1.8 Å resolution) and 60 % CoB9SH for MCRCoB9SH (1.45 Å resolution). The phosphate head- groups are in identical positions to those of CoBSH, CoB5SH, CoB6SH (Figure 4) (5, 33). Both analogues follow the crystallographically observed chain path of bound CoBSH, with the extra atoms displacing the W1 water and placing the thiols closer to the nickel (Figure 6). CoB9SH does have a second conformer that deviates from the CoBSH path, but the thiol position for this conformer and the CoBSH-tracking conformer are identical (Figure 3E and Cedervall et al. Page 8 Biochemistry. Author manuscript; available in PMC 2011 September 7. NIH-PA Author Manuscript NIH-PA Author Manuscript NIH-PA Author Manuscript Supporting Information, Figure S10). Interestingly, the thiol of CoB8SH is not coincident with the CoB9SH carbon that precedes the CoB9SH thiol (Figure 6). CoB8SH is an MCR inhibitor with an apparent Ki of 15 μM (3). CoB9SH has never been tested for inhibition of MCR-catalyzed methane formation, but it is reasonable to assume that it would be an inhibitor. CoBXSH thiol-to-nickel spatial relationship The position of CoBSH in previous crystal structures poses a conundrum (5, 33). In all the proposed catalytic mechanisms, CoBSH must interact with species generated at the nickel. Perplexingly, the sulfur of the CoBSH substrate is 8.8 Å from the Ni(II) in the MCRox1-silent and both MCRred1-silent crystal structures, and 6.4 Å from the thiol of the substrate analogue HSCoM (Figure 1). Modeling studies demonstrated that the addition of a methyl group to HSCoM did not bridge this gap (35, 45, 53). Therefore, a conformational change has been postulated that would enable CoBSH to penetrate deeper into the substrate channel, and thus approach closer to any nickelbound species. The heterodisulfide product in the MCRsilent crystal structure has the CoBSH portion in virtually the same place as in MCRox1-silent, giving no clue to possible structural changes that might occur to facilitate CoBSH reacting with nickel-associated intermediates (5, 33). Trigonometry suggests that if the alkanoyl chain of CoBSH or its analogue is in an extended conformation, each additional unit in the chain would lead to the thiol moving ~1.2 Å towards the Ni. Until this study there have been no crystal structures of CoBSH analogues in complex with MCR, so mechanistic studies using different chain length analogues of CoBSH assumed that shorter analogues would trace the observed path of CoBSH, and longer analogues would penetrate about ~1.2 Å deeper per additional chain unit into the channel. In the case of the shortest analogue CoB5SH, it does indeed follow the path of CoBSH, with the thiol of CoB5SH being 2.8 Å away from the thiol position of CoBSH. However, due to the conformation CoBSH adopts when bound in the substrate channel, the difference in the S-Ni distance is small; the CoB5SH thiol being only 0.5 Å farther from the Ni than CoBSH (8.8 Å for CoBSH vs. 9.3 Å for CoB5SH) (Table 2). This is due to the alkanoyl chain of CoBSH not being in an extended conformation from carbons 4 to 6 (carbon 1 is the carbonyl carbon). CoB6SH on the other hand adopts a conformation that places its thiol in virtually the same position as the thiol of CoBSH (Figure 4 and Table 2). This is consistent with CoB6SH being a substrate. However, the kcat is 1000-fold lower than for CoBSH (3, 35) although its Km value (180 μM, Dey & Ragsdale, in preparation) is similar to that of CoBSH (Km = 75 μM (3)). The reason for this may be that the shorter alkanoyl chain may not enable the analogue thiol to approach the nickel close enough for efficient catalysis, and thus explain why CoB6SH is such a poor substrate. In the case of the longer CoBXSH analogues, the sulfur of CoB8SH is 2.6 Å closer to the Ni ion of F430 than that of CoBSH, and 2.5 Å closer to the thiol of HSCoM (Figure 6 and Table 2). The CoB9SH molecule follows the path of CoBSH, and reaches only a little further into the substrate channel than CoB8SH, with the CoB9SH thiol positioned 2.9 Å closer to the Ni than the thiol of CoBSH (Figure 6 and Table 2). This is only 0.3 Å closer than the distance observed for the CoB8SH thiol, even though they are non-coincident. The distance to the thiol of HSCoM is 2.6 Å closer than that of the substrate, CoBSH; only 0.1 Å closer than the CoB8SH thiol. The two analogue thiols are above an annular hydrophobic aromatic environment created by Pheα330, Tyrα333, Pheα443, Pheβ361 and Tyrβ367 that lies between them and F430 (Figure 6). As a result, penetrating further into the channel may be energetically unfavorable, consistent with the small difference in relative distances between the CoB8SH/CoB9SH thiols and the HSCoM thiol/F430 nickel. The annulus is proposed to be catalytically important in positioning methyl-SCoM and stabilizing the methane product, Cedervall et al. Page 9 Biochemistry. Author manuscript; available in PMC 2011 September 7. NIH-PA Author Manuscript NIH-PA Author Manuscript NIH-PA Author Manuscript and the tyrosines have been proposed to be proton donors associated with mechanism II (Scheme 2B) (5, 33). Thus, there appear to be three preferential distances for thiols (including that of HSCoM) within the MCR substrate channel; HSCoM at 2.4 Å, CoB6SH/CoBSH at 8.7 – 8.8 Å and CoB8SH/CoB9SH at 5.9 – 6.2 Å from the nickel of F430 (Table 2). Recent ENDOR, high-field continuous and pulse EPR work has identified changes in nickel co-ordination when CoBSH is added to MCRred1c (active Ni(I)-MCRred1 + HSCoM) (14, 15, 18, 31). This generates up to 50% MCRred2 which is comprised of two distinct nickel co- ordination geometries; an axial MCRred2a formally assigned as a Ni(III)-hydride, and a rhombic MCRred2r in which the thiol of HSCoM is a Ni(I) ligand (Supporting Information, Scheme S1). Formation of MCRred2 could also be induced by addition of CoBSH substrate analogues CoBS-CH3 and CoBS-CF3, which have a chain length one unit longer than substrate CoBSH (18, 53). The CoBS-CF3 enabled 19F-ENDOR studies to be performed, and demonstrated that following CoBS-CF3 addition the remaining MCRred1 species had Ni(I)-19F distances of 6.2 – 7.7 Å. This distance range agreed with a CoBS-CF3 model created using the CoBSH position observed in the MCRox1-silent crystal structure (53). However, in the MCRred2 species the Ni(I)-19F distances had shortened indicating a movement of ~ 2 Å towards the nickel (MCRred2a Ni(I)-19 F, 4.0 – 5.5 Å; MCRred2r Ni(I)-19F, 4.5 – 5.7 Å). In the case of the CoB8SH analogue (the closest equivalent to CoBS- CF3 used in this study) the thiol to Ni(II) distance lies between the distance ranges observed in the CoBS-CF3 studies, and so the fluorine(s) of the CoBS-CF3 in MCRred2 might penetrate a little further into the hydrophobic annulus in the MCRred2 species. As the alkanoyl chain of CoBSH is not fully extended it could easily undergo a similar conformation change to that observed in the MCRred2 state. Reaction of MCR Ni(III)-alkyl species’ with CoB8SH and CoB9SH The two longer CoBXSH analogues have been shown to undergo alkylation when reacted with MCRPS, a [Ni(III)-alkyl ↔ Ni(II)-alkylsulfonate radical] formed from reaction of Ni(I)-MCRred1 with bromopropanesulfonate (BPS) (Supporting Information, Scheme S1) (20, 23, 30, 45, 54). BPS is a substrate of MCRred1 in a reaction that involves a rapid CoBSH-independent nucleophilic attack by Ni(I) on BPS to displace bromide and generate MCRPS at a rate ~60-fold faster than generation of methane from CoBSH and methyl- HSCoM (20, 45). Certain thiols can eliminate the propylsulfonate to yield a thioether product and regenerate MCRred1, although at a rate 1000-fold slower than methane formation (45). Both CoB8SH and CoB9SH can react with MCRPS to regenerate MCRred1, but CoBSH cannot. The overall second-order rate constant for the reactivation of MCR by CoB8SH is 160 M−1s−1, whereas for CoB9SH the reaction is more sluggish (12 M−1s−1). CoB9SH might be expected to be closer to the proximal Ni ligand. It was therefore proposed that this caused steric interference and explained why CoB9SH was a poorer reactivator of MCR than CoB8SH. Our study has shown that the thiols of these two analogues are placed such that they are approximately the same distance (~3.7 Å) from the thiol of HSCoM ligated to the Ni atom (Table 2). The Ni(II)-HSCoM bond is 2.4 Å, whereas an Ni(III)-alkyl bond is expected to be ~2 Å (24, 33), thus indicating that a conformational change is required to effect the nucleophilic attack of the CoB8SH and CoB9SH thiols on an alkyl- bound species. It would thus appear that a conformational change, such as observed in MCRred2, is required for this chemistry also (53). A Ni(III)-alkyl species is akin to the first intermediate in mechanism I of MCR-catalyzed methane formation, Ni(III)-methyl (MCRMe, Supporting Information, Scheme S1, Scheme 2A) (11, 27). MCRMe has been shown to be capable of generating MCRred1 and methyl- SCoM upon addition of HSCoM (which is the reverse of mechanism I, step 1, Scheme 2A); Cedervall et al. Page 10 Biochemistry. Author manuscript; available in PMC 2011 September 7. NIH-PA Author Manuscript NIH-PA Author Manuscript NIH-PA Author Manuscript similar chemistry to the observed formation of a thioether product from the Ni(III)-alkyl. Further addition of CoBSH following HSCoM treatment of MCRMe led to methane and heterodisulfide formation, the natural products of methanogenesis. Although this lends credence to mechanism I, it should be noted that like MCRPS, MCRMe in these experiments was generated artificially. However, the MCRCoB9SH crystal structure demonstrates that the two additional methylene units in the alkanoyl chain c.f. CoBSH, do not necessarily translate into direct interaction of the thiol with the nickel proximal ligand. However, this could represent the favorable position for a CoBSH thiol interacting with the methyl group of methyl-SCoM. Just as the alkanoyl chain of CoB6SH has a more extended conformation than CoBSH in the substrate channel, CoBSH could also adopt a more extended conformation so that its thiol was in a similar position as the thiol of CoB8SH, priming it for reaction with a nickel bound species. If a significant conformational change is required early in MCR-catalyzed chemistry, which would be a requirement of mechanism I, catalysis may well involve a rearrangement of the aromatic amino acid annulus due to the presence of the methyl of methyl-SCoM, and this might enable deeper penetration of CoBSH into MCR (Figure 6). All the crystal structures in this study, and those solved previously, are of the inactive Ni(II)-MCR, which disfavors close approach to the nickel in the absence of Ni(I)-bound methyl-SCoM, even in the case of CoB9SH. Conclusion The goal of this study was to induce structural changes within the substrate channel and active site of MCR using analogues of coenzyme CoBSH. It was hoped that this would shed light on the nature of conformational changes that have been proposed to occur in MCR catalysis. We have shown that that the CoBXSH analogues do not lead to any significant conformational changes within the context of inactive Ni(II)-MCR. Therefore, it may be that methyl-SCoM is the key coenzyme, in combination with a nickel oxidation state of 1 (and 3), that triggers a conformational change bringing the thiol of CoBSH closer to the nickel. Thus, the crystal structure of the Ni(I)-methyl-SCoM/MCR complex may be required to structurally define the conformational changes required for MCR-mediated chemistry. Supplementary Material Refer to Web version on PubMed Central for supplementary material. Acknowledgments X-ray data were collected at beam-line 14-BM-C and photoreduction studies at 14-ID-B, BioCARS at the Advanced Photon Source (APS), Argonne National Laboratory, Argonne, IL. We thank Vukica Srajer and Yu- Sheng Chen for valuable assistance during data collection. Use of the Advanced Photon Source was supported by the U.S. Department of Energy, Basic Energy Sciences, Office of Science under Contract No. DE- AC02-06CH11357. Use of the BioCARS Sector 14 was supported by the National Institutes of Health, National Center for Research Resources, under grant number RR007707. Computer resources were provided by the Basic Sciences Computing Laboratory of the University of Minnesota Supercomputing Institute, and we thank Can Ergenekan for his support. We also thank Ed Hoeffner at the Kahlert Structural Biology Laboratory (KSBL) at the University of Minnesota. Use of the KSBL was supported by a Minnesota Partnership for Biotechnology and a Medical Genomics Grant SPAP-05-0013-P-FY06. References 1. Thauer RK. Biochemistry of methanogenesis: a tribute to Marjory Stephenson. Microbiology. 1998; 144:2377–2406. [PubMed: 9782487] 2. Thauer RK, Shima S. Methane as fuel for anaerobic microorganisms. Ann N Y Acad Sci. 2008; 1125:158–170. [PubMed: 18096853] Cedervall et al. Page 11 Biochemistry. Author manuscript; available in PMC 2011 September 7. NIH-PA Author Manuscript NIH-PA Author Manuscript NIH-PA Author Manuscript 3. Ellermann J, Hedderich R, Bocher R, Thauer RK. The final step in methane formation. Investigations with highly purified methyl-CoM reductase (component C) from Methanobacterium thermoautotrophicum (strain Marburg). Eur J Biochem. 1988; 172:669–677. [PubMed: 3350018] 4. Ellefson WL, Wolfe RS. Component C of the methylreductase system of Methanobacterium. J Biol Chem. 1981; 256:4259–4262. [PubMed: 6783657] 5. Ermler U, Grabarse W, Shima S, Goubeaud M, Thauer RK. Crystal structure of methyl-coenzyme M reductase: the key enzyme of biological methane formation. Science. 1997; 278:1457–1462. [PubMed: 9367957] 6. Diekert G, Gilles HH, Jaenchen R, Thauer RK. Incorporation of 8 succinate per mol nickel into factors F430 by Methanobacterium thermoautotrophicum. Arch Microbiol. 1980; 128:256–262. [PubMed: 7212929] 7. Diekert G, Jaenchen R, Thauer RK. Biosynthetic evidence for a nickel tetrapyrrole structure of factor F430 from Methanobacterium thermoautotrophicum. FEBS Letters. 1980; 119:118–120. [PubMed: 7428919] 8. Whitman WB, Wolfe RS. Presence of nickel in Factor F430 from Methanobacterium bryantii. Biochem Biophys Res Comm. 1980; 92:1196–1201. [PubMed: 7370029] 9. Albracht SPJ, Ankel-Fuchs D, Böcher R, Ellermann J, Moll J, van der Zwann JW, Thauer RK. Five new EPR signals assigned to nickel in methyl-coenzyme M reductase from Methanobacterium thermoautotrophicum, strain Marburg. Biochim Biophys Acta. 1988; 955:86–102. 10. Dey M, Kunz RC, Lyons DM, Ragsdale SW. Characterization of alkyl-nickel adducts generated by reaction of methyl-coenzyme m reductase with brominated acids. Biochemistry. 2007; 46:11969– 11978. [PubMed: 17902704] 11. Dey M, Telser J, Kunz RC, Lees NS, Ragsdale SW, Hoffman BM. Biochemical and spectroscopic studies of the electronic structure and reactivity of a methyl-Ni species formed on methyl- coenzyme M reductase. J Am Chem Soc. 2007; 129:11030–11032. [PubMed: 17711283] 12. Duin EC, Cosper NJ, Mahlert F, Thauer RK, Scott RA. Coordination and geometry of the nickel atom in active methyl-coenzyme M reductase from Methanothermobacter marburgensis as detected by X-ray absorption spectroscopy. J Biol Inorg Chem. 2003; 8:141–148. [PubMed: 12459909] 13. Duin EC, Signor L, Piskorski R, Mahlert F, Clay MD, Goenrich M, Thauer RK, Jaun B, Johnson MK. Spectroscopic investigation of the nickel-containing porphinoid cofactor F(430). Comparison of the free cofactor in the (+)1, (+)2 and (+)3 oxidation states with the cofactor bound to methyl- coenzyme M reductase in the silent, red and ox forms. J Biol Inorg Chem. 2004; 9:563–576. [PubMed: 15160314] 14. Finazzo C, Harmer J, Bauer C, Jaun B, Duin EC, Mahlert F, Goenrich M, Thauer RK, Van Doorslaer S, Schweiger A. Coenzyme B induced coordination of coenzyme M via its thiol group to Ni(I) of F430 in active methyl-coenzyme M reductase. J Am Chem Soc. 2003; 125:4988–4989. [PubMed: 12708843] 15. Finazzo C, Harmer J, Jaun B, Duin EC, Mahlert F, Thauer RK, Van Doorslaer S, Schweiger A. Characterization of the MCRred2 form of methyl-coenzyme M reductase: a pulse EPR and ENDOR study. J Biol Inorg Chem. 2003; 8:586–593. [PubMed: 12624730] 16. Goubeaud M, Schreiner G, Thauer RK. Purified methyl-coenzyme-M reductase is activated when the enzyme-bound coenzyme F430 is reduced to the nickel(I) oxidation state by titanium(III) citrate. Eur J Biochem. 1997; 243:110–114. [PubMed: 9030728] 17. Harmer J, Finazzo C, Piskorski R, Bauer C, Jaun B, Duin EC, Goenrich M, Thauer RK, Van Doorslaer S, Schweiger A. Spin density and coenzyme M coordination geometry of the ox1 form of methyl-coenzyme M reductase: a pulse EPR study. J Am Chem Soc. 2005; 127:17744–17755. [PubMed: 16351103] 18. Harmer J, Finazzo C, Piskorski R, Ebner S, Duin EC, Goenrich M, Thauer RK, Reiher M, Schweiger A, Hinderberger D, Jaun B. A nickel hydride complex in the active site of methyl- coenzyme m reductase: implications for the catalytic cycle. J Am Chem Soc. 2008; 130:10907– 10920. [PubMed: 18652465] Cedervall et al. Page 12 Biochemistry. Author manuscript; available in PMC 2011 September 7. NIH-PA Author Manuscript NIH-PA Author Manuscript NIH-PA Author Manuscript 19. Hinderberger D, Ebner S, Mayr S, Jaun B, Reiher M, Goenrich M, Thauer RK, Harmer J. Coordination and binding geometry of methyl-coenzyme M in the red1m state of methyl- coenzyme M reductase. J Biol Inorg Chem. 2008; 13:1275–1289. [PubMed: 18712421] 20. Kunz RC, Horng YC, Ragsdale SW. Spectroscopic and kinetic studies of the reaction of bromopropanesulfonate with methyl-coenzyme M reductase. J Biol Chem. 2006; 281:34663– 34676. [PubMed: 16966321] 21. Mahlert F, Bauer C, Jaun B, Thauer RK, Duin EC. The nickel enzyme methyl-coenzyme M reductase from methanogenic archaea: In vitro induction of the nickel-based MCR-ox EPR signals from MCR-red2. J Biol Inorg Chem. 2002; 7:500–513. [PubMed: 11941508] 22. Mahlert F, Grabarse W, Kahnt J, Thauer RK, Duin EC. The nickel enzyme methyl-coenzyme M reductase from methanogenic archaea: in vitro interconversions among the EPR detectable MCR- red1 and MCR-red2 states. J Biol Inorg Chem. 2002; 7:101–112. [PubMed: 11862546] 23. Rospert S, Voges M, Berkessel A, Albracht SP, Thauer RK. Substrate-analogue-induced changes in the nickel-EPR spectrum of active methyl-coenzyme-M reductase from Methanobacterium thermoautotrophicum. Eur J Biochem. 1992; 210:101–107. [PubMed: 1332856] 24. Sarangi R, Dey M, Ragsdale SW. Geometric and electronic structures of the Ni(I) and methyl- Ni(III) intermediates of methyl-coenzyme M reductase. Biochemistry. 2009; 48:3146–3156. [PubMed: 19243132] 25. Tang Q, Carrington PE, Horng YC, Maroney MJ, Ragsdale SW, Bocian DF. X-ray absorption and resonance Raman studies of methyl-coenzyme M reductase indicating that ligand exchange and macrocycle reduction accompany reductive activation. J Am Chem Soc. 2002; 124:13242–13256. [PubMed: 12405853] 26. Telser J, Davydov R, Horng YC, Ragsdale SW, Hoffman BM. Cryoreduction of methyl-coenzyme M reductase: EPR characterization of forms, MCR(ox1) and MCR (red1). J Am Chem Soc. 2001; 123:5853–5860. [PubMed: 11414817] 27. Yang N, Reiher M, Wang M, Harmer J, Duin EC. Formation of a nickel-methyl species in methyl- coenzyme M reductase, an enzyme catalyzing methane formation. J Am Chem Soc. 2007; 129:11028–11029. [PubMed: 17711279] 28. Albracht SPJ, Ankelfuchs D, Vanderzwaan JW, Fontijn RD, Thauer RK. A New Electron- Paramagnetic-Res Signal of Nickel in Methanobacterium-Thermoautotrophicum. Biochim Biophys Acta. 1986; 870:50–57. 29. Telser J, Horng YC, Becker DF, Hoffman BM, Ragsdale SW. On the assignment of nickel oxidation states of the Ox1, Ox2 forms of methyl-coenzyme M reductase. J Am Chem Soc. 2000; 122:182–183. 30. Hinderberger D, Piskorski RR, Goenrich M, Thauer RK, Schweiger A, Harmer J, Jaun B. A nickel- alkyl bond in an inactivated state of the enzyme catalyzing methane formation. Angewandte Chemie-International Ed. 2006; 45:3602–3607. 31. Kern DI, Goenrich M, Jaun B, Thauer RK, Harmer J, Hinderberger D. Two sub-states of the red2 state of methyl-coenzyme M reductase revealed by high-field EPR spectroscopy. J Biol Inorg Chem. 2007; 12:1097–1105. [PubMed: 17690920] 32. Becker DF, Ragsdale SW. Activation of methyl-SCoM reductase to high specific activity after treatment of whole cells with sodium sulfide. Biochemistry. 1998; 37:2639–2647. [PubMed: 9485414] 33. Grabarse W, Mahlert F, Duin EC, Goubeaud M, Shima S, Thauer RK, Lamzin V, Ermler U. On the mechanism of biological methane formation: structural evidence for conformational changes in methyl-coenzyme M reductase upon substrate binding. J Mol Biol. 2001; 309:315–330. [PubMed: 11491299] 34. Grabarse W, Mahlert F, Shima S, Thauer RK, Ermler U. Comparison of three methyl-coenzyme M reductases from phylogenetically distant organisms: unusual amino acid modification, conservation and adaptation. J Mol Biol. 2000; 303:329–344. [PubMed: 11023796] 35. Horng YC, Becker DF, Ragsdale SW. Mechanistic studies of methane biogenesis by methyl- coenzyme M reductase: evidence that coenzyme B participates in cleaving the C-S bond of methyl-coenzyme M. Biochemistry. 2001; 40:12875–12885. [PubMed: 11669624] Cedervall et al. Page 13 Biochemistry. Author manuscript; available in PMC 2011 September 7. NIH-PA Author Manuscript NIH-PA Author Manuscript NIH-PA Author Manuscript 36. Berkessel A. Methyl-Coenzyme-M Reductase - Model Studies on Pentadentate Nickel-Complexes and a Hypothetical Mechanism. Bioorg Chem. 1991; 19:101–115. 37. Jaun B. Coenzyme-F430 from Methanogenic Bacteria - Oxidation of F430 Pentamethyl Ester to the Ni(Iii) Form. Helvetica Chimica Acta. 1990; 73:2209–2217. 38. Signor L, Knuppe C, Hug R, Schweizer B, Pfaltz A, Jaun B. Methane formation by reaction of a methyl thioether with a photo-excited nickel thiolate - A process mimicking methanogenesis in archaea. Chemistry-a European Journal. 2000; 6:3508–3516. 39. Chen SL, Pelmenschikov V, Blomberg MR, Siegbahn PE. Is there a Ni-methyl intermediate in the mechanism of methyl-coenzyme M reductase? J Am Chem Soc. 2009; 131:9912–9913. [PubMed: 19569621] 40. Pelmenschikov V, Blomberg MRA, Siegbahn PEM, Crabtree RH. A mechanism from quantum chemical studies for methane formation in methanogenesis. J Am Chem Soc. 2002; 124:4039– 4049. [PubMed: 11942842] 41. Pelmenschikov V, Siegbahn PE. Catalysis by methyl-coenzyme M reductase: a theoretical study for heterodisulfide product formation. J Biol Inorg Chem. 2003; 8:653–662. [PubMed: 12728361] 42. Duin EC, McKee ML. A new mechanism for methane production from methyl-coenzyme M reductase as derived from density functional calculations. J Phys Chem. 2008; B 112:2466–2482. 43. Bobik TA, Wolfe RS. Physiological importance of the heterodisulfide of coenzyme M and 7- mercaptoheptanoylthreonine phosphate in the reduction of carbon dioxide to methane in Methanobacterium. Proc Natl Acad Sci U S A. 1988; 85:60–63. [PubMed: 3124103] 44. Goenrich M, Duin EC, Mahlert F, Thauer RK. Temperature dependence of methyl-coenzyme M reductase activity and of the formation of the methyl-coenzyme M reductase red2 state induced by coenzyme B. J Biol Inorg Chem. 2005; 10:333–342. [PubMed: 15846525] 45. Kunz RC, Dey M, Ragsdale SW. Characterization of the Thioether Product Formed from the Thiolytic Cleavage of the Alkyl-Nickel Bond in Methyl-Coenzyme M Reductase. Biochemistry. 2008; 47:2661–2667. [PubMed: 18220418] 46. Noll KM, Donnelly MI, Wolfe RS. Synthesis of 7-mercaptoheptanoylthreonine phosphate and its activity in the methylcoenzyme M methylreductase system. J Biol Chem. 1987; 262:513–515. [PubMed: 3100513] 47. Olson KD, McMahon CW, Wolfe RS. Photoactivation of the 2-(methylthio)ethanesulfonic acid reductase from Methanobacterium. Proc Natl Acad Sci U S A. 1991; 88:4099–4103. [PubMed: 1903534] 48. Zehnder AJ, Wuhrmann K. Titanium (III) citrate as a nontoxic oxidation-reduction buffering system for the culture of obligate anaerobes. Science. 1976; 194:1165–1166. [PubMed: 793008] 49. Gunsalus RP, Romesser JA, Wolfe RS. Preparation of coenzyme M analogues and their activity in the methyl coenzyme M reductase system of Methanobacterium thermoautotrophicum. Biochemistry. 1978; 17:2374–2377. [PubMed: 98178] 50. Otwinowski Z, Minor W. Processing of X-ray Diffraction Data Collected in Oscillation Mode. Methods in Enzymology: Macromolecular Crystallography, part A. 1997; 276:307–326. 51. The CCP4 suite: programs for protein crystallography. Acta Crystallogr D Biol Crystallogr. 1994; 50:760–763. [PubMed: 15299374] 52. Emsley P, Cowtan K. Coot: model-building tools for molecular graphics. Acta Crystallogr D Biol Crystallogr. 2004; 60:2126–2132. [PubMed: 15572765] 53. Ebner S, Jaun B, Goenrich M, Thauer RK, Harmer J. Binding of coenzyme B induces a major conformational change in the active site of methyl-coenzyme M reductase. J Am Chem Soc. 2010; 132:567–575. [PubMed: 20014831] 54. Goenrich M, Mahlert F, Duin EC, Bauer C, Jaun B, Thauer RK. Probing the reactivity of Ni in the active site of methyl-coenzyme M reductase with substrate analogues. J Biol Inorg Chem. 2004; 9:691–705. [PubMed: 15365904] Cedervall et al. Page 14 Biochemistry. Author manuscript; available in PMC 2011 September 7. NIH-PA Author Manuscript NIH-PA Author Manuscript NIH-PA Author Manuscript Figure 1. The active site and substrate channel in the MCRox1-silent crystal structure (PDB code 1hbn) (9). Coenzyme F430, CoBSH and HSCoM are drawn as stick colored by atom (carbon: dark grey). The nickel is displayed as a green sphere, water as a red sphere. Interactions are drawn as dashed lines, and the corresponding distance is indicted in Angstroms (Å). The path of the substrate channel was defined in the absence of F430, CoBSH, HSCoM and water, with the surface closest to the viewer cut away. The figure was generated using PyMOL (http://www.pymol.org). Cedervall et al. Page 15 Biochemistry. Author manuscript; available in PMC 2011 September 7. NIH-PA Author Manuscript NIH-PA Author Manuscript NIH-PA Author Manuscript Figure 2. Drawing of CoBSH analogues; (A) N-5-mercaptopentanoylthreonine phosphate (CoB5SH); (B) N-6-mercaptohexanoylthreonine phosphate (CoB6SH); (C) N-8- mercaptooctanoylthreonine phosphate (CoB8SH); (D) N-9-mercaptononanoylthreonine phosphate (CoB9SH). Cedervall et al. Page 16 Biochemistry. Author manuscript; available in PMC 2011 September 7. NIH-PA Author Manuscript NIH-PA Author Manuscript NIH-PA Author Manuscript Figure 3. The active sites and substrate channels of the MCR crystal structures.; (A) MCRCoB5SH; (B) MCRCoB6SH; (C) MCRHSCoM; (D) MCRCoB8SH; (E) MCRCoB9SH. 2Fo-Fc electron density map around the CoBXSH analogues, waters in the CoBSH binding part of the channel and the acetate ion (contoured at 1σ) is shown as a blue mesh. The protein is drawn as cartoon. CoBXSH and acetate are drawn as stick and colored by atom (carbon: CoB5SH orange; CoB6SH pale yellow; CoB8SH light blue; CoB9SH magenta; acetate white). Coenzyme F430 and HSCoM are drawn as stick colored by atom (carbon: F430 dark grey; HSCoM medium grey). The nickel is displayed as a green sphere, and waters as red spheres. The figure was generated using PyMOL (http://www.pymol.org/). Cedervall et al. Page 17 Biochemistry. Author manuscript; available in PMC 2011 September 7. NIH-PA Author Manuscript NIH-PA Author Manuscript NIH-PA Author Manuscript Figure 4. Overlay of CoBSH (from PDB code 1hbn) and the different CoBSH analogues. CoBXSH are drawn as stick with the thiol represented by a sphere, and colored CoB5SH orange; CoB6SH pale yellow; CoBSH light green; CoB8SH light blue; CoB9SH magenta. The protein is drawn as cartoon. Coenzyme F430 and HSCoM are drawn as stick colored by atom (carbon: F430 dark grey; HSCoM medium grey). The nickel is displayed as a green sphere. The figure was generated using PyMOL (http://www.pymol.org/). Cedervall et al. Page 18 Biochemistry. Author manuscript; available in PMC 2011 September 7. NIH-PA Author Manuscript NIH-PA Author Manuscript NIH-PA Author Manuscript Figure 5. Hydrogen bonding diagram for the water structure modeled in MCRHSCoM. The water molecules are named as in Figure 3C (W1–W9); WA, WB and WC are water molecules that are present in all structures (i.e. in concert with the substrate CoBSH and the CoBSH analogues). Interactions between surrounding residues and the water molecules are drawn as dashed lines, and the corresponding distance is indicated in Angstroms (Å). Cedervall et al. Page 19 Biochemistry. Author manuscript; available in PMC 2011 September 7. NIH-PA Author Manuscript NIH-PA Author Manuscript NIH-PA Author Manuscript Figure 6. Stereo image of the annulus of aromatic amino acids distal of coenzyme F430. The protein is drawn as cartoon with the side-chains of the aromatic residues drawn as white stick. CoBSH, (from PDB code 1hbn (9)), CoB8SH and CoB9SH are drawn as stick with the thiols represented by spheres, and colored CoBSH light green; CoB8SH light blue; CoB9SH magenta. Coenzyme F430 and HSCoM are drawn as stick colored by atom (carbon: F430 dark grey; HSCoM medium grey). The nickel is displayed as a green sphere. The figure was generated using PyMOL (http://www.pymol.org/). Cedervall et al. Page 20 Biochemistry. Author manuscript; available in PMC 2011 September 7. NIH-PA Author Manuscript NIH-PA Author Manuscript NIH-PA Author Manuscript Scheme 1. Reaction catalyzed by methyl-coenzyme M reductase Cedervall et al. Page 21 Biochemistry. Author manuscript; available in PMC 2011 September 7. NIH-PA Author Manuscript NIH-PA Author Manuscript NIH-PA Author Manuscript Scheme 2. Two of the proposed catalytic mechanisms for methyl-coenzyme M reductase; (A) mechanism I; (B) mechanism II. Cedervall et al. Page 22 Biochemistry. Author manuscript; available in PMC 2011 September 7. NIH-PA Author Manuscript NIH-PA Author Manuscript NIH-PA Author Manuscript NIH-PA Author Manuscript NIH-PA Author Manuscript NIH-PA Author Manuscript Cedervall et al. Page 23 Table 1 X-ray Data Collection, Processing and Refinement Statistics Data collection and processing statistics Name of data set MCRCoB5SH MCRCoB6SH MCRHSCoM MCRCoB8SH MCRCoB9SH Measured reflections 1969388 2427498 1440665 1160543 1425506 Unique reflections 553755 446253 405349 211803 401701 Resolution (Å) a 50.0–1.30 (1.35–1.30) 50.0–1.40 (1.45–1.40) 50.0–1.45 (1.50–1.45) 50.0–1.80 (1.86–1.80) 50.0–1.45 (1.50–1.45) Completeness (%) a 97.1 (78.1) 99.9 (100.0) 99.5 (99.7) 99.8 (100.0) 98.1 (95.4) R-sym (%) a,b 5.5 (32.9) 7.3 (44.7) 6.2 (44.0) 8.4 (47.7) 5.6 (42.5) I/σI a 22.3 (3.6) 20.4 (4.0) 20.2 (3.2) 21.8 (3.9) 24.3 (3.2) Space group P21 P21 P21 P21 P21 Refinement and model building statistics Resolution (Å) a 20.49–1.30 (1.33–1.30) 19.89–1.40 (1.44–1.40) 20.15–1.45 (1.49–1.45) 19.93–1.80 (1.84–1.80) 20.07–1.45 (1.48–1.45) No. of reflection in working set a 525817 (30239) 423854 (25833) 384868 (25791) 201128 (11193) 381474 (23611) No. of reflection in test set a 27777 (1576) 22348 (1331) 20362 (1319) 10625 (557) 20163 (1210) R-work (%) c 14.32 13.04 13.47 14.95 13.58 R-free (%) d 16.56 15.53 16.22 19.54 16.44 ESU (Å) R-work/R-free 0.044/0.046 0.049/0.051 0.056/0.059 0.121/0.119 0.057/0.060 No. protein atoms 20087 19960 20265 19750 20036 No. coenzyme atoms 218 220 180 224 272 No. ligand atoms 37 62 52 26 49 No. water molecules 2443 2352 2516 1893 2432 RMS bond lengths (Å) 0.033 0.033 0.032 0.028 0.032 bond angles (deg.) 2.693 2.625 2.468 2.059 2.549 Ramachandran plot (%) favored 97.8 97.5 97.6 97.2 97.7 allowed 2.1 2.4 2.3 2.7 2.1 Biochemistry. Author manuscript; available in PMC 2011 September 7. NIH-PA Author Manuscript NIH-PA Author Manuscript NIH-PA Author Manuscript Cedervall et al. Page 24 disallowed 0.1 0.1 0.1 0.1 0.1 Average B-factor (Å2) protein 12.42 13.35 12.12 17.22 12.73 coenzymes 8.20 9.24 7.25 11.24 8.27 ligands 31.95 35.48 28.29 33.76 32.92 waters 22.95 24.89 23.85 26.79 24.09 over all 13.54 14.57 13.40 18.02 13.93 Occupancy of HSCoM per active site (%)e 90/90 50/50 100/100 90/90 90/85 Occupancy of CoBSH per active site (%) e 50/50 50/50 30/30 50/50 40/40 CoBSH analogue, occupancy per active site (%) e CoB5SH, 50/50 CoB6SH, 50/50 CoB8SH, 50/50 CoB9SH, 60/60 Other molecule, occupancy per active site (%) e Acetate, 70/70 aValues in brackets correspond to the highest resolution shell. bR-sym = ΣhklΣNj=1|Ihkl−Ihkl(j)|/ΣhklN*Ihkl, sum over all reflections and all observations N, with Ihkl(j) intensity of the jth observation of reflection hkl and Ihkl mean intensity of the reflection hkl. cR-work = Σ||Fo| − |Fc||/Σ|Fo|, where |Fo| = observed structure factor amplitude and |Fc| = calculated structure factor amplitude. dR-free, R-factor based on 5% of the data excluded from refinement. eOccupancy of model in each of the two crystallographically independent active sites in the ASU Biochemistry. Author manuscript; available in PMC 2011 September 7. NIH-PA Author Manuscript NIH-PA Author Manuscript NIH-PA Author Manuscript Cedervall et al. Page 25 Table 2 Distances from analogue thiols. CoBXS - SCoM distance (Å) CoBXS - Ni distance (Å) CoB5SH 7.11/7.11a 9.30/9.30 CoB6SH 6.26/6.26 8.70/8.70 CoB7SH (substrate) b 6.37/6.39 8.73/8.77 CoB8SH 3.75/3.78 6.16/6.17 CoB9SH 3.71/3.68 5.96/5.91 aDistances in the two crystallographically independent active sites in the ASU bDistances in the 1.16 Å resolution MCRox1-silent structure (PDB code 1hbn)(33) Biochemistry. Author manuscript; available in PMC 2011 September 7.
3M2U
Structural Insight into Methyl-Coenzyme M Reductase Chemistry using Coenzyme B Analogues
"Structural Insight into Methyl-Coenzyme M Reductase Chemistry\nusing Coenzyme B Analogues,†,‡\n(...TRUNCATED)
3M2V
Structural Insight into Methyl-Coenzyme M Reductase Chemistry using Coenzyme B Analogues
"Structural Insight into Methyl-Coenzyme M Reductase Chemistry\nusing Coenzyme B Analogues,†,‡\n(...TRUNCATED)
3M30
Structural Insight into Methyl-Coenzyme M Reductase Chemistry using Coenzyme B Analogues
"Structural Insight into Methyl-Coenzyme M Reductase Chemistry\nusing Coenzyme B Analogues,†,‡\n(...TRUNCATED)
3M31
Structure of the C150A/C295A mutant of S. cerevisiae Ero1p
"Steps in reductive activation of the\ndisulfide-generating enzyme Ero1p\nNimrod Heldman,1 Ohad Vons(...TRUNCATED)
3M32
Structural Insight into Methyl-Coenzyme M Reductase Chemistry using Coenzyme B Analogues
"Structural Insight into Methyl-Coenzyme M Reductase Chemistry\nusing Coenzyme B Analogues,†,‡\n(...TRUNCATED)
End of preview. Expand in Data Studio

primary_citation_with_pmcid.jsonl

This dataset links PDB protein structures with their corresponding primary ciatation text content.


Format

Each line is a JSON object:

{
  "protein_name": "9HCG",
  "structure_title": "Mouse mitoribosome large subunit assembly intermediate bound to NSUN4, MTERF4, and mt-RNAs",
  "main_text": "..."
}
Downloads last month
12