Merck
CN
HomeNuclear Magnetic Resonance (NMR)Isotopic Labeling for NMR Spectroscopy of Biological Solids

Isotopic Labeling for NMR Spectroscopy of Biological Solids

Mei Hong

Department of Chemistry, Iowa State University Ames, Iowa

ISOTEC® Stable Isotopes: Products for Solid State NMR, 2010, 6–11

Isotopic labeling plays an indispensable role in structure determination of proteins and other biomacromolecules using solidstate NMR. It not only enhances the NMR sensitivity but also allows for site-specific interrogation of structures and intermolecular contacts. This article gives a survey of the different isotopic labeling approaches available today for biological solid-state NMR research.

Biosynthetic uniform 13C, 15N labeling

The simplest and most cost-effective biosynthetic labeling method for protein solid-state NMR is to uniformly label all carbon and nitrogen atoms with 13C and 15N. In this way, a single protein sample can in principle provide all the structural constraints – dihedral angles and distances - about the protein. The labeled precursors are typically uniformly (U) 13C-labeled glucose or glycerol, and 15N-labeled ammonium chloride or ammonium sulfate. These compounds can be readily incorporated into the growth media for protein expression. Uniform 13C, 15N-labeling has seen the most widespread application in the development of new magic-anglespinning (MAS) multidimensional correlation techniques for full structure determination of proteins. A number of microcrystalline proteins whose structures are known from X-ray crystallography or solution NMR have been used to demonstrate the ability of solidstate NMR to obtain de novo three-dimensional structures. These microcrystalline proteins include ubiquitin1,2, GB13,4, thioredoxin5, and the a-spectrin SH3 domain6. Uniform 13C and 15N labeling has also been used effectively in structure determination of amyloid fibril proteins, such as transthyretin7, the HET-s prion protein8, and a human prion protein9. A common feature of the proteins amenable to this labeling scheme is that they possess sufficient structural order on the nanometer scale to give highly resolved spectra. Without this high conformational homogeneity and the resulting high spectral resolution, uniform 13C labeling is not recommended since it would cause considerable spectral congestion. Various 2D, 3D1,10,11, and 4D12 correlation techniques have been developed to resolve the signals of uniformly 13C, 15N-labeled proteins and to determine internuclear distances and dihedral angles.

Uniform 13C and 15N labeling has also been applied to a handful of membrane proteins, such as potassium ion channels13, seventransmembrane-helix proteins14,15, light-harvesting complexes16, membrane-bound enzymes17, and bacterial toxins18. Since membrane proteins usually have larger conformational disorder than microcrystalline proteins or fibril-forming proteins, the spectral resolution of membrane proteins is generally lower. Nevertheless, detailed structural information of key regions of these membrane proteins or the global topology of membrane proteins in the lipid bilayer, such as their depth of insertion, could still be obtained even using uniformly 13C, 15N-labeled samples.

The main spectroscopic challenges involved in MAS NMR of uniformly 13C-labeled proteins are three-fold: 1) the limited dispersion of 13C isotropic chemical shifts given the inhomogeneous linewidths of the sample; 2) the 13C-13C scalar couplings that contribute to line broadening; and 3) the dipolar truncation effect that makes it difficult to measure long-range 13C-13C distances in the presence of strong one-bond 13C-13C dipolar couplings. Static 15N NMR of oriented membrane peptides and proteins do not have these challenges, since the spectral dispersion is determined by the much larger anisotropic chemical shift range rather than the isotropic chemical shift range, and because there is no 15N-15N scalar coupling nor any sizeable 15N-15N dipolar coupling in proteins. Therefore, uniform 15N labeling entails few complications for orientation determination of membrane proteins and indeed has seen fruitful applications19,20. On the other hand, it is clearly desirable to increase the information content of the aligned sample spectra by including 13C dimensions. New spectroscopic challenges need to be overcome in 13C NMR of oriented membrane proteins. For example, 13C-13C dipolar couplings of U-13C-labeled proteins are no long removed by MAS in these static samples. Strategies for decoupling the 13C-13C couplings and for correlation experiments under the static condition have been proposed and demonstrated on single crystal model compounds21. Random fractional 13C labeling, which strikes a compromise between resolution and structural information, has also been proposed22.

Biosynthetic selective 13C labeling

Two of the three challenges listed above for studying U-13C labeled proteins are nicely addressed by the complementary approach of selective 13C labeling. In this approach, carbon precursors that contain only specific 13C-labeled sites are incorporated into the protein expression media. These labeled sites are converted, through well-known enzymatic pathways23, to predictable positions in the twenty amino acids, which result in selectively and extensively labeled proteins. All residues of the same amino acid type have the same labeled positions, but different amino acids have different labeled positions due to their distinct enzymatic pathways.

The two main precursors that have been demonstrated are [2-13C] glycerol, which primarily label the Cα carbons of amino acids, and [1,3-13C] glycerol, which label the other sites skipped by [2-13C] glycerol. Each precursor tends to label alternating carbons, thus removing any sizeable 13C-13C scalar couplings and the trivial one-bond dipolar couplings. This selective labeling approach was originally proposed by LeMaster and Kushlan for solution NMR studies and subsequently adopted for solid-state NMR24-26. By far the most important application of selective 13C labeling is distance extraction from 13C-13C correlation spectra. Other amino acid precursors can in principle also be exploited, for example, oxaloacetate, a-ketoglutarate, and pyruvate, as having been done in protein solution NMR. In addition, 13C-labeled carbon dioxide has been used for studying plant cell wall proteins27,28.

Reverse labeling: combining biosynthetic labeling with unlabeled amino acids

Another strategy to reduce the spectral congestion without resorting to amino-acid-specific labeling is to combine a labeled general carbon precursor with unlabeled amino acids, so that only a subset of amino acid types will be labeled. For membrane protein structural studies, one version of this strategy is the TEASE (ten-amino-acid-selective-and-extensive) labeling protocol25. In this approach, [2-13C] glycerol and ten unlabeled amino acids serve as the carbon precursors of the expression media. The ten amino acids are Glu, Gln, Pro, Arg, Asp, Asn, Met, Thr, Ile, and Lys, which are products of the citric acid cycle. Normally, the cycle distributes the 13C labels in glucose or glycerol to produce fractionally labeled sites in these amino acids, so that their signals are more difficult to assign in the NMR spectra than amino acids synthesized from the glycolysis pathway. Due to the approximate hydrophobic versus hydrophilic distinction of the amino acids from the glycolysis pathway versus the citric acid cycle, a membrane protein could in principle be TEASE 13C-labeled to selectively detect the transmembrane segments rich in the hydrophobic residues.

Clearly, this reverse labeling approach is highly flexible and can be adapted for different applications. For example, a U-13C-labeled precursor can be combined with a small set of unlabeled amino acids that are dominant in the protein. Unlabeling of these amino acid types simplifies the NMR spectra considerably14, and does not bring any disadvantages to the protein expression.

Site-specific labeling of synthetic peptides and proteins

Site-specific 13C and 15N labeling continues to provide rich structural information about polypeptides that are too small to be recombinantly expressed or proteins that are too large for uniformly 13C-labeled spectra to be analyzable. For polypeptides shorter than 40 amino acids, chemical synthesis is generally feasible, therefore 13C, 15N-labeled amino acids in their protected forms can be incorporated into the peptide synthesis for sitespecific labeling.

A common site-specific amino acid labeling strategy is the scattered uniform 13C, 15N-labeling of residues. As long as the yield of the peptide synthesis is not prohibitively low, the combination of several samples with different U-13C, 15N-labeled residues can eventually map out the complete structure of the polypeptide of interest. This approach has been used extensively to study amyloid peptides29 and membrane peptides30-32. Non-uniform 13C and 15N labeling of specific amino acid residues has also been applied. The most commonly labeled sites are the 13CO of the polypeptide backbone, and sometimes the sidechain 15N of lysine residues. Applications usually involve distances measurements using heteronuclear REDOR33 or homonuclear 13C recoupling34 experiments.

Since most peptides are synthesized using the Fmoc solid phase chemistry, site-specific amino acid labeling requires Fmoc-protected amino acids. For hydrophobic amino acids, their Fmoc-protected forms are usually commercially available and can also be synthesized readily from their unprotected forms. On the other hand, polar amino acids require both backbone and sidechain protection, thus are more costly and difficult to prepare. While Fmoc solid-phase synthesis is the dominant chemistry in peptide synthesis, t-Boc solid-phase synthesis has also been used for interesting structure determination targets35. Boc-protected 13C, 15N-labeled amino acids are so far much less common. Therefore, increased commercial production and availability of t-Boc-protected amino acids are desirable.

Other isotopic labels for studying macromolecular complexes and protein chemistry

For large macromolecular complexes such as the cell walls of plants and bacteria, and for membrane proteins bound to ligands or inhibitors, it is often important to increase the diversity of isotopic labeling to enable intermolecular distance measurements. Two isotopes are readily available for this purpose: 2H and 19F. 19F is naturally 100% abundant and has a long history of being incorporated into amino acids36-38 as well as non-peptidic molecules such as lipids and pharmaceutical drugs39. Sitespecific 2H labeling is most commonly used for methyl groups of Ala, Leu, and Val, and is an excellent probe of the dynamics of proteins40,41 and DNA42. More recently, perdeuteration of proteins in combination with uniform 13C and 15N labeling has been exploited as a means to obtain high-resolution spectra of proteins, as perdeuteration removes 1H dipolar coupling as a line broadening mechanism. The back-exchanged proteins have 1H spins only at exchangeable positions such as the amide hydrogens and lysine amino groups. These sparse protons can be used as a high-sensitivity detection nucleus. Perdeuterated microcrystalline proteins have been used to study relaxation dynamics of proteins and protein-water interactions43-45.

To produce 13C/15N/2H triply labeled recombinant proteins, one needs to use 2H and 13C labeled glucose, which is commercially available. The main challenge in this type of protein expression is for the cells to tolerate a water-deuterated liquid culture, which usually decreases the protein expression yield.

Future prospects

Isotopic labeling is an essential and versatile tool for NMR structural biology. Creative labeling of NMR-sensitive nuclei (13C, 15N, and 2H), combined with strategic exploitation of naturally 100% abundant nuclei such as 19F and 31P, can advance the structural biology of many insoluble macromolecules important in biology.

For future progress in solid-state NMR structural biology, it will be important to develop a more diverse panel of isotopically labeled compounds and to produce the existing compounds at a more economical level. Since biosynthetically obtained 13C-labeled precursors are ubiquitous and relatively simple to produce, one of the future challenges is a chemical one, which is to produce a diverse array of specifically labeled specifically labeled amino acids and other small biomolecules with isotopic labels at desired positions.

Materials
Loading

References

1.
Hong M. 1999. 15(1):1-14. https://doi.org/10.1023/a:1008334204412
2.
Igumenova TI, McDermott AE, Zilm KW, Martin RW, Paulson EK, Wand AJ. 2004. Assignments of Carbon NMR Resonances for Microcrystalline Ubiquitin. J. Am. Chem. Soc.. 126(21):6720-6727. https://doi.org/10.1021/ja030547o
3.
Franks WT, Zhou DH, Wylie BJ, Money BG, Graesser DT, Frericks HL, Sahota G, Rienstra CM. 2005. Magic-Angle Spinning Solid-State NMR Spectroscopy of the ?1 Immunoglobulin Binding Domain of Protein G (GB1): 15N and13C Chemical Shift Assignments and Conformational Analysis. J. Am. Chem. Soc.. 127(35):12291-12305. https://doi.org/10.1021/ja044497e
4.
Chen L, Olsen RA, Elliott DW, Boettcher JM, Zhou DH, Rienstra CM, Mueller LJ. 2006. Constant-Time Through-Bond13C Correlation Spectroscopy for Assigning Protein Resonances with Solid-State NMR Spectroscopy. J. Am. Chem. Soc.. 128(31):9992-9993. https://doi.org/10.1021/ja062347t
5.
Marulanda D, Tasayco ML, Cataldi M, Arriaran V, Polenova T. 2005. Resonance Assignments and Secondary Structure Analysis ofE.coliThioredoxin by Magic Angle Spinning Solid-State NMR Spectroscopy. J. Phys. Chem. B. 109(38):18135-18145. https://doi.org/10.1021/jp052774d
6.
Pauli J, Baldus M, vanRossum B, Groot Hd, Oschkinat H. 2001 . Backbone and side-chain 13C and 15N signal assignments of the alpha-spectrin SH3 domain by magic angle spinning solid-state NMR at 17.6 Tesla ChemBioChem. 2 272-281.
7.
Jaroniec CP, MacPhee CE, Astrof NS, Dobson CM, Griffin RG. 2002. Molecular conformation of a peptide fragment of transthyretin in an amyloid fibril. Proceedings of the National Academy of Sciences. 99(26):16748-16753. https://doi.org/10.1073/pnas.252625999
8.
Wasmer C, Lange A, Van Melckebeke H, Siemer AB, Riek R, Meier BH. 2008. Amyloid Fibrils of the HET-s(218-289) Prion Form a   Solenoid with a Triangular Hydrophobic Core. Science. 319(5869):1523-1526. https://doi.org/10.1126/science.1151839
9.
Helmus JJ, Surewicz K, Nadaud PS, Surewicz WK, Jaroniec CP. 2008. Molecular conformation and dynamics of the Y145Stop variant of human prion protein in amyloid fibrils. Proceedings of the National Academy of Sciences. 105(17):6284-6289. https://doi.org/10.1073/pnas.0711716105
10.
Rienstra CM, Hohwy M, Hong M, Griffin RG. 2000. 2D and 3D15N?13C?13C NMR Chemical Shift Correlation Spectroscopy of Solids:  Assignment of MAS Spectra of Peptides. J. Am. Chem. Soc.. 122(44):10979-10990. https://doi.org/10.1021/ja001092v
11.
Heise H, Seidel K, Etzkorn M, Becker S, Baldus M. 2005. 3D NMR spectroscopy for resonance assignment and structure elucidation of proteins under MAS: novel pulse schemes and sensitivity considerations. Journal of Magnetic Resonance. 173(1):64-74. https://doi.org/10.1016/j.jmr.2004.11.020
12.
Franks WT, Kloepper KD, Wylie BJ, Rienstra CM. 2007. Four-dimensional heteronuclear correlation experiments for chemical shift assignment of solid proteins. J Biomol NMR. 39(2):107-131. https://doi.org/10.1007/s10858-007-9179-1
13.
Lange A, Giller K, Hornig S, Martin-Eauclaire M, Pongs O, Becker S, Baldus M. 2006. Toxin-induced conformational changes in a potassium channel revealed by solid-state NMR. Nature. 440(7086):959-962. https://doi.org/10.1038/nature04649
14.
Etzkorn M, Martell S, Andronesi O, Seidel K, Engelhard M, Baldus M. 2007. Secondary Structure, Dynamics, and Topology of a Seven-Helix Receptor in Native Membranes, Studied by Solid-State NMR Spectroscopy. Angew. Chem. Int. Ed.. 46(3):459-462. https://doi.org/10.1002/anie.200602139
15.
Shi L, Ahmed MA, Zhang W, Whited G, Brown LS, Ladizhansky V. 2009. Three-Dimensional Solid-State NMR Study of a Seven-Helical Integral Membrane Proton Pump?Structural Insights. Journal of Molecular Biology. 386(4):1078-1093. https://doi.org/10.1016/j.jmb.2009.01.011
16.
Huang L, McDermott AE. 2008. Partial site-specific assignment of a uniformly 13C, 15N enriched membrane protein, light-harvesting complex 1 (LH1), by solid state NMR. Biochimica et Biophysica Acta (BBA) - Bioenergetics. 1777(9):1098-1108. https://doi.org/10.1016/j.bbabio.2008.01.006
17.
Li Y, Berthold DA, Gennis RB, Rienstra CM. 2008. Chemical shift assignment of the transmembrane helices of DsbB, a 20-kDa integral membrane enzyme, by 3D magic-angle spinning NMR spectroscopy. Protein Sci.. 17(2):199-204. https://doi.org/10.1110/ps.073225008
18.
Huster D, Yao X, Jakes K, Hong M. 2002. Conformational changes of colicin Ia channel-forming domain upon membrane binding: a solid-state NMR study. Biochimica et Biophysica Acta (BBA) - Biomembranes. 1561(2):159-170. https://doi.org/10.1016/s0005-2736(02)00340-1
19.
Marassi FM, Ma C, Gratkowski H, Straus SK, Strebel K, Oblatt-Montal M, Montal M, Opella SJ. 1999. Correlation of the structural and functional domains in the membrane protein Vpu from HIV-1. Proceedings of the National Academy of Sciences. 96(25):14336-14341. https://doi.org/10.1073/pnas.96.25.14336
20.
Tian C, Gao PF, Pinto LH, Lamb RA, Cross TA. Initial structural and dynamic characterization of the M2 protein transmembrane and amphipathic helices in lipid bilayers. Protein Science. 12(11):2597-2605. https://doi.org/10.1110/ps.03168503

Reference

21.
Ishii Y, Tycko R. 2000. Multidimensional Heteronuclear Correlation Spectroscopy of a Uniformly15N- and13C-Labeled Peptide Crystal:  Toward Spectral Resolution, Assignment, and Structure Determination of Oriented Molecules in Solid-State NMR. J. Am. Chem. Soc.. 122(7):1443-1455. https://doi.org/10.1021/ja9915753
22.
Filipp FV, Sinha N, Jairam L, Bradley J, Opella SJ. 2009. Labeling strategies for 13C-detected aligned-sample solid-state NMR of proteins. Journal of Magnetic Resonance. 201(2):121-130. https://doi.org/10.1016/j.jmr.2009.08.012
23.
Lehninger AL, Nelson DL, Cox MM. 1993. Principles of Biochemistry . 2nd ed. . New York: Worth Publishers .
24.
Hong M. 1999. Determination of Multiple ?-Torsion Angles in Proteins by Selective and Extensive 13C Labeling and Two-Dimensional Solid-State NMR. Journal of Magnetic Resonance. 139(2):389-401. https://doi.org/10.1006/jmre.1999.1805
25.
Hong M, Jakes K. 1999. 14(1):71-74. https://doi.org/10.1023/a:1008334930603
26.
Castellani F, van Rossum B, Diehl A, Schubert M, Rehbein K, Oschkinat H. 2002. Structure of a protein determined by solid-state magic-angle-spinning NMR spectroscopy. Nature. 420(6911):99-102. https://doi.org/10.1038/nature01070
27.
Cegelski L, Schaefer J. 2005. Glycine Metabolism in Intact Leaves byin Vivo13C and15N Labeling. J. Biol. Chem.. 280(47):39238-39245. https://doi.org/10.1074/jbc.m507053200
28.
Cegelski L, Schaefer J. 2006. NMR determination of photorespiration in intact leaves using in vivo 13CO2 labeling. Journal of Magnetic Resonance. 178(1):1-10. https://doi.org/10.1016/j.jmr.2005.10.010
29.
Petkova AT, Ishii Y, Balbach JJ, Antzutkin ON, Leapman RD, Delaglio F, Tycko R. 2002. A structural model for Alzheimer's  -amyloid fibrils based on experimental constraints from solid state NMR. Proceedings of the National Academy of Sciences. 99(26):16742-16747. https://doi.org/10.1073/pnas.262663499
30.
Cady SD, Mishanina TV, Hong M. 2009. Structure of Amantadine-Bound M2 Transmembrane Peptide of Influenza A in Lipid Bilayers from Magic-Angle-Spinning Solid-State NMR: The Role of Ser31 in Amantadine Binding. Journal of Molecular Biology. 385(4):1127-1141. https://doi.org/10.1016/j.jmb.2008.11.022
31.
Mani R, Cady SD, Tang M, Waring AJ, Lehrer RI, Hong M. 2006. Membrane-dependent oligomeric structure and pore formation of a beta-hairpin antimicrobial peptide in lipid bilayers from solid-state NMR. Proceedings of the National Academy of Sciences. 103(44):16242-16247. https://doi.org/10.1073/pnas.0605079103
32.
Tang M, Waring A, Lehrer R, Hong M. 2008. Effects of Guanidinium?Phosphate Hydrogen Bonding on the Membrane-Bound Structure and Activity of an Arginine-Rich Membrane Peptide from Solid-State NMR Spectroscopy. Angew. Chem. Int. Ed.. 47(17):3202-3205. https://doi.org/10.1002/anie.200705993
33.
Qiang W, Sun Y, Weliky DP. 2009. A strong correlation between fusogenicity and membrane insertion depth of the HIV fusion peptide. Proceedings of the National Academy of Sciences. 106(36):15314-15319. https://doi.org/10.1073/pnas.0907360106
34.
Long JR, Dindot JL, Zebroski H, Kiihne S, Clark RH, Campbell AA, Stayton PS, Drobny GP. 1998. A peptide that inhibits hydroxyapatite growth is in an extended conformation on the crystal surface. Proceedings of the National Academy of Sciences. 95(21):12083-12087. https://doi.org/10.1073/pnas.95.21.12083
35.
Wu Z, Ericksen B, Tucker K, Lubkowski J, Lu W. 2004. Synthesis and characterization of human alpha-defensins 4-6. J Pept Res. 64(3):118-125. https://doi.org/10.1111/j.1399-3011.2004.00179.x
36.
Afonin S, Glaser RW, Berditchevskaia M, Wadhwani P, Gührs K, Möllmann U, Perner A, Ulrich AS. 2003. 4-Fluorophenylglycine as a Label for 19F NMR Structure Analysis of Membrane-Associated Peptides. ChemBioChem. 4(11):1151-1163. https://doi.org/10.1002/cbic.200300568
37.
Grage SL, Ulrich AS. 2000. Orientation-Dependent 19F Dipolar Couplings within a Trifluoromethyl Group Are Revealed by Static Multipulse NMR in the Solid State. Journal of Magnetic Resonance. 146(1):81-88. https://doi.org/10.1006/jmre.2000.2127
38.
Luo W, Mani R, Hong M. 2007. Side-Chain Conformation of the M2 Transmembrane Peptide Proton Channel of Influenza A Virus from19F Solid-State NMR. J. Phys. Chem. B. 111(36):10825-10832. https://doi.org/10.1021/jp073823k
39.
Toke O, Lee Maloy W, Kim SJ, Blazyk J, Schaefer J. 2004. Secondary Structure and Lipid Contact of a Peptide Antibiotic in Phospholipid Bilayers by REDOR. Biophysical Journal. 87(1):662-674. https://doi.org/10.1529/biophysj.103.032706
40.
Cady SD, Goodman C, Tatko CD, DeGrado WF, Hong M. 2007. Determining the Orientation of Uniaxially Rotating Membrane Proteins Using Unoriented Samples:  A2H,13C, and15N Solid-State NMR Investigation of the Dynamics and Orientation of a Transmembrane Helical Bundle. J. Am. Chem. Soc.. 129(17):5719-5729. https://doi.org/10.1021/ja070305e
41.
Williams JC, McDermott AE. 1995. Dynamics of the Flexible Loop of Triose-Phosphate Isomerase: The Loop Motion Is Not Ligand Gated. Biochemistry. 34(26):8309-8319. https://doi.org/10.1021/bi00026a012
42.
Meints GA, Karlsson T, Drobny GP. 2001. Modeling Furanose Ring Dynamics in DNA. J. Am. Chem. Soc.. 123(41):10030-10038. https://doi.org/10.1021/ja010721d
43.
Morcombe CR, Gaponenko V, Byrd RA, Zilm KW. 2005. 13C CPMAS Spectroscopy of Deuterated Proteins:  CP Dynamics, Line Shapes, andT1Relaxation. J. Am. Chem. Soc.. 127(1):397-404. https://doi.org/10.1021/ja045581x
44.
Akbey Ü, Lange S, Trent Franks W, Linser R, Rehbein K, Diehl A, van Rossum B, Reif B, Oschkinat H. 2010. Optimum levels of exchangeable protons in perdeuterated proteins for proton detection in MAS solid-state NMR spectroscopy. J Biomol NMR. 46(1):67-73. https://doi.org/10.1007/s10858-009-9369-0
45.
Lesage A, Emsley L, Penin F, Böckmann A. 2006. Investigation of Dipolar-Mediated Water?Protein Interactions in Microcrystalline Crh by Solid-State NMR Spectroscopy. J. Am. Chem. Soc.. 128(25):8246-8255. https://doi.org/10.1021/ja060866q
登录以继续。

如要继续阅读,请登录或创建帐户。

暂无帐户?