Skip to main content
NIHPA Author Manuscripts logoLink to NIHPA Author Manuscripts
. Author manuscript; available in PMC: 2010 Jun 15.
Published in final edited form as: Anal Biochem. 2009 Mar 28;389(2):174–176. doi: 10.1016/j.ab.2009.03.036

CD spectroscopy has intrinsic limitations for protein secondary structure analysis

Sergei Khrapunov 1
PMCID: PMC2696129  NIHMSID: NIHMS106230  PMID: 19332020

Abstract

Secondary structure content (SSC) cannot be accurately calculated from circular dichroism (CD) spectra for the majority of proteins whose three dimensional structures have been solved. ‘Reliable’ SSC that is significantly different from random SSC can be calculated from CD spectra only for all-α proteins and all-α proteins with canonical β-strand geometry.


The two fields to which the protein CD spectroscopy is applied with well-developed methodology are folding thermodynamics [1; 2] and secondary structure estimation [3]. Most thermodynamic studies rely on relative changes in CD spectra and are therefore relatively independent of calibration with structure. In contrast, the calculation of protein secondary structure content (SSC) requires strict cross-calibration/validation of experimental and reference CD spectra with reference crystallographic or NMR structural data. Following the development and dissemination of reliable CD analysis software via the internet [4; 5; 6; 7], improvements in SSC calculations have resulted from increasing the number of proteins in the protein reference set [4], splitting the ordered fractions of regular and distorted portions [6] and expanding CD spectral analysis to wavelengths below 185 nm using vacuum ultraviolet circular dichroism spectroscopy [8; 9]. Splitting the ordered fractions occurs when α-helices and β-strands are divided into regular and distorted classes [6] yielding six secondary structure classifications: regular α-helix (αR), distorted α-helix (αD), regular β-strand (βR), distorted β-strand (βD), turn and disordered.

The performances of secondary structure calculations are typically characterized by the root-mean-square deviations (RMSD) between the crystal and CD estimates of the secondary-structure content,

RMSD=(YiXi)2N, (1)

where Xi and Yi are the crystallographic and CD estimates of a given type of secondary structure, i, in N reference samples. The overall RMSD is determined by considering all secondary fractions collectively [4; 6; 8]. Lower values of RMSD indicate less discrepancy between the calculated and crystallographic data. It is generally accepted that RMSD measures the predictive power of the method.

Joint application of splitting the ordered fractions and utilization of the lower wavelength CD data (down to 160 nm if obtainable) yields in the best accuracy [8]. Overall RMSD values obtained for 29 of 31 studied proteins [8] are less than the overall RMSD values of 0.091 – 0.098 calculated on the basis of the splitting only [6].

Two questions arise from this and similar results. First, what is the lower limit for RMSD in such calculations? Second, is the accuracy that is reached sufficient to make a reliable and meaningful estimation of SSC for the proteins from their CD spectra? To answer these questions we have compared the overall RMSD for the 31 proteins calculated from their CD spectra [8] with the overall RMSD value obtained for simulated SSC assuming that the main secondary structure types (α-helices, β-strands, turns and disordered) are represented equally. The simulated SSC values were assumed equal to 0.125 each for regular and distorted helices and strands (αR, αD, βR, βD) and 0.25 each for turns and unordered structure.

Table 1 lists the RMSD values comparing crystallographic and CD SSC estimates (RMSDcd) and the RMSD values comparing the crystallographic estimates with those obtained from simulated SSC values of SSC (0.125 or 0.25 in the particular cases; RMSDs). The proteins in Table 1 are grouped accordingly to their tertiary structure class: all-α, all-α and αβ combining α+β and α/β classes [10; 11]. Peroxidase and xylanase are placed in the all-α and all-β groups since their helix/strand ratios are 13/2 and 1/15, respectively.

Table 1. The accuracy of secondary structure content (SSC) calculations of proteins from their CD spectra.

(α- all-α; β - all-β; αβ – combined α+β and α/β classes [10; 11]; RMSDcd –overall RMSD from [8]; RMSDs – simulated overall RMSD (see text); normal – RMSDs values less or comparable with RMSD; italic - RMSDs values less than overall RMSD (0.091) estimated for DSSP assignment [6]; bold - RMSDs values essentially higher than RMSD)

Protein Source PDB code Class RMSDcd RMSDs
Myoglobin Horse heart 1WLA α 0.03 0.205
Hemoglobin Bovine blood 1G08 α 0.048 0.201
Cytochrome C Horse heart 1HRC α 0.073 0.100
Albumin Human serum 1AO6 α 0.036 0.036
Peroxidase Horse radish 1ATJ α 0.038 0.102
Alfa-Lactalbumin Bovine milk 1F6S αβ 0.053 0.078
Lysozyme Hen egg 1HEL αβ 0.048 0.078
Ovalbumin Hen egg 1OVA αβ 0.065 0.064
RNase A Bovine pancreas 1FS3 αβ 0.038 0.042
β-Lactoglobulin Bovine milk 1B8E αβ 0.046 0.076
Pepsin Porcine stomach 4PEP αβ 0.053 0.073
Trypsinogen Bovine pancreas 1TGN αβ 0.047 0.063
α-hymotrypsinogen Bovine pancreas 2CGA αβ 0.016 0.060
Insulin Pig pancreas 4INS αβ 0.103 0.137
Glucose isomerase Strept. rubiginosus 1OAD αβ 0.048 0.083
Lactate dehydrogenase Bovine heart 8LDT αβ 0.031 0.083
Lipase Pseudomonas cepacia 3LIP αβ 0.044 0.048
Transferrin Human serum 1LFG αβ 0.032 0.038
Conalbumin Chicken egg white 1OVT αβ 0.038 0.035
Thioredoxin Escherichia coli 2TRX αβ 0.065 0.054
Catalase Bovine liver 7CAT αβ 0.038 0.055
α-amylase Bacillus subtilis 1BAG αβ 0.032 0.038
Subtilisin A Bacillus subtilis 1SBC αβ 0.047 0.035
Papain Papaya 8PAP αβ 0.065 0.071
Elastase Porcine pancreas 3EST αβ 0.031 0.076
Carbonic anhydrase Bovine erythrocytes 1G6V αβ 0.05 0.058
Trypsin inhibitor Soybean 1AVU β 0.058 0.114
Concanavalin A Jack bean 2CTV β 0.051 0.104
Xylanase Trichoderma 1ENX β 0.108 0.158
Avidin Chicken egg white 1AVE β 0.082 0.110

It is readily evident in the table that except for human serum albumin, only for the proteins belonging to all-α and all-β classes are the simulated RMSD values essentially higher than the experimental ones and values higher than the overall RMSD value of 0.091 estimated for 29 proteins for DSSP assignments [6]. Of the 22 proteins of the αβ class, 10 show simulated RMSDs lower or comparable with experimental ones while the remainder has simulated RMSD lower than the overall value of the DSSP assignment [6]. The only exception is insulin, presumably due to its small dimension, short structural elements and uncertainty in its intermolecular interactions in solution [12]. These properties are expected to influence the CD spectrum of insulin. Moreover individual β-sheets have variable CD spectra due to the variations in the geometry of β-structure in proteins [13]. If β-strands are within an unusual structural motif like the Pentapeptide Repeat Protein fold, the SSC calculated from CD and crystallographic data demonstrate poor correspondence [14]. Thus the analyzed proteins included in all-β class all apparently have canonical β-structure.

As follows from this analysis SSC cannot be accurately calculated from CD spectra for the vast majority of the proteins. Reliable calculation of SSC from CD spectra can be made only for all-α proteins and all-α proteins whose β-strands geometry is apparently canonical. Why does this occur and can the method be ‘repaired’? We propose there are some intrinsic limits to the application of CD spectroscopy to protein secondary structure calculation as summarized below: a) The quality of the Ramachandran plots of the reference crystallographic structures is poor for some reference CD datasets. In fact, the residues for most of the structures in some protein reference sets are under the 90% and even the 80% thresholds for the most favored region of the Ramachandran plot [15]; b) The quality of the proteins used for solution and crystallographic studies may not be consistent. Many of the reference CD spectra are obtained using commercially prepared proteins without purification [4; 8]; c) The consistency of the reference CD database sets is sometimes suspect. The spectrum of some proteins differs in different databases [15]; d) There has been little cross validation of the instruments used to obtain reference and experimental CD spectra. Many reference CD spectra were obtained long ago sometimes on laboratory-specific instruments whose specifications are not documented. Perhaps creation of a central resource of the published and cross-validated CD data files such as the proposed Protein Circular Dichroism Data Bank [16] can help solve these problems.

Lastly, the different algorithms used to calculate protein SSC from crystallographic structures give average contents for particular structures with standard deviations comparable with the RMSD values shown in the Table 1 [6; 9]. Unlike the above problems this one cannot be easily fixed. Its solution requires the mutual agreement of the scientific community on the ceasing the indiscriminate use of the programs DSSP, Procheck, STRIDE, XtlSSTR and PROMOTIF in favor of one. We favor the DSSP algorithm as it is mostly used for PDB files.

We conclude that a reliable and meaningful estimation of SSC from the CD spectra can not be made for proteins with the mixed α and β elements in their structure and apparently for proteins with the noncanonical β-strand geometry. At the same time such estimations can be used as relative measures of the structural changes of the proteins at different conditions such as those observed during folding and unfolding.

Acknowledgments

The author thanks Dr. Michael Brenowitz for his constant support and generous sharing of many valuable suggestions. This study was supported by grant R01-GM079618 from the Institute of General Medical Sciences of the National Institutes of Health.

Footnotes

Publisher's Disclaimer: This is a PDF file of an unedited manuscript that has been accepted for publication. As a service to our customers we are providing this early version of the manuscript. The manuscript will undergo copyediting, typesetting, and review of the resulting proof before it is published in its final citable form. Please note that during the production process errors may be discovered which could affect the content, and all legal disclaimers that apply to the journal pertain.

References

  • 1.Greenfield N. Using circular dichroism collected as a function of temperature to determine the thermodynamics of protein unfolding and binding interactions. Nature Protocols. 2006;1:2876–2880. doi: 10.1038/nprot.2006.204. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 2.Greenfield N. Determination of the folding of proteins as a function of denaturants, osmolytes or ligands using circular dichroism. Nature Protocols. 2006;1:2733–2741. doi: 10.1038/nprot.2006.229. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 3.Greenfield N. Using circular dichroism spectra to estimate protein secondary structure. Nature Protocols. 2006;1:2876–2890. doi: 10.1038/nprot.2006.202. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 4.Sreerama N, Woody R. Estimation of protein secondary structure from circular dichroism spectra: comparison of CONTIN, SELCON, and CDSSTR methods with an expanded reference set. Analitical Biochemistry. 2000;287:252–260. doi: 10.1006/abio.2000.4880. [DOI] [PubMed] [Google Scholar]
  • 5.Whitmore L, Wallace B. DICHROWEB, an online server for protein secondary structure analyses from circular dichroism spectroscopic data. Nucleic Acids Research. 2004;32:W668–W673. doi: 10.1093/nar/gkh371. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 6.Sreerama N, Venyaminov S, Woody R. Estimation of the number of alpha-helical and beta-strand using circular dichroism spectroscopy. Protein Science. 1999;8:370–380. doi: 10.1110/ps.8.2.370. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 7.Sreerama N, Venyaminov S, Woody R. Analysis of Protein Circular Dichroism Spectra Based on the Tertiary Structure Classification. Analitical Biochemistry. 2001;299:271–274. doi: 10.1006/abio.2001.5420. [DOI] [PubMed] [Google Scholar]
  • 8.Matsuo K, Yonehara R, Gekko K. Imroved estimation of the secondary structures of Proteins by vacuum-ultraviolet circular Dichroism spectroscopy. Journal of Biochemistry. 2005;138:79–88. doi: 10.1093/jb/mvi101. [DOI] [PubMed] [Google Scholar]
  • 9.Wallace B, Lees J, Orry A, Lobley A, Janes R. Analyses of circular dichroism spectra of membrane proteins. Protein Science. 2003;12:875–884. doi: 10.1110/ps.0229603. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 10.Levitt M, Chotia C. Structural patterns in globular proteins. Nature. 1976;261:552–558. doi: 10.1038/261552a0. [DOI] [PubMed] [Google Scholar]
  • 11.Manalavan P, Johnson WJ. Sensitivity of circular dichroism to protein tertiary structure class. Nature. 1983;305:831–832. [Google Scholar]
  • 12.Grudzielanek S, Jansen R, Winter R. Solvational Tuning of the Unfolding, Aggregation and Amyloidogenesis of Insulin. Journal of Molecular Biology. 2005;351:879–894. doi: 10.1016/j.jmb.2005.06.046. [DOI] [PubMed] [Google Scholar]
  • 13.Sreerama N, Woody R. Computation and analysis of protein circular dichroism spectra. Methods in Enzymology. 2004;383:318–351. doi: 10.1016/S0076-6879(04)83013-1. [DOI] [PubMed] [Google Scholar]
  • 14.Khrapunov S, Cheng H, Hegde S, Blanchard J, Brenowitz M. Solution Structure and Refolding of the Mycobacterium tuberculosis Pentapeptide Repeat Protein MfpA. Journal of Biological Chemistry. 2008;283:36290–36299. doi: 10.1074/jbc.M804702200. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 15.Janes R. Bioinformatics analysis of circular dichroism protein reference databases. Bioinformatics. 2005;21:4230–4238. doi: 10.1093/bioinformatics/bti690. [DOI] [PubMed] [Google Scholar]
  • 16.Wallace B, Lee W, Janes R. The Protein Circular Dichroism Data Bank (PCDDB): A Bioinformatics and Spectroscopic Resource. Proteins. 2006;62:1–3. doi: 10.1002/prot.20676. [DOI] [PubMed] [Google Scholar]

RESOURCES