International
Tables for
Crystallography
Volume F
Crystallography of biological macromolecules
Edited by M. G. Rossmann and E. Arnold

International Tables for Crystallography (2006). Vol. F. ch. 19.6, pp. 461-462   | 1 | 2 |

Section 19.6.4.7. Visualization, modelling and interpretation of results

T. S. Bakera* and R. Hendersonb

a Department of Biological Sciences, Purdue University, West Lafayette, Indiana 47907-1392, USA, and bMedical Research Council, Laboratory of Molecular Biology, Hills Road, Cambridge CB2 2QH, England
Correspondence e-mail:  tsb@bragg.bio.purdue.edu

19.6.4.7. Visualization, modelling and interpretation of results

| top | pdf |

Once a reliable 3D map is obtained, computer graphics and other visualization tools may be used as aids in interpreting morphological details and understanding biological function in the context of biochemical and molecular studies and complementary X-ray crystallographic and other biophysical measurements.

Initially, for low-resolution studies (>10 Å) and where the structure is unknown, the gross shape (molecular envelope) of the macromolecule is best visualized with volume-rendering programs (e.g. Conway et al., 1996[link]; Sheehan et al., 1996[link]; Spencer et al., 1997[link]). Such programs establish a density threshold, above which all density is represented as a solid and below which all density is invisible (representing possible solvent regions). Choice of the threshold that accurately represents the solvent-excluded density can prove problematic, especially if the microscope CTF is uncorrected (e.g. Conway et al., 1996[link]). For qualitative examination of maps, a threshold at 1.5 or 2 standard deviations above the background noise level provides a practical choice. Another, semi-quantitative, approach is to adjust the threshold to produce a volume consistent with the expected total molecular mass. This procedure is prone to error because the volume is sensitive to small changes in contour level, which, in turn, is highly sensitive to scaling and CTF correction. Caution should therefore be exercised in drawing conclusions based on volume fluctuations of less than 20% of that expected. As a general guide, solid-surface rendering in the range 80 to 120% of the expected volume gives reasonable shape and connectivity.

A complete description of available graphical tools for visualizing 3D density maps is beyond the scope of this discussion, but it is worth noting several of the principles by which 3D data can be rendered. Stereo images (e.g. Liu et al., 1994[link]; Agrawal et al., 1996[link]; Taveau, 1996[link]; Winkler et al., 1996[link]; Nogales et al., 1997[link]; Kolodziej et al., 1998[link]; Gabashvili et al., 2000[link]) provide a powerful way to convey the 3D structure. Also, as in X-ray crystallographic applications, stereo viewing is essential for exploring details of secondary and tertiary structural information in high-resolution 3D maps (e.g. Henderson et al., 1990[link]; Böttcher, Wynne & Crowther, 1997[link]). Additional visualization tools include: use of false colour to highlight distinct components (e.g. Yeager et al., 1994[link]; Cheng et al., 1995[link]; Hirose et al., 1997[link]; Metoz et al., 1997[link]; Zhou et al., 1999[link]) and a variety of computer `sectioning' or image-projection algorithms that produce cut-open views (e.g. Vigers et al., 1986[link]; Cheng et al., 1994[link]; Fuller et al., 1995[link]), spherical sections (e.g. Baker et al., 1991[link]), icosahedrally cut surfaces (Böttcher, Kiselev et al., 1997[link]), polar sections (Fuller et al., 1995[link]), cylindrical sections (e.g. Hirose et al., 1997[link]), radial projections (e.g. Dryden et al., 1993[link]), and radial depth cueing (Spencer et al., 1997[link]), which conveys an immediate, and often quantitative, view of the radial placement of details in a map (Grimes et al., 1997[link]). Animation also provides an alternative approach to enhance the viewer's perception of the 3D structure (e.g. van Heel et al., 1996[link]; Frank et al., 1999[link]). All these rendering methods should always be carefully described so the reader may distinguish representation from result.

Difference imaging and density-map modelling are examples of two additional techniques that can sometimes enhance interpretation of 3D cryo EM data.

Difference imaging is a very powerful tool, long employed by structural biologists outside the cryo EM field, that permits small (or large) differences among closely related structures to be examined. One of the great advantages of cryomicroscopy of ice-embedded specimens over microscopy using negative stains is that cryo EM difference imaging yields more reliable results as confirmed by correlation with biochemical and immunological data (e.g. Baker et al., 1990[link]; Stewart et al., 1993[link]; Vénien-Bryan & Fuller 1994[link]; Yeager et al., 1994[link]; Hoenger & Milligan, 1997[link]; Lawton et al., 1997[link]; Stewart et al., 1997[link]; Böttcher et al., 1998[link]; Conway et al., 1998[link]; Sharma et al., 1998[link]; Zhou, Mcnab et al., 1998[link]). However, reliable interpretation is only possible if the difference maps are carefully calculated (i.e. from two maps calculated to the same resolution and scaled in such a way that the differences are minimized). Subtraction of two maps, each having an intrinsic noise level, guarantees that the difference map will always be noisier than either of the parent maps, and noise in the difference map is what determines the significance of features in it. Careful statistical analysis is an important prerequisite in attributing significance to and interpreting particular features. One critical test is to see whether a difference of a similar size can be found between independent determinations of the same structure. Differences that occur at symmetry axes must be treated cautiously, because the noise level there is greater.

The combination of X-ray and cryo EM data provides a powerful tool for interpreting structures [e.g. see the review by Baker & Johnson (1996[link])]. A high-resolution X-ray model can be docked into a cryo EM density map with greater precision than the nominal resolution of the map. Several similar protocols have been developed for fitting X-ray data to cryo EM reconstructions (e.g. Wikoff et al., 1994[link]; Che et al., 1998[link]; Volkmann & Hanein, 1999[link]; Wriggers et al., 1999[link]). First, because magnification in an electron microscope can vary, the absolute magnification of the reconstruction to within a few per cent must be established. In addition, the relative scale factor for the density must be calculated. Determination of absolute magnification and relative scaling may be accomplished by several means: (1) comparing the EM map with clear features in the X-ray structure of an individual component (Stewart et al., 1993[link]), (2) using radial density profile information derived from scattering experiments (e.g. Cheng et al., 1995[link]), or (3) using the X-ray structure of an entire assembly when this is available (e.g. Speir et al., 1995[link]). When a single component is used for scaling, it is necessary to refine the scale as the proper position and orientation of that component become better determined. Next, the resolution of the density distribution of the reconstruction must be matched to that in the X-ray structure. For example, fitting of the high-resolution model of the adenovirus hexon to the EM density map of the virus itself was accomplished by convoluting the X-ray structure with the point-spread function for the EM reconstruction (Stewart et al., 1993[link]). An alternative procedure is simply to normalize the EM map so that it has the same range of density values as the corresponding X-ray map (e.g. Luo et al., 1993[link]; Wikoff et al., 1994[link]; Ilag et al., 1995[link]). A maximum-entropy approach (Skoglund et al., 1996[link]) was also used to treat CTF effects to improve the correspondence between the adenovirus hexon X-ray structure and the corresponding density in the EM map (Stewart et al., 1993[link]).

The next step in the modelling procedure is to fit the X-ray structure interactively to the cryo EM density using a display program such as O (Jones et al., 1991[link]), at which point a subjective estimate of the quality of the fit can be made. One criterion for the quality of the fit is whether the hand of the structure can be identified. Because 3D density maps are generated from projected views of the structure, an arbitrary hand will emerge during refinement unless explicit steps are taken to determine the absolute hand. Thus, in the absence of relatively high resolution data (which, for example, would reveal the hand of features like α-helices), the absolute hand must be determined from other types of data such as shadowing (e.g. Belnap et al., 1996[link]) or comparison of the orientations of the same particle imaged at different tilt angles (e.g. Finch, 1972[link]; Belnap et al., 1997[link]). The match with X-ray data (of known hand) serves as an unambiguous determination of hand for the EM map. At this stage of fitting it is important to determine whether the X-ray model should be fitted as a single rigid body (e.g. Stewart et al., 1993[link]) or as two or more partly independent domains (e.g. Grimes et al., 1997[link]; Che et al., 1998[link]; Volkmann & Hanein, 1999[link]).

The quality and uniqueness of the X-ray/EM map fit is then assessed, often simply by calculating an R factor between the two maps. It may be necessary first to mask out density that is not part of the structure being fitted by the X-ray data (e.g. Stewart et al., 1993[link]; Liu et al., 1994[link]; Cheng et al., 1995[link]). The uniqueness of the fit is then tested by rotating and shifting the X-ray model in the EM map and noting changes in the R factor (e.g. Che et al., 1998[link]). An objective fitting procedure, either in reciprocal or real space, is necessary for refining and checking the uniqueness of the result of the interactive fitting. The program X-PLOR (Brünger et al., 1987[link]) can be used for reciprocal-space refinement after the two maps are modified to avoid ripple and edge effects due to masking and differences in contrast (e.g. Wikoff et al., 1994[link]; Grimes et al., 1997[link]; Hewat et al., 1997[link]). Real-space refinement has been performed using other programs (e.g. Volkmann & Hanein, 1999[link]; Wriggers et al., 1999[link]). A few examples include fitting the Sindbis capsid protein to the Ross River virion map (Cheng et al., 1995[link]), fitting the VP7 viral protein into the bluetongue virus core map (Grimes et al., 1997[link]), fitting the Ncd motor domain to microtubules (Wriggers et al., 1999[link]), and fitting of two separate macromolecules, the myosin S1 subfragment and the N-terminal domain of human T-fimbrin, to reconstructions of complexes between these molecules and actin filaments (Volkmann & Hanein, 1999[link]). Comparison between results of real- and reciprocal-space fitting can prove informative. For example, reciprocal-space fitting is usually not constrained to avoid interpenetration of the densities, an issue more easily addressed in real-space fitting. If the two approaches yield different fits, it may be necessary to consider conformation changes between the X-ray and the EM structure. This type of analysis is best performed with quantitative model-fitting routines such as those currently being developed (e.g. Volkmann & Hanein, 1999[link]; Wriggers et al., 1999[link]).

References

First citation Wikoff, W. R., Wang, G., Parrish, C. R., Cheng, R. H., Strassheim, M. L., Baker, T. S. & Rossmann, M. G. (1994). The structure of a neutralized virus: canine parvovirus complexed with neutralizing antibody fragment. Structure, 2, 595–607.Google Scholar
First citation Agrawal, R. K., Penczek, P., Grassucci, R. A., Li, Y., Leith, A., Nierhaus, K. H. & Frank, J. (1996). Direct visualization of A-, P-, and E-site transfer RNAs in the Escherichia coli ribosome. Science, 271, 1000–1002.Google Scholar
First citation Baker, T. S. & Johnson, J. E. (1996). Low resolution meets high: towards a resolution continuum from cells to atoms. Curr. Opin. Struct. Biol. 6, 585–594.Google Scholar
First citation Baker, T. S., Newcomb, W. W., Booy, F. P., Brown, J. C. & Steven, A. C. (1990). Three-dimensional structures of maturable and abortive capsids of equine herpesvirus 1 from cryoelectron microscopy. J. Virol. 64, 563–573.Google Scholar
First citation Baker, T. S., Newcomb, W. W., Olson, N. H., Cowsert, L. M., Olson, C. & Brown, J. C. (1991). Structures of bovine and human papilloma viruses: analysis by cryoelectron microscopy and three-dimensional image reconstruction. Biophys. J. 60, 1445–1456.Google Scholar
First citation Belnap, D. M., Olson, N. H. & Baker, T. S. (1997). A method for establishing the handedness of biological macromolecules. J. Struct. Biol. 120, 44–51.Google Scholar
First citation Belnap, D. M., Olson, N. H., Cladel, N. M., Newcomb, W. W., Brown, J. C., Kreider, J. W., Christensen, N. D. & Baker, T. S. (1996). Conserved features in papillomavirus and polyomavirus capsids. J. Mol. Biol. 259, 249–263.Google Scholar
First citation Brünger, A. T., Kuriyan, J. & Karplus, M. (1987). Crystallographic R factor refinement by molecular dynamics. Science, 235, 458–460.Google Scholar
First citation Böttcher, B., Kiselev, N. A., Stel'mashchuk, V. Y., Perevozchikova, N. A., Borisov, A. V. & Crowther, R. A. (1997). Three-dimensional structure of infectious bursal disease virus determined by electron cryomicroscopy. J. Virol. 71, 325–330.Google Scholar
First citation Böttcher, B., Tsuji, N., Takahashi, H., Dyson, M. R., Zhao, S., Crowther, R. A. & Murray, K. (1998). Peptides that block hepatitis B virus assembly: analysis by cryomicroscopy, mutagenesis and transfection. EMBO J. 17, 6839–6845.Google Scholar
First citation Böttcher, B., Wynne, S. A. & Crowther, R. A. (1997). Determination of the fold of the core protein of hepatitis B virus by electron microscopy. Nature (London), 386, 88–91.Google Scholar
First citation Che, Z., Olson, N. H., Leippe, D., Lee, W.-M., Mosser, A. G., Rueckert, R. R., Baker, T. S. & Smith, T. J. (1998). Antibody-mediated neutralization of human rhinovirus 14 explored by means of cryoelectron microscopy and X-ray crystallography of virus–Fab complexes. J. Virol. 72, 4610–4622.Google Scholar
First citation Cheng, R. H., Kuhn, R. J., Olson, N. H., Rossmann, M. G., Choi, H.-K., Smith, T. J. & Baker, T. S. (1995). Nucleocapsid and glycoprotein organization in an enveloped virus. Cell, 80, 621–630.Google Scholar
First citation Cheng, R. H., Reddy, V. S., Olson, N. H., Fisher, A. J., Baker, T. S. & Johnson, J. E. (1994). Functional implications of quasi-equivalence in a T = 3 icosahedral animal virus established by cryo-electron microscopy and X-ray crystallography. Structure, 2, 271–282.Google Scholar
First citation Conway, J. F., Cheng, N., Zlotnick, A., Stahl, S. J., Wingfield, P. T., Belnap, D. M., Kanngiesser, U., Noah, M. & Steven, A. C. (1998). Hepatitis B virus capsid: localization of the putative immunodominant loop (residues 78 to 83) on the capsid surface, and implications for the distinction between c and e-antigens. J. Mol. Biol. 279, 1111–1121.Google Scholar
First citation Conway, J. F., Trus, B. L., Booy, F. P., Newcomb, W. W., Brown, J. C. & Steven, A. C. (1996). Visualization of three-dimensional density maps reconstructed from cryoelectron micrographs of viral capsids. J. Struct. Biol. 116, 200–208.Google Scholar
First citation Dryden, K. A., Wang, G., Yeager, M., Nibert, M. L., Coombs, K. M., Furlong, D. B., Fields, B. N. & Baker, T. S. (1993). Early steps in reovirus infection are associated with dramatic changes in supramolecular structure and protein conformation: analysis of virions and subviral particles by cryoelectron microscopy and image reconstruction. J. Cell Biol. 122, 1023–1041.Google Scholar
First citation Finch, J. T. (1972). The hand of the helix of tobacco mosaic virus. J. Mol. Biol. 66, 291–294.Google Scholar
First citation Frank, J., Heagle, A. B. & Agrawal, R. K. (1999). Animation of the dynamical events of the elongation cycle based on cryoelectron microscopy of functional complexes of the ribosome. J. Struct. Biol. 128, 15–18.Google Scholar
First citation Fuller, S. D., Berriman, J. A., Butcher, S. J. & Gowen, B. E. (1995). Low pH induces swiveling of the glycoprotein heterodimers in the Semiliki forest virus spike complex. Cell, 81, 715–725.Google Scholar
First citation Gabashvili, I. S., Agrawal, R. K., Spahn, C. M. T., Grassucci, R. A., Svergun, D. I., Frank, J. & Penczek, P. (2000). Solution structure of the E. coli 70S ribosome at 11.5 Å resolution. Cell, 100, 537–549.Google Scholar
First citation Grimes, J. M., Jakana, J., Ghosh, M., Basak, A. K., Roy, P., Chiu, W., Stuart, D. I. & Prasad, B. V. V. (1997). An atomic model of the outer layer of the bluetongue virus core derived from X-ray crystallography and electron cryomicroscopy. Structure, 5, 885–893.Google Scholar
First citation Heel, M. van, Harauz, G. & Orlova, E. V. (1996). A new generation of the IMAGIC image processing system. J. Struct. Biol. 116, 17–24.Google Scholar
First citation Henderson, R., Baldwin, J. M., Ceska, T. A., Zemlin, F., Beckmann, E. & Downing, K. H. (1990). Model for the structure of bacteriorhodopsin based on high-resolution electron cryo-microscopy. J. Mol. Biol. 213, 899–929.Google Scholar
First citation Hewat, E. A., Verdaguer, N., Fita, I., Blakemore, W., Brookes, S., King, A., Newman, J., Domingo, E., Mateu, M. G. & Stuart, D. I. (1997). Structure of the complex of an Fab fragment of a neutralizing antibody with foot-and-mouth disease virus: positioning of a highly mobile antigenic group. EMBO J. 16, 1492–1500.Google Scholar
First citation Hirose, K., Amos, W. B., Lockhart, A., Cross, R. A. & Amos, L. A. (1997). Three-dimensional cryoelectron microscopy of 16-protofilament microtubules: structure, polarity, and interaction with motor proteins. J. Struct. Biol. 118, 140–148.Google Scholar
First citation Hoenger, A. & Milligan, R. A. (1997). Motor domains of kinesin and ncd interact with microtubule protofilaments with the same binding geometry. J. Mol. Biol. 265, 553–564.Google Scholar
First citation Ilag, L. L., Olson, N. H., Dokland, T., Music, C. L., Cheng, R. H., Bowen, Z., McKenna, R., Rossmann, M. G., Baker, T. S. & Incardona, N. L. (1995). DNA packaging intermediates of bacteriophage φ X174. Structure, 3, 353–363.Google Scholar
First citation Jones, T. A., Zou, J.-Y., Cowan, S. W. & Kjeldgaard, M. (1991). Improved methods for building protein models in electron density maps and the location of errors in these models. Acta Cryst. A47, 110–119.Google Scholar
First citation Kolodziej, S. J., Klueppelberg, H. U., Nolasco, N., Ehses, W., Strickland, D. K. & Stoops, J. K. (1998). Three-dimensional structure of the human plasmin α2-macroglobulin complex. J. Struct. Biol. 123, 124–133.Google Scholar
First citation Lawton, J. A., Estes, M. K. & Prasad, B. V. V. (1997). Three-dimensional visualization of mRNA release from actively transcribing rotavirus particles. Nature Struct. Biol. 4, 118–121.Google Scholar
First citation Liu, H., Smith, T. J., Lee, W.-M., Mosser, A. G., Rueckert, R. R., Olson, N. H., Cheng, R. H. & Baker, T. S. (1994). Structure determination of an Fab fragment that neutralizes human rhinovirus 14 and analysis of the Fab–virus complex. J. Mol. Biol. 240, 127–137.Google Scholar
First citation Luo, C., Butcher, S. & Bamford, D. H. (1993). Isolation of a phospholipid-free protein shell of bacteriophage PRD1, an Escherichia coli virus with an internal membrane. Virology, 194, 564–569.Google Scholar
First citation Metoz, F., Arnal, I. & Wade, R. H. (1997). Tomography without tilt: three-dimensional imaging of microtubule/motor complexes. J. Struct. Biol. 118, 159–168.Google Scholar
First citation Nogales, E., Wolf, S. G. & Downing, K. H. (1997). Visualizing the secondary structure of tubulin: three-dimensional map at 4 Å. J. Struct. Biol. 118, 119–127.Google Scholar
First citation Sharma, M. R., Penczek, P., Grassucci, R., Xin, H.-B., Fleisher, S. & Wagenknecht, T. (1998). Cryoelectron microscopy and image analysis of the cardiac ryanodine receptor. J. Biol. Chem. 273, 18429–18434.Google Scholar
First citation Sheehan, B., Fuller, S. D., Pique, M. E. & Yeager, M. (1996). AVS software for visualization in molecular microscopy. J. Struct. Biol. 116, 99–105.Google Scholar
First citation Skoglund, U., Öfverstedt, L.-G., Burnett, R. M. & Bricogne, G. (1996). Maximum-entropy three-dimensional reconstruction with deconvolution of the contrast transfer function: a test application with adenovirus. J. Struct. Biol. 117, 173–188.Google Scholar
First citation Speir, J. A., Munshi, S., Wang, G., Baker, T. S. & Johnson, J. E. (1995). Structures of the native and swollen forms of cowpea chlorotic mottle virus determined by X-ray crystallography and cryo-electron microscopy. Structure, 3, 63–78.Google Scholar
First citation Spencer, S. M., Sgro, J.-Y., Dryden, K. A., Baker, T. S. & Nibert, M. L. (1997). IRIS explorer software for radial-depth cueing reovirus particles and other macromolecular structures determined by cryoelectron microscopy and image reconstruction. J. Struct. Biol. 120, 11–21.Google Scholar
First citation Stewart, P. L., Chiu, C. Y., Huang, S., Muir, T., Zhao, Y., Chait, B., Mathias, P. & Nemerow, G. R. (1997). Cryo-EM visualization of an exposed RGD epitope on adenovirus that escapes antibody neutralization. EMBO J. 16, 1189–1198.Google Scholar
First citation Stewart, P. L., Fuller, S. D. & Burnett, R. M. (1993). Difference imaging of adenovirus: bridging the resolution gap between X-ray crystallography and electron microscopy. EMBO J. 12, 2589–2599.Google Scholar
First citation Taveau, J.-C. (1996). Presentation of the SIGMA software: software of imagery and graphics for molecular architecture. J. Struct. Biol. 116, 223–229.Google Scholar
First citation Vigers, G. P. A., Crowther, R. A. & Pearse, B. M. F. (1986). Three-dimensional structure of clathrin cages in ice. EMBO J. 5, 529–534.Google Scholar
First citation Volkmann, N. & Hanein, D. (1999). Quantitative fitting of atomic models into observed densities derived by electron microscopy. J. Struct. Biol. 125, 176–184.Google Scholar
First citation Vénien-Bryan, C. & Fuller, S. D. (1994). The organization of the spike complex of Semliki forest virus. J. Mol. Biol. 236, 572–583.Google Scholar
First citation Winkler, H., Reedy, M. C., Reedy, M. K., Tregear, R. & Taylor, K. A. (1996). Three-dimensional structure of nucleotide-bearing crossbridges in situ: oblique section reconstruction of insect flight muscle in AMPPNP at 23°C. J. Mol. Biol. 264, 302–322.Google Scholar
First citation Wriggers, W., Milligan, R. A. & McCammon, J. A. (1999). Situs: a package for docking crystal structures into low-resolution maps from electron microscopy. J. Struct. Biol. 125, 185–195.Google Scholar
First citation Yeager, M., Berriman, J. A., Baker, T. S. & Bellamy, A. R. (1994). Three-dimensional structure of the rotavirus hemagglutinin VP4 by cryo-electron microscopy and difference map analysis. EMBO J. 13, 1011–1018.Google Scholar
First citation Zhou, Z. H., Chen, D. H., Jakana, J., Rixon, F. J. & Chiu, W. (1999). Visualization of tegument–capsid interactions and DNA in intact herpes simplex virus type 1 virions. J. Virol. 73, 3210–3218.Google Scholar
First citation Zhou, Z. H., Macnab, S. J., Jakana, J., Scott, L. R., Chiu, W. & Rixon, F. J. (1998). Identification of the sites of interaction between the scaffold and outer shell in herpes simplex virus-1 capsids by difference electron imaging. Proc. Natl Acad. Sci. USA, 95, 2778–2783.Google Scholar








































to end of page
to top of page