Temperature-induced DNA density transition in phage λ capsid revealed with contrast-matching SANS
Edited by Uri Raviv, Hebrew University of Jerusalem, Jerusalem, Israel; received December 4, 2022; accepted September 25, 2023 by Editorial Board Member Michael F. Summers
Significance
This work explains the structural origin of the temperature-dependent DNA density transition in bacteriophage λ capsid, occurring close to the physiological temperature favorable for infection (37 °C, human body temperature). Using small-angle neutron scattering, with contrast-matched scattering contribution from viral capsid proteins, we unveiled two coexisting DNA phases in a capsid—a hexagonally ordered high-density shell-DNA phase in the capsid periphery and a low-density, less-ordered DNA phase in the core. At the transition temperature, a density and volume transition occurs in the core-DNA, resulting in lower density and reduced packing defects. This yields increased mobility of the core-DNA phase, facilitating rapid DNA ejection events from phage into a host bacterial cell.
Abstract
Structural details of a genome packaged in a viral capsid are essential for understanding how the structural arrangement of a viral genome in a capsid controls its release dynamics during infection, which critically affects viral replication. We previously found a temperature-induced, solid-like to fluid-like mechanical transition of packaged λ-genome that leads to rapid DNA ejection. However, an understanding of the structural origin of this transition was lacking. Here, we use small-angle neutron scattering (SANS) to reveal the scattering form factor of dsDNA packaged in phage λ capsid by contrast matching the scattering signal from the viral capsid with deuterated buffer. We used small-angle X-ray scattering and cryoelectron microscopy reconstructions to determine the initial structural input parameters for intracapsid DNA, which allows accurate modeling of our SANS data. As result, we show a temperature-dependent density transition of intracapsid DNA occurring between two coexisting phases—a hexagonally ordered high-density DNA phase in the capsid periphery and a low-density, less-ordered DNA phase in the core. As the temperature is increased from 20 °C to 40 °C, we found that the core-DNA phase undergoes a density and volume transition close to the physiological temperature of infection (~37 °C). The transition yields a lower energy state of DNA in the capsid core due to lower density and reduced packing defects. This increases DNA mobility, which is required to initiate rapid genome ejection from the virus capsid into a host cell, causing infection. These data reconcile our earlier findings of mechanical DNA transition in phage.
Sign up for PNAS alerts.
Get alerts for new articles, or get an alert when an article is cited.
The mechanisms by which viruses deliver their genomes into host cells remain poorly understood. Yet, providing a detailed morphological description and classification of viruses and their receptors is essential for studies of traits influencing virus–host interaction (1), viral replication pathways, as well as for the design of antiviral treatments that inhibit genome release from the capsid and block infection (2). Recent findings of pressure-driven viral double-stranded (ds) DNA ejection into cells from phage (3, 4), archaeal viruses (5), and herpesviruses (6) have raised strong interest in understanding the role that intracapsid dsDNA structure plays in genome ejection and packaging during viral replication (7). Intracapsid genome pressure is generated by packaging micrometer-long dsDNA into a nanometer-size capsid by an ATP-driven (adenosine triphosphate) motor complex on the capsid, which is the strongest molecular motor known (8). In phage λ capsid, DNA is tightly packaged with DNA–DNA surface separation of only 7 Å (9, 10). This leads to high interstrand repulsive forces that, along with DNA bending stress, result in an internal genome pressure of tens of atmospheres (atm = 101.325 kPa) (3, 11, 12); e.g., for wild-type (wt) DNA length phage λ, we measured DNA capsid pressure of ~25 atm (3). This pressure powers DNA ejection into a bacterial cell. Similarly, we found that DNA pressure of ~20 atm drives genome ejection from human Herpes simplex virus type 1 (HSV-1) into a cell nucleus (6). However, the small interstrand distance also generates electrostatic sliding friction between neighboring DNA strands (13–15), which limits a packaged genome’s mobility (or fluidity) and restricts rapid DNA ejection from the capsid during cell infection (7).
Using atomic force microscopy (AFM), we have previously shown that restricted intracapsid DNA mobility can be characterized by a DNA-filled phage λ capsid’s solid-like mechanical response to nanoindentation (15, 16). At the same time, DNA packaged in a λ-capsid undergoes a mechanical transition from a solid-like to a fluid-like state triggered by increasing temperature. Fig. 1A shows AFM-determined spring constant in N/m for DNA-filled phage λ capsids measured as a function of temperature. Since spring constant of empty capsids remains unaffected by temperature variation (measured separately in ref. 15), the mechanical transition is reflecting an abrupt stiffness change in intracapsid DNA. The increase in temperature affects the DNA bending stress and its packing defects, leading to the observed genome transition behavior (Discussion) (17). We independently confirmed this intracapsid DNA transition with isothermal titration calorimetry (ITC) measurements; the enthalpy of DNA ejection from the phage displayed an abrupt internal energy transition between 31 °C and 35 °C, see Fig. 1B (15). Below the transition temperature (where DNA is in a solid-like state), we found that the ejection dynamics for a phage λ population are slowed down by an order of magnitude, or ejection is completely stalled, resulting in noninfectious particles. For a phage population infecting a bacterial cell, this translates into desynchronized viral DNA ejection events, where some phage particles eject their DNA instantly while others have a significant time delay (7). However, above the transition temperature (with DNA in a fluid-like state), DNA ejection events from a phage population are synchronized, occurring within seconds (7), which facilitates viral infectivity (15). Synchronization or desynchronization of DNA ejection events from a phage population infecting the same cell alters the course of infection. Competition (or the lack of it) in replication and transcription dynamics between viral genomes entering the infected cell influences the cell’s “decision” between lytic or lysogenic (dormant) replication states, as was found for phage λ replication (18, 19). Interestingly, the intracapsid DNA transition temperature is close to that of the physiological temperature of infection, ~37 °C (human body temperature), which suggests evolutionary physical adaptation of phage λ to its host environment. Thus, the mechanical state of intracapsid DNA is central to viral genome ejection- and replication dynamics, which influence infectivity (7, 18, 19). It is evident that intracapsid DNA mechanics, structure, and packing density are directly related. However, structural evidence of the temperature-induced intracapsid DNA transition in phage was lacking, which prompted this study.
Fig. 1.

We use small-angle neutron scattering (SANS) to investigate temperature-induced change in DNA density and its structural arrangement (through radial density distribution) in phage λ capsid. SANS uses a contrast matching method, which is based on the deuteration of the aqueous buffer solvent, i.e., the exchange of fraction H2O to D2O (20–22). This highlights or hides from the neutron beam selected components of the biological structure (e.g., protein capsid or DNA) depending on the percentage of deuteration of the solvent (23–25); i.e., each component has a characteristic scattering length density (SLD) value, which can be matched by adjusting the SLD of the solvent. Thus, one can use contrast matching to reach the same level of coherent scattering between the buffer and the protein capsid, providing a unique opportunity to study scattering from intracapsid DNA structure alone. The contrast matching SANS technique has been previously used to provide details on size and spatial arrangement of viral constituent components in influenza B virus (26) and phage MS2 (27). We use this powerful approach to analyze transition behavior in the virally packaged genome. Typically, a core–shell model is used for SANS data analysis to describe viral capsid (shell) and its packaged genome (core). In this work, however, we use a core–shell model to explore variation in the density and volume of two coexistent DNA phases within a λ-capsid, observed as variation in the SANS scattering intensity, reflecting intracapsid DNA form factor with temperature. The obtained results validate and explain the nature of a solid-to-fluid-like mechanical DNA transition in phage λ capsid, which was independently observed with AFM and ITC (15); the results also clarify previous SAXS (Small-Angle X-ray Scattering) data analysis (28), which could not explain the observed DNA transition behavior. This was due to the fact that SAXS data alone (without contrast variation) (28) reflects only the structure factor of the hexagonally ordered shell-DNA phase, where DNA density is unaffected by temperature, as we demonstrate below.
To summarize, in this work, we used contrast matching SANS to study the DNA structure packaged in phage λ capsid as a function of temperature. Using SANS data model analysis, combined with our previous cryo-EM (Cryoelectron microscopy) and SAXS structural data for phage λ under the same conditions, we demonstrate a coexistence of two DNA phases—a high-density, hexagonally ordered DNA phase in the periphery of the capsid and a low-density, less-ordered (or disordered) DNA phase in the capsid core. By observing how temperature affected the packing density and volume of both DNA phases, we revealed a correlation between the density change in the core-DNA phase with our earlier AFM/ITC observation of the solid-to-fluid-like mechanical transition of the encapsidated λ genome. The contrast-matching SANS approach presented in this work will help to elucidate universal physico-chemical behavior of tightly packaged nucleic acid condensates in capsids of different viral systems and provide a further understanding of factors influencing viral replication dynamics (7, 18, 19).
Results and Discussion
First, using our earlier cryo-EM single-particle reconstruction and SAXS data for phage λ, we deduce structural information that serves as starting parameters to accurately model SANS data for DNA-filled phage λ capsids (Cryo-EM and SAXS Structural Analysis of DNA Packaged in Phage λ Capsid). We also highlight the limitations of cryo-EM and SAXS techniques in observing the structural changes in the packaged λ-genome (7, 15, 17). These limitations are resolved with the SANS approach, where the capsid protein contribution is contrast-matched, revealing temperature-induced changes in the intracapsid DNA structure (Intracapsid DNA Density Transition Revealed with SANS Contrast Matching), (Fig. 2).
Fig. 2.

Cryo-EM and SAXS Structural Analysis of DNA Packaged in Phage λ Capsid.
Fig. 2 A and B show the central cross sections of the 3D cryo-EM icosahedral reconstruction of phage λ-capsid filled with wt λ-DNA [length 48,500 base pairs (bp)] as well as empty λ-capsid at 20 °C in 10 mmol/L (mM) Mg-Tris buffer. The cryo-EM structure reveals that the entire capsid volume is filled with DNA (red color), extending all the way to the center of the capsid. To better visualize the distribution of DNA within a capsid, we calculated the radial average density profile from the cryo-EM 3D electron density maps using Chimera software, displayed in Fig. 2C. The density profiles on the y axis are shown in arbitrary units relative to averaged background density, but the relevant structural information is in the location of the maximums reflecting the DNA density distribution within the capsid. The density profile for the empty capsid (black line) shows a flat line in the void volume, displaying only the location of the capsid wall. EM cross-sections show identical capsid radii for DNA-filled and empty capsids, as observed from overlapping capsid wall density peak positions at ~300 Å (Fig. 2C). The increase in the DNA density profile from the capsid wall toward the capsid center is attributed to an artifact resulting from icosahedral averaging for single-particle reconstruction, as we verified with SANS data analysis below. [The effect of DNA density increase toward the capsid center is due to the Fourier-based reconstruction strategies typically used for EM data analysis. The interpolation strategy used during Fourier reconstruction can produce a radial falloff in the real-space image due to the entire particle volume being filled with DNA matter, scattering the electron beam. For comparison, an empty capsid is missing DNA matter in the capsid interior (besides buffer background) and the reconstruction process of image averaging does not contribute to the radial density from the inner capsid wall to its center. Other authors have observed the same trend in radial DNA distribution using similar reconstruction algorithms, e.g., see ref. 29 showing DNA distribution in phage P22. Also, it can be noted that there is a negative density in the radial distribution from the cryo-EM map on both sides of the intensity maxima corresponding to the protein capsid wall. This is attributed to imperfections in estimating and correcting the objective aberrations introduced during data collection along with intensity normalization procedure. Thus, it is the relative and not the absolute intensity values that are meaningful.] The density profile for the DNA-filled capsid shows, however, how DNA is structurally arranged from the inner capsid wall toward the center. Along the capsid wall, there are well-ordered, multiple concentric DNA layers (∼7 to 8 layers). The layers are evenly spaced, suggesting that DNA has adopted an ordered liquid crystalline state (shown as dashed grey vertical lines). The spacings between the DNA layers were determined by fitting a Lorentzian function to each of the local maxima on the DNA density curve versus the distance from the capsid center (Fig. 2C). Those maxima provide the location of ordered DNA layers in the capsid. The interlayer distance displays periodicity for DNA in the periphery of the capsid (9). However, the DNA density profile in Fig. 2C shows that DNA layer periodicity is lost toward the capsid center, suggesting that intracapsid DNA structure coexists in two phases—ordered in the capsid periphery and less ordered (or disordered) in the center. The cutoff radius for the less-ordered core-DNA phase from the capsid center, rcore, can be determined from the density profile plot (rcore ∼ 95 Å), shown as a red dashed vertical line (Fig. 2C). [We locate the position for each DNA density maximum in Fig. 2C using a Lorentzian function. We previously determined the average interlayer distance of aH = 22.7 Å from cryo-EM map for phage λ (9). Then, we calculated the interlayer distance towards the capsid center between the density peaks within a deviation value of 10% from its average value. Based on this criterion, we defined the cutoff radius for the disordered DNA phase. The oscillations beyond the cutoff radius are outside of our criterion].
Complementary to the cryo-EM data, we performed SAXS analysis of DNA-filled and empty phage λ particles in the temperature range 10 °C to 40 °C in TM-buffer (10 mM Tris+10 mM MgCl2); see Fig. 3. Small-angle scattering methods (SAXS and SANS) are powerful experimental techniques for the structural analysis of biological matter due to their range of resolution, from some tens of Ångstroms to a few hundreds of nanometers. Ordered structures, such as icosahedral capsids or packaged DNA ordered in the periphery of the capsid, produce Bragg scattering peaks. Fig. 3A shows SAXS scattering intensity versus magnitude of the scattering vector q (called wave number) for empty λ-capsid (green curve) and wt DNA-filled phage λ (red curve). The interpretation of the structure reflected by the scattering intensity pattern depends on the value of the wave number q, which is inversely proportional to the length scale L of the analyzed structure, i.e., . Thus, the scattering signal from DNA-filled phage λ in the low q-region ~0.006 to 0.06 Å−1 provides scattering information at the large-length scales (tens to hundreds of nm), corresponding to the whole viral particle and reflecting structures of both capsid and global packaged DNA arrangement inside (called the “form factor” in the scattering analysis). On the contrary, at high q-region ≥ 0.1 Å−1, Bragg peak reflects the coherent scattering at short-length scales (tens of Angstroms) from the averaged local ordering of dsDNA helices packaged in a capsid (DNA “structure factor” in the scattering analysis), providing average distance between ordered DNA centers, ds (10, 30). This is illustrated in Fig. 3A. For comparison, the scattering profile from an empty λ-capsid is missing the DNA Bragg peak in the high q-region, while the capsid structural features are visible as ripples in the low q-region (Fig. 3A). SAXS data for DNA-filled λ-capsid can be interpreted by fitting the scattering intensity versus q. Our fit combines two models that separately describe the data at low and high q-regions (black line), since they reflect different structural features of the packaged DNA, as explained above. At low q, a simple spherical model (approximating the icosahedral capsid with a solid sphere, not differentiating between the protein shell and the packaged DNA) describes the scattering profile, providing information on the capsid size. At high q, the broad peak (BP) model is used to fit the observed Bragg peak, which can then be used to accurately locate the position of peak maximum, which reflects the interaxial distance between ordered DNA strands in the capsid (31, 32). The BP model consists of a power law term (usually referred to as linear background) and a Lorentzian term (33, 34). As mentioned above, previous X-ray scattering analysis of bulk DNA phases condensed by an osmolyte at packing density corresponding to that in phage λ (10) has demonstrated hexagonal DNA ordering (10, 30). The interaxial distance of <∼30 Å corresponds to a hexagonal order (35–37) consistent with the packaged DNA structure found in phage λ capsid. The disordered DNA in the core of the capsid should not contribute to the SAXS scattering signal at high q (due to the incoherent scattering, the disordered structures do not produce a Bragg peak). The characteristic Bragg peak for the hexagonally ordered DNA structure in the capsid periphery (10, 30) in Fig. 3A displays a maximum at qBragg ~0.26 Å−1. Using the relationship for a hexagonal lattice , we obtained ds-value of 27.6 Å, which remained constant in the entire temperature range of 10 to 40 °C, see Fig. 3B. This SAXS-obtained interaxial distance value corresponds to the interlayer distance, aH, between radially ordered DNA layers in the cryo-EM map cross-section above, through a relationship for a hexagonally ordered DNA lattice, connecting SAXS and cryo-EM data together (6, 9).
Fig. 3.

If we assume that full-length λ-DNA is hexagonally ordered through the capsid volume with interaxial DNA–DNA distance ds = 27.6 Å, then a volume unoccupied by DNA would exist toward the center of the capsid (12, 38). However, as observed by cryo-EM reconstruction in Fig. 2, DNA is in fact dispersed throughout the entire λ-capsid volume without a void in the center. As described above, the apparent interlayer ordering in the cryo-EM cross section disappears toward the center of the capsid, suggesting that DNA is in a disordered state in the capsid core. Similar dsDNA distributions within the capsids were also observed for other viruses (39–43). As shown in Fig. 3B, the interaxial distance ds for ordered DNA in the capsid periphery is constant in the entire temperature range (10 to 40 °C), suggesting that its structure is unaffected (SAXS raw data and model fits at all temperatures are shown in SI Appendix, Fig. S1). At the same time, we earlier observed with ITC and AFM an abrupt transition in intracapsid DNA properties at ~32 °C, see Fig. 1 (16). We hypothesize that a temperature-induced mechanical and density DNA transition occurs in the core of the capsid, rather than in the capsid periphery, affecting the boundary between coexisting high-density (in the capsid periphery) and low-density (in the capsid center) DNA phases. We use a spherical core–shell model based on cryo-EM and SAXS data (Figs. 2 and 3) to fit the SANS data for DNA structure packaged in phage λ capsid as a function of temperature, where contribution from the capsid protein is contrast matched by a deuterated solvent. This provides a stringent test of our hypothesis. Here, structural analysis data are used to describe packaged DNA structure in a viral capsid as a coexisting two-phase system of different DNA densities. Alternatively, one-phase, hexagonally ordered DNA in the capsid [as was proposed in the earlier works (10, 44)] cannot explain the observed DNA transition behavior and is not supported by the cryo-EM density data (9).
Intracapsid DNA Density Transition Revealed with SANS Contrast Matching.
As mentioned above, SAXS provides structural information from the scattering mainly attributed to both protein capsid and DNA, as well as contributions from the cross-correlated scattering terms from all viral components (including phage tail proteins). We are interested in the unique scattering contribution from the global DNA structure confined in the capsid, reflected in the low-q region. However, the scattering intensity contributions from the capsid and the cross-correlation terms do not allow us to fit the SAXS scattering data with a model suggested by cryo-EM data, which separates the scattering signal contribution from the apparent two-phase packaged DNA structural densities from the capsid scattering signal. To resolve this issue, we performed SANS contrast matching experiments, where the scattering contributions from all virus protein components are removed, allowing us to resolve the large-scale structural features of the genome packaged in phage λ capsid [in contrast to the small-scale DNA–DNA spacings determined in the past from SAXS Bragg peak at high q (10)].
SAXS provides a scattering signal from the interaction of X-ray radiation with the electron cloud of atoms, while SANS provides scattering from the short-range interaction of a neutron beam with the nuclei of atoms. Hence, one of the principal advantages of a neutron beam is its selectivity to specific nuclear isotopes, which are not differentiated by X-rays. Of particular interest are the dissimilar values of both coherent and incoherent neutron scattering from hydrogen (H) and deuterium (D). As described above, since biological matter is “bathed” in an aqueous buffer solution containing hydrogen atoms, we can change the SLD of the buffer by replacing the fraction of water with heavy water (deuterium oxide), resulting in different ratios of H/D atoms, allowing us to achieve a large range of SLDs for the buffer solution. Thus, the SLD for both viral capsid (proteins) and DNA can be contrast matched to an aqueous buffer containing different fractions of deuterium, making either of these components “invisible” to the neutron beam (i.e., indistinguishable from buffer SLD and resulting in zero scattering contribution upon buffer signal subtraction), see Fig. 4A. For X-rays, the parallel method of contrast matching is challenging to achieve since contrast depends primarily on the differences in the electron densities of molecules; it can be difficult to obtain sufficient contrast match between biomacromolecules and the surrounding buffer without significantly affecting buffer properties [e.g., by addition of high sucrose fraction (45)].
Fig. 4.

Fig. 4A shows the variation in SLD for the TM-buffer used in this study (Materials and Methods), phage λ proteins (capsid and tail) and λ-DNA versus D2O/H2 O fraction in the buffer solution. SLDs are computed specifically for phage λ, taking into account all amino acids of virus protein components, and for λ-DNA, taking into account the sequence of the 48,500 bp λ-genome. Buffer SLD increases linearly with increasing fraction of D2O (46). As shown in Fig. 4A, the contrast match point for phage λ proteins (termed “capsid” in the figure legend and throughout the text) occurs at 43% volume fraction of D2O/H2O in TM-buffer, where the capsid SLD line intercepts the buffer SLD line. Fig. 4B shows SANS measurement of scattering intensity for empty λ-capsid at contrast match point in 43% D2O/H2O TM-buffer mixture and at full contrast in 100% D2O TM-buffer. Due to the neutron scattering properties of deuterium (47, 48), the SANS scattering signal at 100% D2O full contrast accentuates the structural features of the icosahedral-shaped empty capsid, showing characteristic ripples. On the contrary, at 43% D2O/H2O contrast match, the SANS scattering signal lacks any structural features in the entire q-range, showing essentially a flat line at the level of buffer contribution to the scattering signal (buffer signal contribution has not been subtracted here since it would result in predominantly noise). The observed slight increase in SANS intensities at 100% and 43% D2O/H2O toward the low q-limit suggests small particle aggregation (stable monodisperse capsid suspension should display a plateau in the low q -limit). This is due to the protocol used to empty λ-capsids from DNA, where DNA-filled phage solution is first heated to 78 °C and subsequently cooled to room temperature (49). This procedure destabilizes the portal cap proteins, leading to DNA ejection, while the rest of the capsid structure remains intact (as we demonstrated in ref. 50). The ejected DNA is then digested with DNase I prior to SANS measurement. Heating, however, can lead to partial aggregation of a fraction of broken capsid (50). Regardless of the partial capsid aggregation, this SANS measurement shows a stark difference in the scattering signals between contrast-matched and full-contrast empty phage capsids, illustrating the effectiveness of using the contrast-matching method to visualize packaged DNA structure in DNA-filled phage without scattering contribution from the capsid structure.
Next, we conducted SANS measurements on DNA-filled phage λ with contrast matching of the protein capsid in 43% D2O/H2O TM-buffer solution. This allows us to isolate and explore the structural changes of the encapsidated DNA as temperature is increased from 20 °C to 40 °C [our AFM data showed that the solid-to-fluid-like DNA transition in phage λ occurs at ~32 °C (51, 52)]. We separately confirmed with SAXS that the scattering profile from empty λ-capsids remains unchanged between 20 °C and 40 °C, showing that capsid structure is unaffected (17). Fig. 4C shows a representative scattering intensity profile, I versus wave number q for DNA in phage λ capsids at 20 °C, where scattering contribution from the solvent has been subtracted to further highlight the scattering features (I versus q SANS raw data for DNA in phage λ capsids in 43% D2O/H2O TM-buffer solution at 20 °C, 25 °C, 30 °C, 35 °C, and 40 °C are shown in SI Appendix, Fig. S2). The scattering profile shows a plateau in the low q-limit, confirming that no particle aggregation occurs. SANS data are typically analyzed by fitting the scattering data with a model of the investigated structure (53). Since we are interested in the effect of temperature on the behavior of two coexisting phases of encapsidated DNA (hexagonally ordered shell-DNA and less-ordered core-DNA), we have chosen a spherical core–shell model approximation (33, 34) (Materials and Methods), shown schematically in the inset of Fig. 4C. Core-DNA radius, rcore, is measured from the capsid center, and thickness of shell-DNA is provided by t = rshell – rcore, where rshell is the outer DNA radius confined by the capsid wall, measured from the capsid center, see details in Materials and Methods section. We assume that the outer shell-DNA radius is equal to that of the inner capsid radius. The inner capsid radius was determined by subtracting the λ-capsid wall thickness of 2 nm [obtained by cryo-EM (54)] from the outer capsid radius value obtained by fitting the SAXS data with a sphere model in the low q-region (Fig. 3A). The spherically averaged inner λ-capsid radius is ~29.9 nm and is constant at all temperatures (15, 17). [It can be noted that our SANS data for DNA-filled phage with contrast-matched capsid suggests that the radius of the outer DNA layer is ~1.6 nm smaller than the inner capsid wall radius, likely due to the repulsion between DNA and the negatively charged capsid wall (9, 54)]. This “empty” space between the outer DNA layer and the inner capsid wall was indeed observed with cryo-EM (Fig. 2A). However, the DNA-capsid wall separation distance does not change with temperature, as verified by our SANS data, and for simplicity is omitted in the core–shell model.
Our core–shell SANS model has two fitting variables (or degrees of freedom)—core-DNA radius (rcore) and core-DNA SLD (ρcore). The remaining model parameters in the core–shell fitting model have been derived from the SAXS and cryo-EM data above, see Materials and Methods. Shell-DNA thickness, t, can be calculated for any given value of rcore since the outer shell radius is constant (rshell = 28.3 nm). The initial fitting value for rcore at 20 °C was determined from the cryo-EM DNA density plot versus distance from the capsid center, shown with the red dashed vertical line in Fig. 2D, where DNA layer periodicity is lost (the phage sample was incubated at 20 °C prior to cryo-freezing to collect the EM density map). Note that the initial rcore and ρcore values were determined (see Materials and Methods section) only to start the computational iteration. A representative core–shell model fit to the SANS data for DNA-filled phage λ with contrast matched capsid at 20 °C shows a good agreement between the SANS data and the fitting model (Fig. 4C). (Core–shell model fits to SANS data at all temperatures in the range 20 °C to 40 °C are shown in SI Appendix, Fig. S2). In the low q -region, the DNA structure displays ripples (subsidiary maxima) that reflect the overall shape and structural features of the entire λ-genome packaged in the capsid (with characteristic distance correlation from tens to a few hundreds of nm). At higher q-values, the intensity declines to low levels comparable to the buffer scattering intensity, which results in predominantly noise when the buffer scattering background is subtracted from the sample scattering signal. In SI Appendix, Fig. S3, we plot all spherical core–shell model fits (I versus q) at different temperatures in one plot in a narrow q-range of 0.015 to 0.038 Å−1, focusing on DNA form factor. Since total DNA size is constrained by the capsid wall and is constant with temperature, the change is observed only in the absolute intensity of the form factor peak. SI Appendix, Table S1 summarizes all model fitting parameters and shows goodness of each fit (χ2).
Fig. 5A shows the variation in the core-DNA radius, rcore, with increasing temperature obtained from the core–shell model fit to our SANS data between 20 °C and 40 °C for the DNA-filled phage λ sample with capsid contrast matched at 43% D2O/H2O (as mentioned above, both SLDs, ρcore and rcore, are model fitting parameters varying with temperature). This allows us in turn to calculate the temperature-induced change in the shell-DNA thickness, Fig. 5A. Using values for the core-DNA radius and shell-DNA thickness obtained from the core–shell SANS data model, we computed the variation in volume for the core- and shell-DNA phases with temperature, Fig. 5B. Since DNA is organized in a hexagonal lattice in the shell-DNA phase, the DNA density (bp/nm3) in the shell phase was calculated from , where 0.34 nm is the average distance between each bp and ds = 2.76 nm (17). This yields DNA density in the shell phase of ~0.446 bp/nm3, which remains unchanged in the entire temperature range since the DNA–DNA interaxial distance is constant with temperature, as shown with SAXS in Fig. 3B. This implies that the SLD for shell-DNA (ρshell) also remains constant with temperature, ρshell = 3.68 × 1010 cm−2 at 43% D2O/H2 O (it includes contributions from both DNA and water hydrating the DNA, ρshell = ϕDNA x SLDDNA + ϕD2O/H2O x SLDD2O/H2O, where ϕDNA = VDNA/Vshell and the volume fraction of water in the shell is ϕD2O/H2O = VD2O/H2O/Vshell); see the details in Materials and Methods. Knowing the volume of the shell-DNA phase as well as the constant density of DNA ordered in a hexagonal lattice, we determined the number of bp of DNA that fits into the shell phase, shown as % DNA bp relative to wt λ-DNA length (48,500 bp), versus temperature; see Fig. 5C. Once the DNA fraction that fits in the shell phase is known, the remaining DNA must fit into the core phase region; see Fig. 5C. Hence, the DNA density in the core can be computed using the core-DNA volume values. Core-DNA density variation with temperature is shown in bp/nm3 in Fig. 5D. In summary, our core–shell model SANS data analysis provides a comprehensive description of the changes in DNA packing density and volume distribution between two coexisting DNA phases in a phage capsid, occurring as temperature is increased from 20 °C to 40 °C. [It is important to reiterate that ρshell and ρcore are treated as SLDs of two components (shell-DNA and core-DNA component, respectively) in a finite volume of the capsid, where each component contains both DNA and D2O/H2O at different ratios and not pure DNA alone with an infinite volume. DNA density in the shell-DNA phase component is homogeneous and is constant with temperature due to unchanged ds-value of the ordered DNA phase, which leads to a constant ρshell value with temperature, based on our definition of ρshell above. However, the extent of core-DNA density ordering is unknown, and ρcore values obtained from the model fit do not follow the trend of the core-DNA density variation with temperature [see SI Appendix, Table S1 showing ρcore (T)]. This further confirms that DNA density in the core phase is not homogeneous.
Fig. 5.

Fig. 5A shows that the core-DNA radius increases linearly between 20 °C and 30 °C but then declines as temperature is further increased to 40 °C. The change in shell-DNA thickness shows the opposite behavior since capsid volume is constant and is coupled to changes in core-DNA radius. Shell-DNA thickness shows a decrease between 20 °C and 30 °C followed by an increase between 35 °C and 40 °C. In Fig. 5B, we also calculate volume variation with temperature of core- and shell-DNA phases. It is interesting to note that the volume of the less-ordered core-DNA phase is approximately doubled as the temperature is raised from 20 °C to 30 °C, suggesting that temperature has a significant effect on the extent of DNA disordering in the capsid. This expansion of the less-ordered DNA phase in the capsid core occurs at the expense of the highly ordered DNA phase in the shell (where DNA always remains hexagonally ordered). This is illustrated in Fig. 5C; when temperature is increased from 20 °C to 30 °C, the number of bp in the ordered shell phase is reduced as DNA layers are moved from the shell into the less-ordered core phase. When temperature is further increased from 35 °C to 40 °C, this behavior is reversed. Fig. 5D shows that the DNA density in the core is increasing fivefold as the temperature increases from 20 °C to 30 °C. However, the core density change reverses and begins to show a decline between 35 °C and 40 °C, suggesting a DNA density transition occurring between 30 °C and 35 °C. This temperature interval for the apparent density transition in core-DNA phase correlates directly with the DNA enthalpy transition temperature interval measured with ITC (15) (shown as solid grey area in all graphs in Fig. 5) and with our earlier AFM observation of solid-to-fluid like mechanical DNA transition in the capsid occurring at ~32 °C (15). To illustrate this, we plotted both the AFM data for the spring constant (N/m) of DNA-filled phage λ capsids (red triangles) and the intracapsid DNA density change obtained from the SANS data analysis (Fig. 5D). The spring constant shows an initial increase in DNA stiffness between 20 °C and 30 °C, which is in agreement with increasing core-DNA density, as higher DNA packaging density is expected to yield higher DNA stiffness. Following the DNA transition at ~32 °C, there is an abrupt drop in DNA stiffness followed by a continued decrease in stiffness between 33 °C and 37 °C. The decrease in stiffness correlates with the observed decrease in the core-DNA density in the same temperature range. It should be noted that the SANS data analysis shows that core-DNA density is significantly lower than shell-DNA density in the whole temperature range (e.g., ~2 times lower at the transition temperature of ~32 °C). The opposite trend was observed in the cryo-EM density map (Fig. 2C), where DNA packing density gradually increased toward the capsid center. This discrepancy can be attributed to an artifact in cryo-EM analysis, resulting from icosahedral averaging of EM data. It is worth mentioning that a two-phase coexistence of DNA structure in a capsid with lower DNA density in the center has been theoretically predicted with a continuum approach to the equation of state for encapsidated DNA (55). This model predicts a coexistence between line hexatic (high density) and cholesteric (low density) phases, which translates to a nonhomogeneous DNA distribution inside the capsid, consistent with our SANS and cryo-EM observations. However, in this model (55), the temperature is kept constant and temperature effect on the coexisting DNA phases was not investigated.
The observed changes in the volume and density of the core-DNA can be explained by a temperature effect on DNA–DNA interactions, bending stress, and packing defects (30, 56, 57)—all of which determine tertiary DNA structure in the capsid. The interplay between these parameters with increasing temperature can explain the initial increase in the density and volume of the core-DNA phase (between 20 and 30 °C, with corresponding decrease in the volume of the shell-DNA phase, which retains constant density), the density transition (at 30 to 35 °C), and the decrease in the core-DNA volume and density (between 35 and 40 °C, with shell-DNA volume increasing again). First, we consider temperature effect on DNA–DNA interaction energy. We previously analyzed the effect of temperature on DNA–DNA interaction energy alone, without contribution from DNA bending and packing defects induced by viral capsid confinement (15, 17). We used SAXS to determine the DNA interaxial spacing for bulk DNA arrays condensed in solution by an osmotic stress polymer (polyethylene glycol, PEG), under ionic conditions identical to the SANS measurements in this work (17, 58). We found that DNA–DNA interaxial distance remained constant in the temperature range 5 °C to 50 °C. This indicates that repulsive interactions for linearly packaged DNA are not affected by temperature at a DNA packing density corresponding to that in phage λ capsid. Next, we review the scenario when DNA is confined in phage λ capsid. The intracapsid confinement requires DNA to bend along radii that are energetically unfavorable given the internal λ-capsid radius of ∼30 nm (54) and the dsDNA persistence length of 50 nm (11, 59), which creates bending stress on the packaged genome. (Persistence length defines the stiffness of a polymer, describing the minimum radius of curvature it can adopt by the available thermal energy. Bending it to a smaller radius requires additional work). To relieve the bending stress, helices are packed closer to the capsid wall, decreasing the bending radius and also decreasing the spacing, therefore increasing the interaction energy. At the same time, the repulsive DNA–DNA interactions will push DNA strands as far from each other as possible, filling the entire capsid volume and maximizing the interstrand separations (9). Thus, there is a trade-off between bending and interaction energies. Furthermore, when DNA is bending inside the capsid, the initial correlation between two helices that have slightly different radii of curvature is lost, and the mutual orientation between helices must be reestablished. This leads to packing defects, which are absent for linear packaging of DNA in solution but are necessary here to reestablish a favorable phosphate–phosphate “phasing” of helices to reduce the repulsive interactions due to bending (9, 57, 60). The higher temperature likely hinders optimum correlation between the helices. The spacing remains the same in the hexagonally ordered shell-DNA phase since DNA is simply filling a volume (Fig. 3B), but the interhelical repulsion will increase. This is confirmed by an initial increase in the stiffness of DNA in the capsid measured by AFM, when temperature is increased from 25 to 32 °C (preceding the DNA transition, Figs. 1B and 5D). In parallel with the increasing interstrand repulsions, the increase in temperature will decrease the DNA persistence length (61), leading to less bending stress. If the bending stress decreases and if there is room to expand in the capsid, then spacings would increase and interaction energy decrease. While volume of the capsid is fixed and DNA cannot expand in total volume, it can expand into the core volume of the capsid, where DNA density is significantly lower than in the hexagonally ordered DNA phase in the periphery of the capsid (shell-DNA). SANS data in Fig. 5B (volume of DNA phases) and Fig. 5 C and D (% bp and density of DNA phases) show that this partial DNA expansion occurs; a portion of DNA in the high-density ordered shell phase is forced into the core phase in the capsid with significantly lower packing density. As a result, the volume of the core-DNA is increased while the volume of the shell-DNA is decreased (shell-DNA density always remains constant). It is striking to observe that DNA density in the core phase is increasing by a factor of ~5 (from ~0.05 bp/nm3 to ~0.25 bp/nm3) when temperature is increased from 20 °C to 30 °C, prior to the DNA phase transition.
The number of DNA packing defects, however, also increases with increasing DNA density and volume in the core [packing defects occur more frequently closer to the capsid center due to higher DNA bending curvature resulting in poor alignment between neighboring DNA helices (57)], leading to increasing energy in the core-DNA phase. (As described above, our SANS data does not provide information on the ordering of DNA but only on the size and shape of core- and shell-DNA phases. Hence, partial ordering can occur in the core-DNA phase, which leads to packing defects in both core- and shell-DNA phases). There appears to be a maximum threshold core-DNA density value at ~30 °C before density transition occurs and the core-DNA adopts a new interhelical conformation with packing density decreasing between 35 and 40 °C (Fig. 5B). This reduces the number of packing defects at temperatures >30 °C due to decreased density and volume. The DNA is moved from core phase back to shell-DNA phase, so the volume of the core is decreasing again with a corresponding increase in the shell-DNA volume (Fig. 5B). Thus, as shown in Fig. 5, our core–shell model of contrast matched SANS data for phage λ as a function of temperature (describing the effect of temperature on the two-phase coexistence in intracapsid DNA) confirms a structural DNA transition in the capsid core between 30 °C and 35 °C, which supports the AFM- and ITC-measured solid-to-fluid like transition in the same temperature range. The transition yields a lower energy state of DNA in the capsid core (with lower density and reduced packing defects), increasing its mobility, which is required to initiate rapid DNA ejection from the core of the virus capsid into a host cell (7). As mentioned above, core-DNA is packaged last into the capsid volume during virus assembly (due to high DNA bending stress) and is therefore ejected first during infection. Furthermore, it should be mentioned that the 10 mM Mg2+ buffer concentration used in this study corresponds to the most favorable physiological Mg-ion conditions for phage λ adsorption to Escherichia coli cells and subsequent infection (17). Indeed, we have previously observed that phage λ replication dynamics is significantly increased above the DNA transition temperature of ~32 °C (17).
Concluding Remarks
Structural virology has been instrumental in providing a detailed morphological description and classification of viruses and their receptors (1). The structure of a virally encapsidated genome plays a major role in the processes of virion assembly and viral genome release during infection, as well as for viral genome replication dynamics in a host cell (7, 18, 19, 62). Yet, knowledge of virally packaged genome structures is currently limited. Most of the structural studies on viruses have employed X-ray crystallography (63), cryo-EM (64), and SAXS (65), where focus has been mainly on the viral capsids due to their symmetry, rather than on the encapsidated genome, where the lack of symmetry prevents high-resolution analysis. Furthermore, cryo-EM-obtained structural information on a genome packaged in a capsid is obstructed by artifacts from icosahedral averaging. SAXS has been used to investigate packaged DNA structure in phage (10). However, SAXS data analysis is limited to the small-scale DNA–DNA spacing in a phage capsid, determined from the structure factor for hexagonally ordered DNA reflected by a Bragg peak in high q-region. As explained above, SAXS does not resolve the global DNA structure in the capsid reflected by the form factor in the low q -region since the DNA scattering intensity is obstructed by the high-intensity scattering from the icosahedral capsid structure. We have previously shown with AFM that packaged DNA mobility in phage λ capsid undergoes a transition from a solid-like to a fluid-like state close to physiological temperature of infection (15). This mechanical transition controls the dynamics of viral genome ejection into a host bacterial cell (as we demonstrated with ITC), which is suggested to significantly influence the course of infection, leading to either lytic viral replication or a lysogenic state (7). Hence, the SAXS data, reflecting only the hexagonally ordered shell-DNA phase, where structure and density are unaffected by temperature (15, 17), could not explain the observation of the temperature-induced solid-to-fluid-like DNA transition (28). In our previously published SAXS data for phage λ (15), discussed in this work, we also analyzed the change in area of the Bragg peak with temperature at high q’s corresponding to X-ray diffraction from the hexagonally ordered shell-DNA phase in phage λ. The area is directly proportional to the number of ordered DNA base pairs in the shell phase. We found an initial decrease in the Bragg peak area with increasing temperature followed by a slight increase above 35 °C (corresponding to the DNA density transition temperature range). While this observation suggests an initial decrease in the volume of the shell-DNA phase followed by a subsequent volume increase after the transition, in agreement with our SANS data, a direct comparison between SAXS high q-range structure factor data and SANS low q-range form factor data should not be made. As discussed above, the presented SANS data analysis does not reflect on the local ordering of DNA in the capsid but only on the volume and density change of DNA in the core and shell phases, respectively.
SANS with contrast matching of capsid proteins allowed us to visualize temperature-induced changes in the global DNA structure in the capsid in the low q-region, with data analysis discriminating between two coexisting DNA phases in the core- and shell volumes of the capsid. We found that core-DNA volume constitutes only ~8% of the whole capsid volume at the temperature of transition (Fig. 5B). However, it is precisely the core-DNA that is packaged last and consequentially ejected first from the capsid, starting the infection. Thus, a density transition with temperature in the core-DNA structure is essential to facilitate rapid DNA ejection into a cell. In summary, this work reveals a direct link between the density distribution of virally packaged DNA in phage λ and its AFM-measured mobility (15). Furthermore, for the first time, to our knowledge, SAS (small-angle scattering) analysis reveals the form factor for DNA packaged in a viral capsid besides local DNA–DNA ordering analysis performed in the previous studies (10, 15, 30).
It can also be mentioned that it is plausible that the core-DNA portion of the packaged genome is oriented perpendicular to the hexagonally ordered shell phase, which could explain how large bending stress on the core-DNA is minimized. In parallel, sharp bending or kinking of dsDNA could occur in the center of the capsid, which excites local melting of a few base pairs, unwinding the DNA where the sharp bending occurs (66–69). Excitation of melted base pairs provides flexible local joints of DNA, which relax both bending and twisting stress of DNA, reducing the order of packing. This is also consistent with the more fluid-like behavior of the packed DNA observed after the transition.
Material and Methods
Phage λ Purification.
wt bacteriophage λ, strain cI857, with 100% wt (48.5kb) was purified according to the protocols described in ref. 7. Briefly, phage λ production was achieved by thermally inducing the lysogenic E. coli strain AE1 derived from S2773. The purified virus particles were dialyzed and stored in TM buffer (10 mM MgCl2, 50 mM Tris HCl, pH = 7.4). To produce empty phage λ particles, DNA-filled capsids were heated to 78 °C to trigger genome ejection through the phage tail while capsid remains intact (50). The ejected viral DNA was digested with DNase to avoid its contribution to the scattering signal.
Cryo-EM.
Cryo-EM samples were stored at 20 °C in TM-buffer (50 mM TrisHCl + 10 mM MgCl2) prior to plunge freezing. Cryo-EM data were collected on Tecnai F20 Twin transmission electron microscope operating at 200 keV equipped with Gatan 4kx4k CCD using the Leginon automated electron microscopy package, see ref. 70 for more details. Cryo-EM cross sections from icosahedral reconstructions and average radial density were obtained using Chimera software (71).
Solution SAXS.
SAXS measurements were carried out at the 12-ID B station at Advanced Photon Source at Argonne National Laboratory. A 12keV X-ray beam was used to illuminate the sample. Forty scans with 1s X-ray exposure time were collected and averaged. A buffer solution of the dialysis buffer (TM-buffer) was measured using the same SAXS setup and buffer scattering signal subtracted as the background (17).
SANS.
SANS measurements were performed at Center for High Resolution Neutron Scattering (CHRNS) NG3 vSANS beamline at NIST Center for Neutron Research (NCNR), Gaithersburg MD, USA and at ILL D11 beamline, Grenoble, France. Because the neutron flux is lower than X-rays, the sample was exposed several minutes at each configuration of the experiment, without damage or degradation of the sample in 1-mm pathlength quartz cells. The solvent exchange from 100% H2O to 43% D2O/H2 O ratio was done by dialysis. The scattering signal from TM-buffer was measured separately and subtracted from the sample (DNA-filled phage λ in TM-buffer) scattering signal as background. The data reduction and model analysis were done with the NCNR SANS macros in Igor Pro (34). SANS scattering intensity I(q), on an absolute scale, versus wave number q (magnitude of scattering vector , where is the neutron wavelength and is the scattering angle) was modeled with a spherical core-shell model with DNA in both the core and shell. We used NCNR macros to compute the SD values between the experimental data points and the fitting model parameters.
The core-DNA is the less-ordered/disordered DNA in the center in Fig. 2A and the shell-DNA is the ordered DNA near the periphery in Fig. 2A with water in between the strands. The initial value for the core SLD, ρcore = 3.21 × 1010 cm−2 at 43% D2O/H2O, was obtained by fitting the SANS data at 20 °C and the initial value for rcore = 9.51 nm was calculated from the cryo-EM density map at 20 °C shown in Fig. 2C. We kept ρshell fixed at 3.68 × 1010 cm−2 at 43% D2O/H2O (corresponding to the constant DNA packing density in the entire temperature interval, which includes the SLD for both DNA and the water-filled fraction between the DNA strands at 43% D2O/H2O). The ρshell value was obtained from the cryo-EM DNA density map in Fig. 2C, where the initial values of core-DNA radius and shell-DNA thickness at 20 °C were converted to volumes, Vcore and Vshell (note that Vshell refers to the volume of the shell thickness where Vcapsid = Vcore + Vshell). Based on the hexagonal lattice model for DNA packaged in the shell thickness (12), we calculated the length of DNA that can fit into Vshell. Thus, we were able to calculate the volume occupied by DNA, VDNA (by considering the cross-sectional area of a dsDNA helix). The difference between Vshell and VDNA is occupied by water molecules at 43% D2O/H2O ratio. SLDDNA and SLD43%D2O/H2O are known values (72–74). Hence, based on the volume fraction of DNA and water in Vshell, we estimated ρshell value. If we define the volume fraction of DNA in the shell thickness as ϕDNA = VDNA/Vshell and the volume fraction of water in the shell as ϕD2O/H2O = VD2O/H2O/Vshell, then ρshell = ϕDNA x SLDDNA + ϕD2O/H2O x SLDD2O/H2O.
The spherical core–shell model used to fit SANS data is given by equation below:
where is the core-DNA radius; is the shell-DNA radius = 28.3 nm (equal to the inner capsid radius), t is the shell-DNA thickness; is the core (i = core) or capsid (i = capsid) volume; is the SLD for core-DNA (i = core), shell-DNA (i = shell) and solvent (i = solv), φ is a scaling factor that was kept constant at all temperatures, and q is wave number. SI Appendix, Fig. S3 shows a panel of the resultant core–shell profiles at all temperatures with all curves combined within a selected low q-range, reflecting DNA form factor. SI Appendix, Table S1 summarizes the SANS fitting parameters obtained with the smeared core-shell model.
Disclaimer.
Certain commercial equipment, materials, software, or suppliers are identified in this paper to foster understanding. Such identification does not imply recommendation or endorsement by the National Institute of Standards and Technology nor does it imply that the materials or equipment identified are necessarily the best available for the purpose.
Data, Materials, and Software Availability
All study data are included in the article and/or SI Appendix.
Acknowledgments
We greatly acknowledge Ralf Schweins, Tom Goddard, Jacob Kirkensgaard, and Julian Oberdisse for valuable discussions during SANS scattering analysis. We acknowledge Alberto Brandariz for assistance with sample preparation. Access to the NG3 vSANS was provided by the CHRNS, a partnership between the National Institute of Standards and Technology and the NSF under Agreement No. DMR-2010792. This work was supported by the Swedish Research Council grants (Vetenskapsrådet) 349- 2014-3962 and 2019-05192 (all to A.E.), and Mats Paulsson Foundation to A.E.
Author contributions
A.E. designed research; J.R.V.V., E.T., and A.E. performed research; J.R.V.V., E.T., S.K., and A.E. contributed new reagents/analytic tools; J.R.V.V., S.K., and A.E. analyzed data; and J.R.V.V. and A.E. wrote the paper.
Competing interests
The authors declare no competing interest.
Supporting Information
Appendix 01 (PDF)
- Download
- 1.57 MB
References
1
R. Gallet, S. Kannoly, I. N. Wang, Effects of bacteriophage traits on plaque formation. BMC Microbiol. 11, 181 (2011).
2
A. Brandariz-Nunez, S. J. Robinson, A. Evilevitch, Pressurized DNA state inside herpes capsids-A novel antiviral target. PLoS Pathog. 16, e1008604 (2020).
3
A. Evilevitch, L. Lavelle, C. M. Knobler, E. Raspaud, W. M. Gelbart, Osmotic pressure inhibition of DNA ejection from phage. Proc. Natl. Acad. Sci. U.S.A. 100, 9292–9295 (2003).
4
M. Castelnovo, A. Evilevitch, DNA ejection from bacteriophage: Towards a general behavior for osmotic-suppression experiments. Eur. Phys. J. E Soft Matter 24, 9–18 (2007).
5
K. J. Hanhijarvi, G. Ziedaite, M. K. Pietila, E. Haeggstrom, D. H. Bamford, DNA ejection from an archaeal virus–a single-molecule approach. Biophys. J. 104, 2264–2272 (2013).
6
D. W. Bauer, J. B. Huffman, F. L. Homa, A. Evilevitch, Herpes virus genome, the pressure is on. J. A. Chem. Soc. 135, 11216–11221 (2013).
7
A. Evilevitch, The mobility of packaged phage genome controls ejection dynamics. Elife 7, e37345 (2018).
8
D. E. Smith et al., The bacteriophage phi 29 portal motor can package DNA against a large internal force. Nature 413, 748–752 (2001).
9
G. C. Lander et al., DNA bending-induced phase transition of encapsidated genome in phage lambda. Nucleic Acids Res. 41, 4518–4524 (2013).
10
W. C. Earnshaw, S. C. Harrison, DNA arrangement in isometric phage heads. Nature 268, 598–602 (1977).
11
S. Tzlil, J. T. Kindt, W. M. Gelbart, A. Ben-Shaul, Forces and pressures in DNA packaging and release from viral capsids. Biophys. J. 84, 1616–1627 (2003).
12
P. K. Purohit, J. Kondev, R. Phillips, Mechanics of DNA packaging in viruses. Proc. Nat. Acad. Sci. U.S.A. 100, 3173–3178 (2003).
13
I. S. Gabashvili, A. Grosberg, Dynamics of double stranded DNA reptation from bacteriophage. J. Biomol. Struct. Dyn. 9, 911–920 (1992).
14
Z. T. Berndsen, N. Keller, S. Grimes, P. J. Jardine, D. E. Smith, Nonequilibrium dynamics and ultraslow relaxation of confined DNA during viral packaging. Proc. Natl. Acad. Sci. U.S.A. 111, 8345–8350 (2014).
15
T. Liu et al., Solid-to-fluid-like DNA transition in viruses facilitates infection. Proc. Natl. Acad. Sci. U.S.A. 111, 14675–14680 (2014).
16
U. Sae-Ueng et al., Solid-to-fluid DNA transition inside HSV-1 capsid close to the temperature of infection. Nat. Chem. Biol. 10, 861–867 (2014).
17
D. Li et al., Ionic switch controls the DNA state in phage lambda. Nucleic acids Res. 43, 6348–6358 (2015).
18
L. Zeng et al., Decision making at a subcellular level determines the outcome of bacteriophage infection. Cell 141, 682–691 (2010).
19
Q. Shao, A. Hawkins, L. Zeng, Phage DNA dynamics in cells with different fates. Biophys. J. 108, 2048–2060 (2015).
20
M. M. Castellanos, A. McAuley, J. E. Curtis, Investigating structure and dynamics of proteins in amorphous phases using neutron scattering. Comput. Struct. Biotechnol. J. 15, 117–130 (2017).
21
Y. Wei, M. J. Hore, Characterizing polymer structure with small-angle neutron scattering: A Tutorial. J. Appl. Phys. 129, 171101 (2021).
22
M. J. Hollamby, Practical applications of small-angle neutron scattering. Phys. Chem. Chem. Phys. 15, 10566–10579 (2013).
23
B. Jacrot, C. Chauvin, J. Witz, Comparative neutron small-angle scattering study of small spherical RNA viruses. Nature 266, 417–421 (1977).
24
M. Comellas-Aragonès et al., Solution scattering studies on a virus capsid protein as a building block for nanoscale assemblies. Soft Matter 7, 11380–11391 (2011).
25
P. Timmins, J. Witz, A neutron scattering study of the structure of compact and swollen forms of southern bean mosaic virus. Virology 119, 42–50 (1982).
26
S. Cusack, R. W. H. Ruigrok, P. C. J. Krygsman, J. E. Mellema, Structure and composition of influenza virus: A small-angle neutron scattering study. J. Mol. Biol. 186, 565–582 (1985).
27
D. A. Kuzmanovic, I. Elashvili, C. Wick, C. O’Connell, S. Krueger, Bacteriophage MS2: Molecular weight and spatial distribution of the protein and RNA components by small-angle neutron scattering and virus counting. Structure 11, 1339–1348 (2003).
28
M. d. Frutos et al., Can changes in temperature or ionic conditions modify the DNA organization in the full bacteriophage capsid?. J. Phys. Chem. B 120, 5975–5986 (2016).
29
J. Tang et al., Peering down the barrel of a bacteriophage portal: The genome packaging and release valve in p22. Structure 19, 496–502 (2011).
30
X. Qiu et al., Salt-dependent DNA-DNA spacings in intact bacteriophage lambda reflect relative importance of DNA self-repulsion and bending energies. Phys. Rev. Lett. 106, 028102 (2011).
31
D. Wang, J. Lal, D. Moses, G. C. Bazan, A. J. Heeger, Small angle neutron scattering (SANS) studies of a conjugated polyelectrolyte in aqueous solution. Chem. Phys. Lett. 348, 411–415 (2001).
32
D. Wang, D. Moses, G. C. Bazan, A. J. Heeger, J. Lal, Conformation of a conjugated polyelectrolyte in aqueous solution: Small angle neutron scattering. J. Macromol. Sci. Part A 38, 1175–1189 (2001).
33
B. Hammouda, Probing nanoscale structures-the sans toolbox (National Institute of Standards and Technology, 2008), pp. 1–717.
34
S. R. Kline, Reduction and analysis of SANS and USANS data using IGOR Pro. J. Appl. Crystallogr. 39, 895–900 (2006).
35
C. Stanley, D. C. Rau, Evidence for water structuring forces between surfaces. Curr. Opin. Colloid Interface Sci. 16, 551–556 (2011).
36
D. C. Rau, V. A. Parsegian, Direct measurement of the intermolecular forces between counterion-condensed DNA double helices. Evidence for long range attractive hydration forces. Biophys. J. 61, 246–259 (1992).
37
E. Raspaud, D. Durand, F. Livolant, Interhelical spacing in liquid crystalline spermine and spermidine-DNA precipitates. Biophys. J. 88, 392–403 (2005).
38
P. K. Purohit et al., Forces during bacteriophage DNA packaging and ejection. Biophys. J. 88, 851–866 (2005).
39
F. P. Booy et al., Liquid-crystalline, phage-like packing of encapsidated DNA in herpes simplex virus. Cell 64, 1007–1015 (1991).
40
W. Jiang et al., Structure of epsilon15 bacteriophage reveals genome organization and DNA packaging/injection apparatus. Nature 439, 612–616 (2006).
41
M. E. Cerritelli et al., Encapsidated conformation of bacteriophage T7 DNA. Cell 91, 271–280 (1997).
42
A. Huet, J. F. Conway, L. Letellier, P. Boulanger, In vitro assembly of the T = 13 procapsid of bacteriophage T5 with its scaffolding domain. J. Virol. 84, 9350–9358 (2010).
43
L. R. Comolli et al., Three-dimensional architecture of the bacteriophage phi29 packaged genome and elucidation of its packaging process. Virology 371, 267–277 (2008).
44
A. Leforestier, F. Livolant, The bacteriophage genome undergoes a succession of intracapsid phase transitions upon DNA ejection. J. Mol. Biol. 396, 384–395 (2010).
45
J. San Emeterio, L. Pollack, Visualizing a viral genome with contrast variation small angle X-ray scattering. J. Biol. Chem. 295, 15923–15932 (2020).
46
K. L. Sarachan, J. E. Curtis, S. Krueger, Small-angle scattering contrast calculator for protein and nucleic acid complexes in solution. J. Appl. Crystallogr. 46, 1889–1893 (2013).
47
T. Zemb, P. Lindner, Neutron, X-Rays and Light. Scattering Methods Applied to Soft Condensed Matter (North Holland, 2002), chap. 7 by P. Schurtenberger.
48
R.-J. Roe, Methods of X-Ray and Neutron Scattering in Polymer Science (Oxford University Press on Demand, 2000).
49
D. W. Bauer et al., Exploring the balance between DNA pressure and capsid stability in herpesviruses and phages. J. Virol. 89, 9288–9298 (2015).
50
D. W. Bauer et al., Exploring the balance between DNA pressure and capsid stability in Herpes and phage. J. Virol. 89, 9288–9298 (2015), https://doi.org/10.1128/JVI.01172-15.
51
D. Li et al., Ionic switch controls the DNA state in phage λ. Nucleic Acids Res. 43, 6348–6358 (2015).
52
T. Liu et al., Solid-to-fluid–like DNA transition in viruses facilitates infection. Proc. Natl. Acad. Sci. U.S.A. 111, 14675–14680 (2014).
53
J. S. Pedersen, Analysis of small-angle scattering data from colloids and polymer solutions: Modeling and least-squares fitting. Adv. Colloid Interface Sci. 70, 171–210 (1997).
54
G. C. Lander et al., Bacteriophage lambda stabilization by auxiliary protein gpD: Timing, location, and mechanism of attachment determined by cryo-EM. Structure 16, 1399–1406 (2008).
55
R. Podgornik, M. A. Aksoyoglu, S. Yasar, D. Svensek, V. A. Parsegian, DNA equation of state. In vitro vs in viro. J. Phys. Chem. B 120, 6051–6060 (2016).
56
G. C. Lander et al., DNA bending-induced phase transition of encapsidated genome in phage lambda. Nucleic Acids Res. 41, 4518–4524 (2013), https://doi.org/10.1093/nar/gkt137.
57
S. Y. Park, D. Harries, W. M. Gelbart, Topological defects and the optimum size of DNA condensates. Biophys. J. 75, 714–720 (1998).
58
V. A. Parsegian, R. P. Rand, N. L. Fuller, D. C. Rau, Osmotic stress for the direct measurement of intermolecular forces. Methods Enzymol. 127, 400–416 (1986).
59
J. Kindt, S. Tzlil, A. Ben-Shaul, W. M. Gelbart, DNA packaging and ejection forces in bacteriophage. Proc. Natl. Acad. Sci. U.S.A. 98, 13671–13674 (2001).
60
A. A. Kornyshev, D. J. Lee, S. Leikin, A. Wynveen, Structure and interactions of biological helices. Rev. Mod. Phys. 79, 943–996 (2007).
61
S. Geggier, A. Kotlyar, A. Vologodskii, Temperature dependence of DNA persistence length. Nucleic Acids Res. 39, 1419–1426 (2011).
62
L. Zeng, I. Golding, Following cell-fate in E. coli after infection by phage lambda. J. Vis. Exp. 56, e3363 (2011), https://doi.org/10.3791/3363.
63
M. G. Rossmann, M. C. Morais, P. G. Leiman, W. Zhang, Combining X-ray crystallography and electron microscopy. Structure 13, 355–362 (2005).
64
T. S. Baker, N. H. Olson, S. D. Fuller, Adding the third dimension to virus life cycles: Three-dimensional reconstruction of icosahedral viruses from cryo-electron micrographs. Microbio. Mol. Bio. Rev. 63, 862–922 (1999).
65
W. C. Earnshaw, R. W. Hendrix, J. King, Structural studies of bacteriophage lambda heads and proheads by small angle X-ray diffraction. J. Mol. Biol. 134, 575–594 (1979).
66
J. Widom, DNA bending and kinking. Nature 309, 312–313 (1984).
67
T. E. Cloutier, J. Widom, Spontaneous sharp bending of double-stranded DNA. Mol. Cell 14, 355–362 (2004).
68
R. Vafabakhsh, T. Ha, Extreme bendability of DNA less than 100 base pairs long revealed by single-molecule cyclization. Science 337, 1097–1101 (2012).
69
Q. Du, A. Kotlyar, A. Vologodskii, Kinking the double helix by bending deformation. Nucleic Acids Res. 36, 1120–1128 (2008).
70
G. C. Lander et al., DNA bending-induced phase transition of encapsidated genome in phage λ. Nucleic Acids Res. 41, 4518–4524 (2013).
71
E. F. Pettersen et al., UCSF Chimera—A visualization system for exploratory research and analysis. J. Comput. Chem. 25, 1605–1612 (2004).
72
C. L. Schildkraut, J. Marmur, P. Doty, Determination of the base composition of deoxyribonucleic acid from its buoyant density in CsCl. J. Mol. Biol. 4, 430–443 (1962).
73
B. Jacrot, The study of biological structures by neutron scattering from solution. Rep. Progress Phys. 39, 911 (1976).
74
B. Hammouda, Probing nanoscale structures–The SANS toolbox (National Institute of Standards and Technology, 2012).
Information & Authors
Information
Published in
Classifications
Copyright
Copyright © 2023 the Author(s). Published by PNAS. This open access article is distributed under Creative Commons Attribution-NonCommercial-NoDerivatives License 4.0 (CC BY-NC-ND).
Data, Materials, and Software Availability
All study data are included in the article and/or SI Appendix.
Submission history
Received: December 4, 2022
Accepted: September 25, 2023
Published online: October 30, 2023
Published in issue: November 7, 2023
Keywords
Acknowledgments
We greatly acknowledge Ralf Schweins, Tom Goddard, Jacob Kirkensgaard, and Julian Oberdisse for valuable discussions during SANS scattering analysis. We acknowledge Alberto Brandariz for assistance with sample preparation. Access to the NG3 vSANS was provided by the CHRNS, a partnership between the National Institute of Standards and Technology and the NSF under Agreement No. DMR-2010792. This work was supported by the Swedish Research Council grants (Vetenskapsrådet) 349- 2014-3962 and 2019-05192 (all to A.E.), and Mats Paulsson Foundation to A.E.
Author contributions
A.E. designed research; J.R.V.V., E.T., and A.E. performed research; J.R.V.V., E.T., S.K., and A.E. contributed new reagents/analytic tools; J.R.V.V., S.K., and A.E. analyzed data; and J.R.V.V. and A.E. wrote the paper.
Competing interests
The authors declare no competing interest.
Notes
This article is a PNAS Direct Submission. U.R. is a guest editor invited by the Editorial Board.
Authors
Metrics & Citations
Metrics
Altmetrics
Citations
Cite this article
Temperature-induced DNA density transition in phage λ capsid revealed with contrast-matching SANS, Proc. Natl. Acad. Sci. U.S.A.
120 (45) e2220518120,
https://doi.org/10.1073/pnas.2220518120
(2023).
Copied!
Copying failed.
Export the article citation data by selecting a format from the list below and clicking Export.
Cited by
Loading...
View Options
View options
PDF format
Download this article as a PDF file
DOWNLOAD PDFLogin options
Check if you have access through your login credentials or your institution to get full access on this article.
Personal login Institutional LoginRecommend to a librarian
Recommend PNAS to a LibrarianPurchase options
Purchase this article to access the full text.