• No results found

Nanocrystal imaging using intense and ultrashort X-ray pulses

N/A
N/A
Protected

Academic year: 2022

Share "Nanocrystal imaging using intense and ultrashort X-ray pulses"

Copied!
6
0
0

Loading.... (view fulltext now)

Full text

(1)

Nanocrystal imaging using intense and ultrashort X-ray pulses

Carl Caleman, G¨osta Huldt, Carlos Ortiz, Filipe R. N. C. Maia, Erik G. Marklund, Fritz G. Parak, David van der Spoel and Nicu¸sor Tˆımneanu

Physik Department E17, Technische Universit¨at M¨unchen, James-Franck-Strasse, DE-85748 Garching, Germany,Department of Cell and Molecular Biology, Biomedical Centre, Uppsala University, Box 596, SE-751 24 Uppsala, Sweden, andDepartment of Physics and Material Science, Uppsala University, Box 530, SE-751 21 Uppsala, Sweden Submitted to Proceedings of the National Academy of Sciences of the United States of America

Structural studies of biological macromolecules are severely limited by radiation damage. Traditional crystallography curbs the effects of damage by spreading damage over many copies of the molecule of interest in the crystal. X-ray lasers offer an additional oppor- tunity for limiting damage by out-running damage processes with ultrashort and very intense X-ray pulses. Such pulses may allow the imaging of single molecules, clusters or nanoparticles, but coherent flash imaging will also open up new avenues for structural studies on nano- and micro-crystalline substances. This paper addresses the potentials and limitations of nanocrystallography with extremely in- tense coherent X-ray pulses. We use urea nanocrystals as a model for generic biological substances, and simulate the primary and sec- ondary ionization dynamics in the crystalline sample. The results establish conditions for diffraction experiments as a function of X- ray fluence, pulse duration, and the size of nanocrystals.

nanocrystallography | X-ray Free Electron Laser | radiation damage | coherent

flash imaging

Abbreviations: XFEL, X-ray Free Electron Laser; LCLS, Linac Coherent Light Source;

MD, Molecular Dynamics

New light sources have had a significant influence on natural sci- ences throughout history. Radio transmitters, X-ray sources and optical lasers triggered fundamental transformations both in sci- ence and society. As a consequence, expectations are high, regard- ing the impact of the long-awaited first hard X-ray lasers, which will start user operation this year. These lasers produce ultra-short and ex- tremely intense coherent X-ray pulses with a peak brilliance, exceed- ing that of conventional synchrotron sources by more than billion times. The FLASH soft X-ray free-electron laser in Germany [1]

was the first to reach into the X-ray frequencies, and it is a fully operational user facility today. The LINAC Coherent Light Source (LCLS) [2] in the USA is a hard X-ray laser that has produced first light and it is lasing at 1.5 ˚A . Similar projects are under way in Japan [3] and in Europe. In addition to these linear accelerator- based machines, table-top X-ray lasers, driven by optical lasers, have started making their mark [4–7]. Short, intense, coherent, hard X-ray pulses can be exploited for new experiments in disciplines, ranging from experimental astrophysics to structural biology [8]. Such X-ray pulses could open the door to single molecule imaging, i.e. retriev- ing atomic structures from large biomolecules without the need of a crystalline sample [9, 10].

Structures could also be determined from nanocrystalline mate- rials. In the process of crystallization, many macromolecules (e.g.

membrane proteins) do not form large crystals. However, they often form sub-micron crystals but these are usually too small to gener- ate useful diffraction data at a conventional synchrotron source. It has been suggested [11–13] that such nanocrystals could be used for structural studies with X-ray lasers.

Any sample exposed to an intense X-ray pulse will be ionized, and extensive ionization destroys the sample. The time scale on which this process occurs is critical for obtaining an interpretable diffraction pattern to yield an atomic structure of the sample. In principle, the X-ray pulse must be short enough such that the entire pulse passes through the sample before a major disarrangement of the

Fig. 1.Crystal size and the extent of secondary electron cascades. The figure in (a) shows the overall dimensions of electron clouds produced during the thermaliza- tion of an 8 keV photoelectron and a 0.4 keV Auger electron (ejected from a nitrogen atom) inside a large urea crystal. Similar cascade sizes are produced in protein crys- tals, in an X-ray diffraction experiment. The total number of ionizations was 18 in the Auger cascade, and 118 in the photoelectron cascade at 100 fs after the emission of the primary electrons. At this point, the radius of gyration of the photoelectron cascade reached 2µm, and that of the Auger electron cascade 200 nm. The photo- electron cascade is significantly bigger than a typical nanocrystal/microcrystal under considerations here (b). Using lysozyme as an example, the protein nanocrystal would contain about 300,000 unit cells (c).

atomic and electronic configurations takes place. The ionizations due to the direct photoabsorption and subsequent secondary processes af- fect the ability to get useful structural information from the diffrac- tion pattern in three ways. (i) Ionization decreases the elastic X-ray scattering power of the atoms. (ii) Removal of electrons from the atoms leaves behind positively charged ions that repel each other due to Coulomb forces, leading to the destruction of the structure. (iii)

Reserved for Publication Footnotes

(2)

Fig. 2.Evolution of secondary electron cascades in a urea crystal over time. (a) Number of secondary ionizations produced by a photoelectron of 8 KeV and by Auger electrons (impact energies: 250 eV for carbon, 400 eV for nitrogen, 500 eV for oxygen). (b) Spatial evolution of the secondary electron cloud from a photoelectron in a large urea crystal depicted through the radial electron density as a function of time. (c) Secondary electron cloud from an Auger electron (nitrogen). The termalization of electrons from oxyen and carbon has similar features. Black lines show changes in the radius of gyration (defined in Methods) of the electron clouds.

Free electrons either leave the sample, if their energy is high enough, or remain in the sample as a background electron gas, in which case they will contribute to noise in the diffraction pattern.

There are no experiments on the dynamics of radiation dam- age from FEL pulses at ˚angstr¨om wavelengths. Experiments pub- lished so far reach into the soft X-ray regime (down to about 13.5 nm wavelength) [14–17]. Data about this regime come from experi- ments performed at the FLASH free-electron laser in Hamburg. The- oretical models extend the picture into the unexplored hard X-ray regime [9, 18–20]. The explosion mechanism strongly depends on sample size. Electrons ejected from atoms during exposure prop- agate through the sample, and cause further ionization by eliciting secondary electron cascades. The extent of ionization through this mechanism depends on the size of the sample. Photoelectrons re- leased by X-rays of 1.5 ˚A wavelength are fast (53 nm/fs), and they can escape from small samples early in an exposure (Figure 1). In contrast, Auger electrons are slow (9.5 nm/fs for carbon) and it is likely that they will thermalize even in a small sample (Figure 1).

In late phases of an exposure, a significant fraction of the emitted electrons will not be able to escape the increased positive potential of the sample even if the sample is small. For small samples, the explosion is dominated by Coulomb processes. This is driven by the repulsion of the positive ions left behind by electrons leaving the sample. In big samples, electrons will be trapped simply because they lose energy before reaching the surface. Trapped electrons increase the kinetic energy of the sample through thermal processes, while slowing the Coulomb explosion by partially screening the positively charged core. Predictions point to a transition from Coulomb explo- sion to a hydrodynamic explosion. A positively charged surface layer is formed and it peels off, burning the sample from outside towards the core. The expansion of the core is driven by thermal processes as the electron pressure grows [18, 19].

The aim of this work is to study the damage caused by ion- ization in nanocrystals of biological material, with sizes up to one micrometer. Crystals larger than one micrometer are normally con- sidered viable and diffract good enough at conventional synchrotron sources. Our study aims at providing a screening tool for usable sam- ple sizes in nanocrystallography experiments, with regard to sample size and X-ray laser pulse parameters. Serial crystallography experi- ments with sub-micron protein crystals have recently been performed at a synchrotron [21], and showed that powder diffraction data can be obtained using a continuous microjet of nanocrystals. It has also been found that longitudinal coherence properties of the X-ray lasers limit the resolution of single-particle diffraction imaging [22]. At a wavelength of 1.5 ˚A the particles have to be smaller than 500 nm in diameter to achieve imaging with a resolution length of less than 2 ˚A.

Detailed description of electron impact ionization and secondary electron cascades has been presented in the literature for different ma-

terials [23–25]. The dynamics of photoelectrons in protein crystals have been investigated [26] (without consideration to Auger emission or secondary electron cascades. The results suggest that radiation damage can be limited by reducing the crystal size. The present paper steps beyond these studies, and treats photo-emission, Auger emis- sion and cascade processes during exposure of a biological nanocrys- tal to an XFEL pulse. We investigate the effect of radiation damage for samples of several sizes and different X-ray pulse lengths, and the consequences it has on diffractive imaging of biological samples as they are exposed to an XFEL pulse.

Electron impact ionization and secondary cascades Figure 1 shows the simulated dynamics of ionization in an infinitely large urea crystal exposed to an X-ray pulse with 1.5 ˚A wavelength.

A nanocrystal is smaller than the electron cascades and most photo- electrons may leave the sample before thermalization. In a step-wise approach, we first treat the thermalization of electrons with various energies, corresponding to photoelectrons or different Auger elec- trons in urea. In the next step, we combine the primary and secondary ionization effects to describe the entire dynamics of the system during and after the X-ray pulse (see the Methods section). Detailed descrip- tion of the model, electron scattering on atoms, treatment of electron- hole recombinations and electron-phonon interactions is presented in [25].

At low atomic numbers, a single photo-ionization releases elec- trons at two distinctly different energies. The energy of the photo- electron corresponds to the difference between the photon energy (8.3 keV in this case) and the K-shell binding energy, while Auger elec- trons carry kinetic energy dependent on the Auger process. Auger energies for carbon, nitrogen and oxygen in the urea target are ap- proximately 250 eV, 400 eV and 500 eV, which is more than an order of magnitude lower than the energy of a photoelectron (≈8 keV). In the energy range considered here, the number of secondary ioniza- tions produced by the inelastic scattering of a single electron, scales linearly with the energy of the initial electron [23, 25] (Figure 2a).

The onset of the electron cascade scales also with the incident electron energy. The electron cloud initiated by an energetic photo- electron thermalizes much slower than electrons in Auger cascades, as the electron travels further between each scattering event in the crystal due to its higher energy and low interaction cross section (Fig- ure 2b). However, when the energy distribution in the electron cloud has reached impact ionization threshold and no more ionizations can occur, the cloud has generated ten times as many secondary electrons than an Auger electron. In the same time, the photoelectron cloud is more than an order of magnitude larger than the Auger electron induced cloud. Figure 2b shows radial electron density from the pho- toelectron and Auger electron cloud as these develop in time. At each time point, the radial density is normalized to give the total number

(3)

Fig. 3.Evolution of secondary electron cascades as a function of the X-ray pulse length in a urea crystal. The pulse is centered at t=0. (a) The probability distribution (normalized to 1) for the emission of a photoelectron (1.5 ˚A X-ray wavelength) and the subsequent emission of an Auger electron from carbon, nitrogen or oxygen during a 10 fs X-ray pulse (full width at half maximum, FWHM). Auger electron lifetimes are 11.3 fs s for carbon, 8.3 fs for nitrogen and 6.6 fs for oxygen. (b) Evolution of the photoelectron cascades in the urea crystal as a function of pulse length. The data are normalized to a single photoionization occurrence. (c) Evolution of the Auger electron cascades from carbon, nitrogen, and oxygen under the same conditions as in (b). The black line shows when 99.5% of the pulse has passed.

of secondary ionizations (Figure 2a). For comparison, the radii of gyration (defined in the Method section) of the photoelectric cloud and Auger cloud are presented with black lines. As shown, radius of gyration describes well the spatial extent of the electron clouds and it will be used to characterize electron clouds throughout the paper (Figure 2b).

Sample damage is caused by ionization. At a wavelength of 1.5 ˚A, the ratio between elastically (coherently) scattered photons and photoionization is 1:32 for oxygen, 1:26 for nitrogen and 1:20 for carbon [27]. Most of the incoming photons will primarily con- tribute to ionization in the sample and only a few will generate a coherent diffraction image. The loss of an electron from a carbon atom will decrease its scattering power by about 17%. This is 14%

for nitrogen and about 12% for oxygen. However, ionizations occur- ring after the X-ray pulse has left the sample will no longer influence the diffraction pattern. Therefore, more photoionizations can be al- lowed when using a very short X-ray pulse, as only few secondary electrons will be generated during the pulse as compared to longer pulses (Figure 3).

In a very large crystal the ejected electrons have nowhere to es- cape, and no Coulomb explosion is possible. In such a system the pressure of the ejected electrons drives a hydrodynamic expansion of the sample, and heats the system. At 8.3 keV photon energy, a single photoelectron will liberate about 390 electrons as it comes to a ther- mal equilibrium in a large sample (Figure 2a). This free electron gas will contribute to a Thomson background in the diffraction pattern.

Some of these ionizations can be avoided by using smaller crystals.

0 10 20 30 40 50 60 70 80 90 100

0 2 4 6 8 10 12 14 16 18

Pulse length (FWHM, fs)

Number of ionizations cascadeAuger

0 50 100 150 200 250 Radius of gyration (nm) cascade x 0.1Photo-

Fig. 4.Number of secondary ionizations (left) and the radius of gyration (right, for definition see Methods) at the point when 99.5% of the pulse passed the sample.

These values are plotted as a function of pulse length (FWHM). The solid red curves for the photoelectron cascades are scaled down by a factor 10. The Auger cascades (blue dotted line) contains all secondary electrons from C, N and O combined, and are normalized to a single photoionization event.

In a sample that is small compared to the size of the X-ray beam, photoionization events will occur with equal probability throughout the entire sample. If the sample is smaller than the mean free path of photoelectrons or Auger electrons, many of the high-energy electrons are expected to escape the system during the exposure.

When investigating the feasibility of imaging crystals of various sizes, we calculate the total number of electrons generated for each crystal size based on the electron cloud dynamics as a function of pulse length (Figure 3). If the electron cloud is expected to be larger than the crystal, electrons will escape and it is necessary to compen- sate for that effect in the calculation. Figure 3 provides the basis for calculating the tolerable levels of radiation damage, as a function of pulse length and size of electron cloud.

It is worth noting the differences between the present method of treating the entire dynamics of the electron clouds versus pulse length (Figures 3 and 3) and the treatment of single electron thermalization (Figure 2). For short X-ray pulses which are comparable to Auger lifetimes, the single electron approximation overestimates the Auger cascade but provides a good approximation for the photoelectron cas- cades. For pulses longer than 10 femtoseconds, the single electron method (Figure 2) underestimates the ionization and development of the photo-cascades.

Imaging nanocrystals

Three-dimensional (3D) structural studies require a 3D data set.

Since the XFEL pulse will destroy the sample, structure determina- tion relies on the fact that the experiment can be repeated, i.e. that many crystals can be produced from the same material, and then delivered into the X-ray beam in a repetitive and controlled man- ner. Rather than building up a complete X-ray diffraction data set by rotating the crystal and collecting a sequence of diffraction im- ages, as is done in conventional crystallography, one will be need to scale together individual diffraction images from many of differ- ent nanocrystals, in order to build up complete 3D data set. It is yet to be proven that it is possible to effectively combine such data, but it is reasonable to expect that this computational problem can be solved, as it has in the case of continuous diffraction pattern [28].

Assuming that the XFEL provides enough X-ray photons per pulse to record a diffraction pattern from a single shot, a crystal with 100 unit cells produces a discrete diffraction pattern, just as any large crystal, and conventional X-ray phasing technique can be used. Fur- thermore, oversampling techniques for direct phase retrival may also be employed for a 3D structural determination [29] For an average size protein molecule, like Deacetoxycephalosporin (DAOCS) with a unit cell size of (a=10.7 nm, b=10.7 nm, c=7.01 nm) [30] a crystal edge of 100 nm corresponds to around 1000 unit cells. In the case of a single molecule, where a continuous diffraction image is generated, different reconstruction algorithms have to be employed [16, 28, 31].

(4)

0 0.2 0.4 0.6 0.8 1

-15 -10 -5 0 5 10 15

0 0.02 0.04 0.06 0.08 0.1

Relative intensity of Bragg peak RMSD atomic positions (nm)

Time (fs) Pulse

RMSD urea

(110) (220) (330) (440)

Fig. 5.Degradation of Bragg peaks during exposure to an X-ray pulse in a urea crystal. Pulse length: 10 femtoseconds (FWHM) centered at t=0, 1.5 ˚A wavelength.

The pulse intensity is adjusted to create one ionization per atom when 99.5% of the pulse passes through the sample. Peak intensities at different resolutions are repre- sented by the (hkl) reflections and are normalized to one, based on intensities from undamaged crystals. The (440) reflection corresponds to 2 ˚A resolution (scattering angle at 45) and has an average degradation of 50% due to ionization and atomic displacement. The dashed black line shows the average root mean square deviations (RMSD) in atomic positions during illumination.

To put damage caused by secondary ionization into a perspec- tive of what resolution one can expect to achieve in the reconstructed structure, damage needs to be related to photon scattering. We model the noise to be Poisson distributed and require a minimum of 9 pho- tons per Bragg peak. This corresponds to a signal-to-noise ratio (SNR) of three, with SNR defined as the expected signal over the standard deviation of the noise. At this level, it should be possible to detect a diffraction peak, even without averaging over many diffrac- tion images. Using averaging one could probably detect Bragg peaks even with a SNR below one. We aim for a resolution of the recon- structed structure of 2 ˚A, and require that each Bragg peak scattered at the angle θ=45 to have a photon count higher than 9 photons.

Using Equation 5 one can calculate the number of scattered photons from the incident X-ray beam. This gives us a tool to map up the pa- rameter space (crystal dimensions, XFEL pulse parameters), e.g. to obtain a 2 ˚A resolution structure. Naturally a larger number of unit cells gives a higher signal. On the other had the radiation damage puts restrictions on the maximal crystal size. As pointed out ear- lier [26] larger crystals suffer more ionizations due to the trapping of photoelectrons and the subsequent secondary cascades. Assum- ing that only one ionization per non-hydrogen atom is allowed, one can predict the maximum crystal size where ionization is kept below this maximum ionization threshold. Figure 6 illustrates the number of photons scattered in a Bragg peak at 2 ˚A resolution, as a function of unit cell size and crystal size. The solid line shows the limit of 9 photons per peak, where the achieved signal is considered sufficient.

Our choice for a threshold of one ionization per atom in aver- age is a reasonably conservative choice. The scattering power of the atoms is reduced, leading to a drop of up to 30% of scattered intensity.

The drop can be accounted for by employing correction algorithms on diffraction patterns based on statistical methods [32]. At the same time, ionization leads to destruction of the crystalline order and loss of signal in the Bragg peaks. The total drop in the Bragg peak in- tensity is shown in Figure 3, for a 10 fs long pulse, considering an average of one ionization per atom. The results presented below can only improve if a lower threshold will be chosen.

The number of electrons generated within the exposure of the X- ray pulse increases with pulse length (Figure 3). Limiting radiation damage enforces the use of shorter pulses for higher photon fluxes.

Shortening the X-ray pulse length is usually challenging, therefore it might be better from the imaging perspective to use lower X-ray fluxes at longer pulse lengths.

To quantify these results we list four examples of biomolecules with different unit cell sizes. Naturally a larger crystal can host more molecules, and is therefore suitable for imaging of larger molecules.

Due to the fact that the high-energy electrons thermalize rather slowly, the shorter the pulse, the larger the molecule that can be im- aged. We have made calculations for pulses up to 100 fs long, fow which atoms will have enough time to disorder. However, keeping in mind that the samples explode in shells, i.e. the outer part of the sample is destroyed faster than the inner [18, 19] useful structural in- formation might still be left in the center of a large crystal after 100 fs. In the following four examples the pulse is assumed to consist of 1011photons, and with a length shorter than 100 fs (Figure 6b).

For a small protein such as Lysozyme (1IEE) [33], with a unit cell of (a=7.7 nm, b=7.7 nm, c=3.7 nm) a cubic crystal with a side of 150 nm would generate a good image. To reach the same SNR for an average sized protein, like DAOCS (1UNB) [30] (a=10.7 nm, b=10.7 nm, c=7.01 nm), the crystal would have to be have a side of at least 200 nm. To image Rubisco (8RUC) [34], a rather large pro- tein molecule, (a= 17.1 nm, b=14.3 nm, c=12.7 nm) the crystal side would need to be as large as 250 nm. For imaging a large crystallized virus, like MS2 (2MS2) [35] (a=28.8 nm, b=28.8 nm, c=65.3 nm), one would need to have crystals larger than 600 nm.

For a lower pulse intensity, larger crystals are needed. With pulses of 1010photons (which is in the range of what tabletop laser based free electron lasers are expected to achieve [36]), a crystal un- der 1 micron is sufficient for obtaining a good image of a protein as large as Rubisco (Figure 6a). Obtaining a usable diffraction pat- tern from a submicron crystal of MS2 virus is on the border of what is achievable, see Figure 6a. With pulses of 1012 photons (which is what LCLS is expected to deliver), the use of smaller crystals is possible, but there the damage puts a limit on the pulse length. The requirement of average ionization per atom of less than one will con- strain the pulse length to be shorter than 50 fs, or, alternatively, crystal sizes less than 160 nm (Figure 6c).

The calculation of the number of photons per Bragg peak (Equa- tion 5) presented in Figure 6 is valid only for crystals. This ap- proximation breaks down when the crystal size approaches the unit cell size, as the diffracted image becomes less discrete. It has been shown [9] that it is possible to reconstruct the structure of a Lysozyme in a 5×5×5 cluster to a resolution higher than 2 ˚A (assuming 9 pho- tons/pixel), using a 10 fs pulse with 5×1012photons in a 100 nm diameter spot size. Applying these pulse parameters to Equation 5 we can conclude that the calculations in the present work agree well with those of [9].

Conclusion

We present a theoretical study of the dynamics of the electrons gen- erated in a biological sample placed in a XFEL beam, based on sim- ulations of electron cloud development in urea crystals.

Due to the higher inelastic electron cross section at lower ener- gies [25] the secondary electron cascade caused by the Auger elec- tron is generated faster than the corresponding cascade from the pho- toelectron (Figure 2). The electron density associated with the Auger cloud is higher and more localized around the point where the initial electron was created (Figure 1). This leads inherently to two electron energy regimes and two electron cloud sizes, which occur simultane- ously in the sample during the pulse exposure.

When deciding which parameters of the X-ray laser pulse and sample characteristics one should use, there are mainly two effects at play which are driven mainly by the photoelectrons (Figure 3); i) Escape of the photoelectrons: if the photoelectrons escape the sam- ple the total number of ionizations in the sample will be significantly reduced. In other words, the sample has to be smaller than the to- tal size of the photoelectric cloud, to ensure that the total number of ionizations per atom is low (lower than one per atom). In this case, the length of the pulse plays little role in controlling the damage. ii)

(5)

Fig. 6.Integrated Bragg peak intensity for a reflection at 2 ˚A resolution (wavelength 1.5 ˚A) as a function of crystal size and unit cell dimensions for three different X-ray pulse intensities with (a) 1×1010photons, (b) 1×1011photons, and (c) 1×1012photons in a focal spot of 1µm diameter (FWHM). The solid line corresponds to a scattered signal of 9 photons in this Bragg peak, when peak degradation is not taken into account. Imaging is considered feasible if there are more than 9 photons/peak at 2 ˚A resolution.

For high photon intensities in (c), ionization is a limiting factor and constrains the crystals to be smaller and pulses shorter than 50 fs (dashed line).

Short pulses: the photoelectric cloud does not have time to develop to reach higher charges. In this case, size of the sample less rele- vant, as the charge density buildup is restrained to be low, and will be constrained by coherence requirements due to the pulse length.

The above considerations lead to the following scenario: For a photon flux of 1012photons per pulse, a pulse length of 50 fs comes out as a pivotal number. For pulses shorter than 50 fs, the free elec- tron density buildup is lower than our required threshold, meaning that one can investigate any crystal size (provided it produces a good enough signal for reconstruction). For pulses longer than 50 fs, en- forcing the threshold of 1 ionization per atom makes it necessary that the size of the crystal is smaller than the size of the photoelec- tric cloud after 50 fs, leading to a limit of the sample/crystal size of

∼160 nm, Figure 6c.

For a photon flux of 1011or 1010photons per pulse, ionization is not a limiting factor, since the level of one ionization per atom is never reached. The maximum number of electrons generated by the photo electron is around 390, see Figures 2 and 6. From the photoion- ization cross sections it can be deduced that less than every 500th atom will be photo ionized. This implies that imaging nanocrystals of small proteins with a small unit cell, such as Lysozyme, will be feasible at sources with lower photon fluxes.

In the case of 1013photons per pulse, the direct photoionization will ionize every third atom. This constrains the pulse length to be shorter than the average Auger decay time. For such intense pulses the pulse length must be below 5 fs, with submicron sample sizes.

Our calculations show that pulses of 1011 photons provide an ideal case for nanocrystal imaging. This number is within reach both by the imminent X-ray laser at LCLS as well as the upcoming table- top X-ray lasers, like the one under construction in Munich-Centre for Advanced Photonics. The near future holds the prospect of struc- tural determination from submicron protein crystals.

Materials and Methods

Simulations of the electron cascade dynamics in crystalline urea (CO(NH2)2) were performed using the spatial electron dynamics program,EHOLE, that is a part of theGROMACS [37] Molecular Dynamics software package. Urea was chosen as model for a biological sample due to three reasons: it has a well known crystalline structure [38], it has an atomic composition of biological character, and its unit cell is small. The urea crystal is also among the simplest crystalline organic materials known, with 16 atoms per unit cell in the tetragonal space groupP 421m. In earlier work [25], the inelastic electron cross sections for urea has been derived from first principles calculations. Based on these we have calculated the number of secondary electrons generated by an impact electron in a urea crystal. The inelastic cross section for electron scattering in urea is comparable in magnitude with that for water [25].

Thus, urea crystals are a good model for protein nanocrystals, known to contain 30%- 60% water. We refer to [24, 25, 39, 40] for further details on these calculations. The degradation of Bragg peaks in Figure 3 has been calculated from MD simulations on an urea crystal, usingGROMACS with stochastic interaction of X-ray photons with atoms [9]. The intensity of Bragg peaks is defined by integrating around each peak

over a rectangular area centered on the Bragg peak and with sides of length equal to 1/10 of the separation between adjacent peaks [41]. The spectral width∆λ/λis not taken into account here. The degradation of the Bragg peak is expected to be smaller when integrating through the thickness∆λ/λ2of the Ewald sphere.

We assume that the X-ray pulse can be described by a Gaussian centered at timet0= 0and will consider the wavelength of the incoming X-ray photons to be 1.5 ˚A. Following this pulse, several primary ionizations are treated- the photoelectric effect resulting in an ejection of a high energy electron (8 keV), accompanied by an Auger effect which provides an electron of a lower energy, depending on atomic species. The emission for these electrons is described by normalized probability distri- butions. The photoelectric effect is instantaneous so the emission probability follows the same Gaussian profile as the X-ray pulse. The probability for an Auger process to happen is then a convolution of a Gaussian with the exponential decay characteristic for each individual atomic species (see Figure 3a). The exponential decays are taken with corresponding life times of 11.3 fs for carbon, 8.3 fs for nitrogen and 6.6 fs for oxygen. The probability for photoionization in urea is determined by the cross sec- tion of the atoms, which are well known [27]. For the three atomic species that can undergo an Auger process, the contribution from the atoms C, N, and O, is weighted according to the photoionization cross section on the respective atoms,σC,σN,

σO, and normalized to the total photoelectric cross section for the urea molecule.

After calculating the dynamics of a secondary electron cascade from an incident electron with a specific energy (Figure 2), the entire electron cascade following an X-ray pulse impinging on a crystal can be calculated in the following manner. For pho- toelectrons, this is simply given by convoluting the emission rate with the evolution of the photoelectric cascade as given by our simulations (see Figure 3a). Similarly, for Auger electrons, the Auger emission rates are convoluted with the cascade following their impact on the crystal. Contribution from all atomic species (C, N, O) is consid- ered and weighted accordingly. The following equations were used, for describing the emission probability of photoelectrons

P (t) = N e(t−t0)

2

2w2 , [ 1 ]

and of Auger electrons

A(t) = N Z

dt0e(t0 −t0)

2

2w2 1

τe(t−t0 )τ , [ 2 ] wherewis the width of the pulse,τrepresents the Auger life time for a certain atomic species andNis a normalization constant which normalizes the entire probability to 1.

The secondary electron cascade from incident photoelectrons and associated Auger electrons (as shown in Figure 3), weighted for all atomic species in the sample, is given by the convolution

C(t) = X

i=C,N,O

niσi

Z

dt0N e(t0 −t0)

2

2w2 Cphoto(t, t0)+

+ Z

dt0N e(t0 −t0)

2

2w2 1

τe(t−t0 )τ CAuger(t, t0) ff

,

[ 3 ]

whereCphoto(t, t0)andCAuger(t, t0)represent the cascade development with time for a single electron starting from timet0. These cascades are obtained from MD simualtions and are represented by the ionization rate as a function of time (Figure 2a fort0= 0), or radii of gyration, Figure 2b and c.

(6)

The radius of gyration, used in Figure 3, is described by

Rg(t) =

„ P

iri(t)2mi

P

imi

«1/2

, [ 4 ]

wheremiis the mass of electroniandrithe position of electroniwith respect to the center of mass of all free electrons.

Through calculations similar to those in [42], the average number of photons scattered elastically by a protein crystal within a Bragg peak centered on the direc- tion0, ϕ0)can be approximated by the expression

Z Z

I(θ, ϕ) dΩ ≈ Φr2eλ2A4 a6

X

atoms

fatom2 0)1 + cos20)

2 , [ 5 ]

whereis the solid angle spanned by the Bragg peak,Φthe fluence of the incoming X-ray beam,rethe classical electron radius,λthe wavelength,Athe crystal width,

athe unit cell width,fatomthe atomic scattering factor, andθ0 the polar angle between the incident pulse and the center of the Bragg peak. It is assumed that the unit cell structure factor is constant within the Bragg peak, and that adjacent Bragg peaks do not overlap; both these approximations improve with the ratioA/a. The squared structure factor of the unit cell is represented by its average value at high scattering angles [43]. For numerical evaluation, the unit cell was assumed to have a density 1/30 ˚A−3carbon-equivalent atoms (corresponding to a unit cell consisting of 50% non-structural water and protein with density approximately 1.35 g/cm3), and the scattering factor of carbon was calculated from the Cromer-Mann parameters [44].

ACKNOWLEDGMENTS. The Swedish Research Council is acknowledged for finan- cial support as well as the DFG Cluster of Excellence: Munich-Centre for Advanced Photonics. The authors would like to thank Janos Hajdu, Magnus Bergh, Gerard Kleywegt, Inger Andersson, Richard Neutze, Martin Svenda, Rosmarie Friemann and Karin Valeg˚ard, for their valuable input.

References

1. Ackermann W, et al. (2007) Operation of a free-electron laser from the extreme ultraviolet to the water window. Nat. Photonics 1:336–342.

2. DiMauro LF, et al. (2007) Progress report on the LCLS XFEL at SLAC. J. Phys.: Conf. Ser.

88:012058.

3. Shintake T, et al. (2008) A compact free-electron laser for generating coherent radiation in the extreme ultraviolet region. Nat. Photonics 2:555–559.

4. Geddes CGR, et al. (2004) High-quality electron beams from a laser wakefield accelerator using plasma-channel guiding. Nature 431:538–541.

5. Mangles SPD, et al. (2004) Monoenergetic beams of relativistic electrons from intense laser- plasma interactions. Nature 431:535–538.

6. Faure J, et al. (2004) A laser-plasma accelerator producing monoenergetic electron beams.

Nature 431:541–544.

7. Sandberg RL, et al. (2008) High numerical aperture tabletop soft x-ray diffraction microscopy with 70-nm resolution. Proc. Natl. Acad. Sci. U.S.A. 105:24–27.

8. Gaffney KJ, Chapman HN (2007) Imaging atomic structure and dynamics with ultrafast x-ray scattering. Science 316:1444–1448.

9. Neutze R, Wouts R, van der Spoel D, Weckert E, Hajdu J (2000) Potential for biomolecular imaging with femtosecond X-ray pulses. Nature 406:752–757.

10. Hajdu J (2000) Single-molecule x-ray diffraction. Curr. Opin. Struct. Biol. 10:569–573.

11. Neutze R, Huldt G, Hajdu J, van der Spoel D (2004) Potential impact of an X-ray free electron laser on structural biology. Radiat. Phys. Chem. 71:905–916.

12. Boutet S, Robinson I (2008) Coherent x-ray diffractive imaging of protein crystals. J. Syn- chrotron Rad. 15:576–583.

13. Chapman HN (2009) X-ray imaging beyond the limits. Nat. Mater. 8:299–301.

14. Hau-Riege SP, et al. (2007) Subnanometer-scale measurements of the interaction of ultrafast soft x-ray free-electron-laser pulses with matter. Phys. Rev. Lett. 98:145502.

15. Hau-Riege SP, et al. (2007) Damage threshold of inorganic solids under free-electron-laser irradiation at 32.5 nm wavelength. Appl. Rev. Lett. 90:173128.

16. Chapman HN, et al. (2007) Femtosecond time-delay x-ray holography. Nature 448:676–680.

17. Chalupsk´y J, et al. (2007) Characteristics of focused soft x-ray free-electron laser beam deter- mined by ablation of organic molecular solids. Opt. Expr. 15:6036–6043.

18. Bergh M, Tˆımneanu N, van der Spoel D (2004) A model for the dynamics of a water cluster in a X-ray FEL beam. Phys. Rev. E 70:051904.

19. Hau-Riege SP, London RA, Sz˝oke A (2004) Dynamics of biological molecules irradiated by short x-ray pulses. Phys. Rev. E 69:051906.

20. Jurek Z, Faigel G, Tegze M (2004) Dynamics in a cluster under the influence of intense femtosecond hard X-ray pulses. Eur. Phys. J. D 29:217–229.

21. Shapiro DA, et al. (2008) Powder diffraction from a continuous microjet of submicrometer protein crystals. J. Sync. Rad. 15:593–599.

22. Hau-Riege SP (2008) Effect of the coherence properties of self-amplified-spontaneous- emissionx-ray free electron lasers on single-particle diffractive imaging. Opt. Express 16:2840–

2844.

23. Ziaja B, London RA, Hajdu J (2005) Unified model of secondary electron cascades in diamond.

J. Appl. Phys. 97:064905.

24. Ortiz C, Caleman C (2007) Secondary electron cascade dynamics in KI and CsI. J. Phys.

Chem. C. 111:17442–17447.

25. Caleman C, et al. (2009) Radiation damage in biological material: electronic properties and electron impact ionization in urea. Europhys. Lett. 85:18005.

26. Nave C, Hill MA (2005) Will reduced radiation damage occur with very small crystals? J.

Synchrotron Rad. 12:299–303.

27. Berger M, et al. (1998) XCOM: photon cross sections database.

(http://physics.nist.gov/PhysRefData/Xcom/Text/XCOM.html).

28. Fung R, Shneerson V, Saldin DK, Ourmazd A (2009) Structure from fleeting illumination of faint spinning objects in flight. Nat. Phys. 5:64–67.

29. Miao J, Hodgson KO, Sayre D (2001) An approach to three-dimensional structures of biomolecules by using single-molecule diffraction images. Proc. Natl. Acad. Sci. U.S.A.

98:6641–6645.

30. Valeg˚ard K, et al. (2004) Deacetoxycephalosporin c synthase complexed with 2-oxoglutarate and ampicillin. Nat. Struct. Biol. 11:95–10.

31. Huldt G, Sz˝oke A, Hajdu J (2003) Diffraction imaging of single particles and biomolecules. J.

Struct. Biol. 144:219–227.

32. Hau-Riege SP, London RA, Chapman HN, Szoke A, Tˆımneanu N (2007) Encapsulation and diffraction-pattern-correction methods to reduce the effect of damage in x-ray diffraction imag- ing of single biological molecules. Phys. Rev. Lett. 98:198302.

33. Sauter C, et al. (2001) Structure of tetragonal hen egg-white lysozyme at 0.94 ˚a from crystals grown by the counter-diffusion method. Acta Cryst. D 57:1119–1126.

34. Knight S, Andersson I, Branden C (1990) Crystallographic analysis of ribose 1,5-biphosphate carboxylase from spinach at 2.4 ˚a resolution - subunit interactions and active-site. J. Mol. Biol.

215:113–160.

35. Valeg˚ard K, Liljas L, Fridborg K, Unge T (1990) The 3-dimensional structure of the bacterial- virus MS2. Nature 345:36–41.

36. G¨unter F, et al. (2007) Design considerations for table-top, laser-based VUV and X-ray free electron lasers. Appl. Phys. B 86:431–435.

37. van der Spoel D, et al. (2005) GROMACS: Fast, Flexible and Free. J. Comp. Chem. 26:1701–

1718.

38. Pryor AW, Sanger PL (1970) Collection and interpretation of neutron diffraction measurements on urea. Acta Cryst. A 26:543–558.

39. Tˆımneanu N, Caleman C, Hajdu J, van der Spoel D (2004) Auger electron cascades in water and ice. Chem. Phys. 299:277–283.

40. Ziaja B, van der Spoel D, Sz˝oke A, Hajdu J (2001) Radiation induced electron cascade in diamond and amorphous carbon. Phys. Rev. B 64:214104.

41. van der Spoel D, Maia FRNC, Caleman C (2008) Structural studies of melting on the picosec- ond time scale. Phys. Chem. Chem. Phys. 10:6344–6349.

42. James RW (1948) The optical principles of the diffraction of X-rays (Oxbow press, Wood- bridge).

43. Wilson AJC (1949) The probability distribution of X-ray intensities. Act. Cryst. 2:318–321.

44. Cromer DT, Mann JB (1968) X-ray scattering factors computed from numerical Hartree-Fock wave functions. Acta Cryst. A 24:321–326.

References

Related documents

X-ray, free-electron laser, XFEL, diffraction analysis, structure determination, nanocrystal, molecular dynamics, GROMACS, biomolecular imaging, ubiquitin, trajectory, explosion

Re-examination of the actual 2 ♀♀ (ZML) revealed that they are Andrena labialis (det.. Andrena jacobi Perkins: Paxton & al. -Species synonymy- Schwarz & al. scotica while

Our results indicate that irradiating a solid target with a pulse from the existing FLASH facility will not heat it enough, but that molecular cluster targets at the coming hard

This report describes the first X-ray Free Electron Laser (FEL) imaging experiments at the Linac Coherent Light Source, at Stanford Linear Accelerator Centre, where a number of

A study on the types of forces that the particles are subjects to during the sample injection and delivery process could also help in reducing the damages by these forces on

• We have combined Monte Carlo simulation of the X-ray interaction in a scintillator coated CMOS pixel detector with advanced electrical device simulation of the

Each particle presented to the x-ray free electron laser (XFEL) beam can only be measured once, in a single orientation.. This means it will give structural information about a

The cascaded theoretical model can be considered generic for hybrid detectors and be evaluated for different X-ray inputs, sensor materials, charge diffusion data, pixel sizes