research papers\(\def\hfill{\hskip 5em}\def\hfil{\hskip 3em}\def\eqno#1{\hfil {#1}}\)

Journal logoJOURNAL OF
SYNCHROTRON
RADIATION
ISSN: 1600-5775

Diffraction of X-ray free-electron laser femtosecond pulses on single crystals in the Bragg and Laue geometry

aM. V. Lomonosov Moscow State University, 119992 GSP-2 Moscow, Russia
*Correspondence e-mail: vabushuev@yandex.ru

(Received 25 September 2007; accepted 3 July 2008; online 22 July 2008)

A solution of the problem of dynamical diffraction for X-ray pulses with arbitrary dimensions in the Bragg and Laue cases in a crystal of any thickness and asymmetry coefficient of reflection is presented. Analysis of pulse form and duration transformation in the process of diffraction and propagation in a vacuum is conducted. It is shown that only the symmetrical Bragg case can be used to avoid smearing of reflected pulses.

1. Introduction

In the near future, X-ray free-electron lasers (XFELs) with wavelength λ ≃ 0.1 nm will become available for a wide community of users. Therefore it is of interest to consider dynamical diffraction as a tool for controlling and tailoring parameters of femtosecond pulses and for developing methods of X-ray laser pulse diagnostics. Three XFEL projects are now actively developed: the European XFEL Facility in Germany (Altarelli et al., 2006[Altarelli, M. et al. (2006). Editors. Report DESY 2006-097. DESY, Hamburg, Germany. (http://xfel.desy.de/tdr/index_eng.html .)]), the LCLS (Linac Coherent Light Source) in the USA (Arthur et al., 2002[Arthur, J. et al. (2002). LCLS Conceptual Design Report. LCLS, USA. (http://www-ssrl.slac.stanford.edu/lcls/cdr/ .)]) and the SCSS (SPring-8 Compact SASE Source) in Japan (Tanaka & Shintake, 2005[Tanaka, T. & Shintake, T. (2005). SCSS X-FEL Conceptual Design Report, edited by Takashi Tanaka and Tsumoru Shintake. SCSS XFEL, R&D Group, RIKEN Harima Institute/SPring-8, Japan. (http://www-xfel.spring8.or.jp/SCSSCDR.pdf .)]). In these machines, X-ray bunches of duration ∼100–200 fs leave an undulator as a result of self-amplification of spontaneous radiation of 15 GeV electrons. Theoretical calculations show that these pulses will have an irregular multiple-peak internal structure and consist of several hundreds of supershort independent sub-pulses of duration τ0 ≃ 0.1 fs, separated by time intervals Δt ≃ 0.3–0.5 fs. A typical pulse has a transversal size r0 ≃ 50 µm at the undulator exit, angular divergence ≃ 1 µrad, peak power ≃ 10 GW and average power ≃ 40 W (Saldin et al., 2004[Saldin, E. L., Schneidmiller, E. A. & Yurkov, M. V. (2004). Report TESLA-FEL 2004-02. DESY, Hamburg, Germany.]).

The analysis of diffraction of XFEL radiation has been restricted so far to the approximation of a plane (unlimited) wavefront for the Bragg case (Chukhovskii & Förster, 1995[Chukhovskii, F. N. & Förster, E. (1995). Acta Cryst. A51, 668-672.]; Shastri et al., 2001a[Shastri, S. D., Zambianchi, P. & Mills, D. M. (2001a). J. Synchrotron Rad. 8, 1131-1135.],b[Shastri, S. D., Zambianchi, P. & Mills, D. M. (2001b). Proc. SPIE, 4143, 69-77.]; Graeff, 2004[Graeff, W. (2004). J. Synchrotron Rad. 11, 261-265.]) and for the Laue case (Shastri et al., 2001b[Shastri, S. D., Zambianchi, P. & Mills, D. M. (2001b). Proc. SPIE, 4143, 69-77.]; Graeff, 2002[Graeff, W. (2002). J. Synchrotron Rad. 9, 82-85.]; Malgrange & Graeff, 2003[Malgrange, C. & Graeff, W. (2003). J. Synchrotron Rad. 10, 248-254.]). The time structure of the incident pulse has been approximated either by a δ function (Chukhovskii & Förster, 1995[Chukhovskii, F. N. & Förster, E. (1995). Acta Cryst. A51, 668-672.]; Shastri et al., 2001a[Shastri, S. D., Zambianchi, P. & Mills, D. M. (2001a). J. Synchrotron Rad. 8, 1131-1135.],b[Shastri, S. D., Zambianchi, P. & Mills, D. M. (2001b). Proc. SPIE, 4143, 69-77.]; Graeff, 2002[Graeff, W. (2002). J. Synchrotron Rad. 9, 82-85.]; Malgrange & Graeff, 2003[Malgrange, C. & Graeff, W. (2003). J. Synchrotron Rad. 10, 248-254.]) or by a Gaussian (Shastri et al., 2001a[Shastri, S. D., Zambianchi, P. & Mills, D. M. (2001a). J. Synchrotron Rad. 8, 1131-1135.]; Graeff, 2004[Graeff, W. (2004). J. Synchrotron Rad. 11, 261-265.]). Although giving some insight into the physics of short-pulse diffraction, such an approach cannot in principle take into account the presence of transverse mode structure and, even more essential, a non-uniform distribution of the field phase inside a pulse. However, such a phase distribution will inevitably arise at large, of the order of 100 m (Saldin et al., 2004[Saldin, E. L., Schneidmiller, E. A. & Yurkov, M. V. (2004). Report TESLA-FEL 2004-02. DESY, Hamburg, Germany.]), distances from the undulator to the sample or monochromator crystal. Besides, all analysis so far has been limited to the reflected pulse field on the exit surface of a crystal, whereas significant practical interest is for spatial (transversal) and temporal (longitudinal) smearing of pulses during their further propagation in vacuum.

In the present article a general theory of dynamical diffraction of X-ray pulses with an arbitrary spatial and temporal structure, described by a field Ein(r, t), on crystals with arbitrary thickness and asymmetry coefficient in the Bragg and in the Laue cases is developed. Such an approach allows us to analyse the structure of fields Eg(r, t) of forward-diffracted (transmitted, g = 0) and diffracted (reflected, g = h) pulses at any distance from the crystal, and also the degree of space and time coherence of these pulses and their relation with the statistical properties of the XFEL radiation field.

2. Theory

We shall consider diffraction reflection and transmission of a pulse of X-ray radiation Ein(r, t) = Ain(r, t) exp(iK0riω0t), which is incident on a single-crystal plate of thickness d. The field on the entrance crystal surface z = 0 can be written as

[E_{\rm{in}}(x,t)=A_{\rm{in}}(x,t)\exp(iK_{0x}x-i\omega_0t),\eqno(1)]

where Ain(x, t) is a slowly varying complex amplitude (the envelope of a wave packet), K0x = K0sinθ0, K0 = ω0/c = 2π/λ and c is the light speed in a vacuum; the axis x is directed along the crystal surface and the axis z is directed inside the crystal along the normal n to the surface (Fig. 1[link]). The projection of the incident wavevector on the axis z is K0z = K0γ0, where γ0 = cos(K0·n) = cosθ0. The angle of incidence of the radiation to the normal n is θ0 = ψθBΔθ, where θB is the Bragg angle for the central (average) frequency ω0, which is determined by the expression 2K0sinθB = h, where h is the modulus of the reciprocal lattice vector h = (hcosψ, −hsinψ), Δθ is the angular deviation from the exact Bragg angle, which is determined by the expression K0h = −K0hsin(θB + Δθ), and ψ is the inclination angle of reflecting crystal planes to the normal n. The restriction |ψθB| < π/2 on angle ψ follows from the condition γ0 > 0. Representation of a pulse by the form (1)[link] is correct as long as the characteristic cross-section size of a pulse r0 [\gg] λ, and its duration τ0 [\gg] λ/c.

[Figure 1]
Figure 1
Geometry in real space of X-ray pulse diffraction in the Bragg and Laue cases. 1, incident pulse Ein(r, t); 2, transmitted pulse E0(r, t); 3, reflected pulse Eh(r, t) in the Bragg case; 4, reflected pulse in the Laue case; 5, crystal; θ0 and θh are the angle of incidence of the initial pulse and the angle of reflection of the diffracted pulse, respectively, relative to the axis z.

Let us now write the field Ein(x, t) (1)[link] in the form of a two-dimensional Fourier integral,

[E_{\rm{in}}(x,t)=\textstyle\int\!\!\int{E}_{\rm{in}}(k_{0x},\omega)\exp(ik_{0x}x-i\omega{t})\,{\rm{d}}k_{0x}\,{\rm{d}}\omega,\eqno(2)]

where

[E_{\rm{in}}(k_{0x},\omega)=(2\pi)^{-2}\textstyle\int\!\!\int{E}_{\rm{in}}(x,t)\exp(-ik_{0x}x+i\omega{t})\,{\rm{d}}x\,{\rm{d}}t.\eqno(3)]

Here and further on, all integrations are carried out over the infinite limits from −∞ to +∞. Substituting the field Ein(x, t) (1)[link] into (3)[link] and introducing new variables

[q=k_{0x}-K_{0x},\quad\quad\Omega=\omega-\omega_0,\eqno(4)]

one obtains a set of Fourier amplitudes of the field, Ein(k0x, ω) = Ain(q, Ω), with

[A_{\rm{in}}(q,\Omega)=(2\pi)^{-2}\textstyle\int\!\!\int{A}_{\rm{in}}(x,t)\exp(-iqx+i\Omega{t})\,{\rm{d}}x\,{\rm{d}}t.\eqno(5)]

Expression (2)[link] describes a set of plain monochromatic waves with amplitudes Ain(q, Ω), wavevectors k0 = (k0x, k0z) and frequencies ω, where k0x = K0x + q, k0z = (k02k0x2)1/2 and k0 = (ω0 + Ω)/c, which are incident on a crystal surface. In accordance with the known results of the plane-wave dynamical theory of X-ray diffraction, each single component wave in (2)[link] is transmitted and reflected with the amplitude coefficients of transmission T(q, Ω) and reflection R(q, Ω). As a result we shall obtain the distribution of fields Eg(x, z, t) for transmitted (g = 0) and reflected (g = h) pulses at any point of space (x, z) outside the crystal and at any moment of time t,

[E_g({\bf{r}},t)=\textstyle\int\!\!\int{B_g}(q,\Omega)A_{\rm{in}}(q,\Omega)\exp(i{\bf{k}}_g{\bf{r}}-i\omega{t})\,{\rm{d}}q\,{\rm{d}}\Omega,\eqno(6)]

where B0 = T, Bh = R.

Here it is taken into account that owing to a condition of continuity of the tangential components of the wavevectors at the entrance and exit crystal surfaces, the values of projections of wavevectors kg in a vacuum will take the following form,

[k_{gx}=K_{gx}+q,\quad\quad k_{gz}=\sigma_g(k_0^2-k_{gx}^2)^{1/2},\eqno(7)]

where Kgx = K0x + gx, g = 0, h; σ0,h = 1 in the Laue case; σ0 = 1 and σh = −1 in the Bragg case; z ≤ 0 for a reflected pulse in the Bragg case and zd in the Laue case and for a transmitted pulse in the Bragg case. Contrary to the usual notation of wavevectors, namely denoting wavevectors in a vacuum by Kg and in the crystal by kg, Kg denotes the average wavevectors of the pulses and kg takes into account the q- and ω-spectra of the incident, reflected and transmitted pulses. Throughout this paper all wavevectors are restricted to a vacuum.

Note that the Fourier-transform-based approach, used here, is more simple and productive in comparison with the time-dependent Takagi–Taupin differential equations used by Chukhovskii & Förster (1995[Chukhovskii, F. N. & Förster, E. (1995). Acta Cryst. A51, 668-672.]), Wark & He (1994[Wark, J. S. & He, H. (1994). Laser Part. Beams, 12, 507-513.]) and Wark & Lee (1999[Wark, J. S. & Lee, R. W. (1999). J. Appl. Cryst. 32, 692-703.]).

Representing the square root (7)[link] in the form of a series over small parameters q/K0 and Ω/ω0, which is truncated discarding terms of the third order and higher, and substituting this result into the two-dimensional integral (6)[link], we obtain a general expression for the electric fields of X-ray pulses [see Appendix A[link], equations (30)[link] and (31)[link]],

[E_g({\bf{r}},t)=A_g({\bf{r}},t)\exp(i{\bf{K}}_g{\bf{r}}-i\omega_0t),\eqno(8)]

where Kgx = K0x + gx = K0sinθg, Kgz = σg(K02Kgx2)1/2 = K0γg, γg = cosθg. The angle of diffraction reflection with respect to the crystal normal is θh = ψ + θBbΔθ, where b = γ0/γh is the asymmetry coefficient of the reflection. In the Bragg case, γh < 0, b < 0, and angle ψ must satisfy the condition |ψπ/2| < θB. The slow-varying amplitudes are

[A_g(x,z,t)=\textstyle\int\!\!\int{B_g}(q,\Omega)A_{\rm{in}}(q,\Omega)\exp(iS_g+iD_g)\,{\rm{d}}q\,{\rm{d}}\Omega,\eqno(9)]

where

[S_g(q,\Omega)=q(x-\tan\theta_gz)-\Omega(t-z/c\gamma_g),\eqno(10)]

[D_g(q,\Omega)=-\left[q-(\Omega/c)\sin\theta_g\right]^2z/(2K_0\gamma_g^3).\eqno(11)]

The phase Sg (10)[link] determines the displacement of pulse centres in x and t with distance z from the crystal. The phase Dg (11)[link], which is quadratic in [q − (Ω/c)sinθg] and proportional to z, describes the curvature of the wavefront and the diffraction smearing of pulses during their propagation in a vacuum. It is necessary to take into account the terms of the order of ∼q2, Ω2 and qΩ to obtain a correct solution of expression (7)[link] and to analyse diffraction broadening of pulses in space and in time. In all previous articles this extremely important aspect was not taken into account. Expression (11)[link] describes the effect of the curvature of the asymptotes of the dispersion surface far away from the reflecting crystal for pulses limited in time and in space. The influence of curved asymptotes has been considered theoretically earlier, for example, by Bauspiess et al. (1976[Bauspiess, W., Bonse, U. & Graeff, W. (1976). J. Appl. Cryst. 9, 68-80.]) in the case of the incident spherical X-ray or neutron waves on the interferometer. Integral (9)[link] is equivalent to the integral formula of Kirchhoff–Helmholtz, generalizing the Huygens–Fresnel principle, since in the quasi-optical approximation the spherical wavefront of point sources can be replaced by a parabolic wavefront, which is justified in the paraxial region.

Let us write the reflection coefficient R(q, Ω) and transmission coefficient T(q, Ω) in (6)[link] and in (9)[link] for a crystal with any thickness d in the general form. In the Bragg case,

[\eqalign{&R=(R_1-pR_2)/(1-p),\cr&T=[\exp(i\varphi_1)-p\exp(i\varphi_2)]/(1-p),}\eqno(12)]

where

[\eqalign{&R_{1,2}=(\alpha_1\pm Q)/2C\chi_{\bar h},\quad\quad p=(R_1/R_2)\exp[i(\varphi_1-\varphi_2)],\cr& \varphi_{1,2}=k_0\varepsilon_{1,2}d,\quad\quad \varepsilon_{1,2}=(2\chi_0 +\alpha_1\pm Q)/4\gamma_0,\cr&\alpha_1=\alpha{b}-\chi_0(1-b),\quad\quad Q=(\alpha_1^2 + 4bC^2 \chi_h\chi_{\bar h})^{1/2}.}]

Here χg are the Fourier components of the dielectric susceptibility of the crystal; C = 1 and C = cos2θB for σ- and π-polarized radiation, respectively. Parameter α = [k02 − (k0 + h)2]/k02, determining the deviation from the exact Bragg condition, has the form

[\eqalignno{\alpha(q,\Omega)={}&2\sin2\theta_{\rm{B}}\left[\Delta\theta-q/K_0\gamma_0\right.\cr&\left.+\,(\Omega/\omega_0)\sin\psi/\gamma_0\cos\theta_{\rm{B}}\right],&(13)}]

where Δθ is the departure of the incident pulse from the Bragg angle.

In the Laue case,

[B_g=A_g^{(1)}\exp(i\varphi_1)+A_g^{(2)}\exp(i\varphi_2),\eqno(14)]

where

[A_0^{(1,2)}=(Q\mp\alpha_1)/2Q,\quad\quad A_h^{(1,2)}=\pm Cb\chi_h/Q.]

The characteristic angular width ΔθB of the diffraction reflection coefficients R (12)[link] and R = Bh (14)[link] depends on the ratio between the thickness of the crystal d and the extinction length Λ = λ(γ0|γh|)1/2/πC|χh|. In the case of a thick crystal (d [\gg] Λ) this width is equal to ΔθB = C|χh|/|b|1/2sin2θB, or, in the other designations, ΔθB = λ|γh|/πΛsin2θB. In the case of a thin crystal, when dΛ, the angular width of the reflection is ΔθBλ|γh|/πdsin2θB. The reflection coefficient in the Laue case is maximal if the crystal thickness satisfies the condition d = (π/2)Λ(1 + 2n), where n = 0, 1, 2,….

The intensities of the transmitted and reflected pulses are determined by the expression Ig(x, z, t) = |Ag|2. It is easy to show that the total energy of a pulse, Wg = [\textstyle\int\!\!\int]Igdxdt, does not depend on the distance z and the time t, which means conservation of energy during the pulse propagation in a vacuum,

[W_g=(2\pi)^2\textstyle\int\!\!\int|B_g(q,\Omega)|^2|A_{\rm{in}}(q,\Omega)|^2\,{\rm{d}}q\,{\rm{d}}\Omega.]

As an example, we shall further consider everywhere a Gaussian incident pulse,

[\eqalignno{A_{\rm{in}}(x,t)={}&\exp\left[-(x\gamma_0/r_0)^2+i\varphi_0(x)\right.\cr&\left.-\,(t-x\sin\theta_0/c)^2/\tau_0^2\right],&(15)}]

where r0 and τ0 are the transverse size and the pulse duration of the incident pulse, respectively, φ0(x) = α0(xγ0/r0)2 is the phase, and the parameter α0 is equal to the phase at |x| = r0/γ0 [see Appendix B[link], expression (45)[link]].

Depending on the ratio between r0 and τ0 it is possible to introduce the concept of a long pulse, for which pulse duration τ0 [\gg] (r0/c)tanθ0, and a short pulse with wide front with r0 [\gg] cτ0cotanθ0. In the first case, only a limited area of the crystal surface with |x| ≤ x0 = r0/γ0 is involved in the scattering, whereas in the second case the incident pulse with size Δxcτ0/sinθ0 [\ll] x0 propagates along the crystal surface with speed c/sinθ0, higher than the speed of light in a vacuum. It is the latter situation that will be realised for femtosecond pulses.

For narrow and short pulses the angular divergence Δθ0λ/πr0 and spectral width ΔΩ0 ≃ 2/τ0 are comparable or even exceed the angular width ΔθB and spectral width ΔΩB = ΔθBω0cotanθB of a Bragg reflection. This leads to a sharp change in the form and to reduction of intensity of a reflected pulse, but also to its smearing in time as well as in space. The degree of smearing in the general case increases with the distance z (see §3[link]).

3. Results and discussion

Let us explore some examples of typical pulse parameters. We shall always consider the 220 reflection of σ-polarized radiation with λ = 0.154 nm from a silicon single crystal at a departure angle Δθ = (1 − b)Re(χ0)/2sin2θB, which corresponds to the maximal reflected intensity, where θB = 23.65°. In the symmetric Bragg case (b = −1), the Bragg width for a thick crystal ΔθB = 12.4 µrad, and the extinction length Λ = 2.16 µm. In the symmetric Laue case (b = 1), Λ = 4.92 µm. From a general point of view it is clear that for the reduction of heat absorption the thickness of a crystal should best be chosen small (d ≤ 1–3Λ), but at the same time large enough to provide sufficiently high X-ray reflection coefficient values.

The expressions (8)[link]–(14)[link] give a general solution of the problem of transmission and reflection of X-ray pulses in the Bragg and Laue cases. Let us discuss some special cases. If the field amplitude Ain does not depend on x and t, then, in agreement with expressions (5)[link] and (13)[link], Ain(q, Ω) = δ(q)δ(Ω), α = 2Δθsin2θB, and we find the well known result for a plane monochromatic wave: Ag = Bg(Δθ). Formally this means that in (15)[link] one should assume r0 → ∞ and τ0 → ∞. In a real situation the approximation of a plane monochromatic wave will be realised at Δθ0 [\ll] ΔθB and ΔΩ0 [\ll] ΔΩB, i.e. in the case of a source of size rs [\gg] λ/(πΔθB) [see expression (44)[link] at αs = 0] and of pulse duration τ0 [\gg] 2/ΔΩB. For example, for λ = 0.154 nm in the case of symmetric Bragg reflection Si(220) this leads to the following requirements: r0 [\gg] 4 µm, τ0 [\gg] 6 fs.

For a less restrictive approximation of a monochromatic X-ray beam, the amplitude Ain(x) depends only on one coordinate, i.e. τ0 → ∞ in equation (15)[link]. In this case, Ain(q, Ω) = Ain(q)δ(Ω), α = 2sin2θB(Δθq/K0γ0) and

[A_g(x,z)=\textstyle\int{B_g}(q)A_{\rm{in}}(q)\exp(iS_g+iD_g)\,{\rm{d}}q,\eqno(16)]

where

[S_g=q(x-\tan\theta_gz),\quad\quad D_g=-q^2z/(2K_0\gamma_g^3).]

Expression (16)[link], which is valid at any z and d, is more general in comparison with that obtained earlier using the Green function method for diffraction reflection of a limited X-ray beam from a semi-infinite crystal at z = 0 in the Bragg case (Afanas'ev & Kohn, 1971[Afanas'ev, A. M. & Kohn, V. G. (1971). Acta Cryst. A27, 421-430.]) and at z = d in the Laue case (Slobodetzkii & Chukhovskii, 1970[Slobodetzkii, I. Sh. & Chukhovskii, F. N. (1970). Kristallografiya, 15, 1101-1107. (In Russian.)]).

It is of further interest to analyse diffraction of a short pulse with a wide wavefront when the longitudinal size of the incident pulse l0 = cτ0 [\ll] r0. In this case it is possible to neglect boundary effects, i.e. to exclude the field dependence on the incident pulse transverse coordinate. Then Ain(x, t) = Ain(txsinθ0/c) and, in agreement with (5)[link], Ain(q, Ω) = Ain(Ω)δ(qΩsinθ0/c). As a result, from (9)[link] it is easy to show that

[A_g(x,z,t)=\textstyle\int{B_g}(\Omega)A_{\rm{in}}(\Omega)\exp(iS_g+iD_g)\,{\rm{d}}\Omega,\eqno(17)]

where

[S_g=(\Omega/c)\left[\sin\theta_0x+(1-\sin\theta_0\sin\theta_g)z/\gamma_g\right]-\Omega{t},\eqno(18)]

[D_g=-\Omega^2F_gz,\quad\quad F_g=(\sin\theta_0-\sin\theta_g)^2/(2K_0c^2\gamma_g^3).\eqno(19)]

The substitution of expression q = Ωsinθ0/c into (13)[link] results in a known expression for the value of α in the case of an incident non-monochromatic plane wave,

[\alpha(\Omega)=2\sin2\theta_{\rm{B}}[\Delta\theta+(\Omega/\omega_0)\tan\theta_{\rm{B}}].\eqno(20)]

For convenience, the analysis of the space and time structure of amplitudes Ag(x, z, t) (17)[link] can be carried out in a new Cartesian system of coordinates (x[_g^{\,\prime}], z[_g^{\,\prime}]) with transition rules x = x[_g^{\,\prime}]cosφ[_g^{\,\prime}] + z[_g^{\,\prime}]sinφ[_g^{\,\prime}], z = z[_g^{\,\prime}]cosφ[_g^{\,\prime}]x[_g^{\,\prime}]sinφ[_g^{\,\prime}], in which the axis z[_g^{\,\prime}] makes an angle φ[_g^{\,\prime}] with the crystal normal n. This angle φ[_g^{\,\prime}] is selected in such a way that the phase Sg (18) becomes independent of the transversal coordinate x[_g^{\,\prime}]. From (18) it is easy to obtain

[\tan\varphi_g^\prime=\gamma_g\sin\theta_0/(1-\sin\theta_0\sin\theta_g).\eqno(21)]

Note that the axes of coordinates (x[_h^{\,\prime}], z[_h^{\,\prime}]) and (xp, zp) in Appendix C[link] are parallel to each other; however, the system (xp, zp) moves with the speed of light in a vacuum along the direction of the wavevector Kh [see Fig. 8 and the formulae (53)[link], (55)[link]]. It is easy to be convinced that the angle φ[_h^{\,\prime}] = θh + φh, where the angle φh is defined from equation (56)[link].

In the new coordinate system the phase Sg = −Ω(tz[_g^{\,\prime}]/Vg), where Vg is the speed of the pulse along the longitudinal axis z[_g^{\,\prime}],

[V_g=c|\gamma_g|/(1-2\sin\theta_0\sin\theta_g+\sin^2\theta_0)^{1/2}.\eqno(22)]

From expressions (19)[link], (21)[link] and (22)[link] it follows that, for transmitted pulses (g = 0) both in the Bragg case and in the Laue case, φ0′ = θ0, V0 = c and D0 = 0. In other words the incident pulse with a wide front is transmitted along its initial direction K0 with the speed of light and is not deformed during the pulse transmission in a vacuum, i.e. remains a plane non-monochromatic wave with time dependence A0(tz0′/c), which differs in the general case from Ain(tz0′/c).

Quite a different situation takes place for the reflected pulses. In the general case the directions of the wavevector Kh and the normal N to the pulse do not coincide (see Appendix C[link] and Fig. 8). This is stipulated by the fact that at a fixed incidence angle θ0 various spectral components of a field Ah(Ω) are reflected under different angles θh(Ω) to the crystal normal. As long as the reciprocal lattice vector h in kh (7)[link] has a non-zero projection hx ≠ 0 along the x axis, part of the longitudinal impulse of the wavevector kh is transferred to the crystal and the angle of reflection θh(Ω) is different for various spectral components Ω: θh(Ω) = θh + Δθh(Ω), where

[\Delta\theta_h(\Omega)=-2(\Omega/\omega_0)\sin\theta_{\rm{B}}\cos\psi/\gamma_h.]

Superposition of these plane waves gives rise to non-trivial propagation of the reflected pulse in a vacuum. Earlier the speed Vh (22)[link] was not quite correctly named as the `group velocity' (Malgrange & Graeff, 2003[Malgrange, C. & Graeff, W. (2003). J. Synchrotron Rad. 10, 248-254.]). For a more detailed discussion of this, see Appendix C[link].

The only exception is the symmetric Bragg case (b = −1), for which ψ = π/2, sinθ0 = sinθh, Δθh = 0, Vh = c and Dh = 0, i.e. smearing of pulses in a vacuum does not take place. If |b| ≠ 1 in the Bragg case, and in any Laue case, Vh < c and the form of the pulse Ah(tz[_h^{\,\prime}]/Vh) varies during the propagation in a vacuum (Figs. 3–7). For the Laue case, expressions (21)[link] and (22)[link] were given earlier by Malgrange & Graeff (2003[Malgrange, C. & Graeff, W. (2003). J. Synchrotron Rad. 10, 248-254.]) without, however, taking into account the pulse smearing effects, caused by the phase Dh (19)[link], which is quadratic in Ω.

The distance RD from a crystal along the wavevector Kh, at which a substantial smearing of the reflected pulse begins, is determined from the condition |Dh| ≃ 1, from which RD ≃ (ΔΩE2|Fhγh|)−1, where in the approximation of the Gaussian forms for R(Ω) and Ain(Ω) the effective spectral width ΔΩE = ΔΩ0ΔΩB/(ΔΩ02 + ΔΩB2)1/2. The expression given above for the distance RD coincides with equation (52)[link] in Appendix C[link]. The effect of smearing and broadening is increased with the reduction of pulse duration, when the incident spectrum width exceeds the spectral width of the Bragg reflection ΔΩ0 [\gg] ΔΩB and therefore ΔΩEΔΩB. If, for example, τ0 = 0.1 fs, then in the Bragg-case Si(220) reflection with b = −2 and wavelength λ = 0.154 nm the distance RD ≃ 64 cm, and in the symmetric Laue case RD ≃ 8 cm (see also Fig. 9).

We shall consider first the reflection of a long incident pulse of duration τ0τB and with small transversal size r0Λ. Fig. 2[link] shows the space distributions of the modulus of amplitudes of the incident pulse |Ain(xs, zs)| with duration τ0 = 10 fs (a) and the reflected pulse |Ah(xp, zp)| in the symmetric Bragg case (b). In this case the longitudinal size of the pulse l0 = τ0c = 3 µm, as well as its transverse size r0 = 10 µm, are comparable with the X-ray extinction length Λ. Hereinafter the system of coordinates (xp, zp) moves together with the reflected pulse [see Fig. 8 and expressions (53)[link], (55)[link]]. After reflection the pulse is strongly stretched in the transverse direction zp and its maximum intensity decreases by more than four times (Fig. 2a[link]). The degree of distortion of the form of the pulse increases with increase in distance R from the crystal to the reflected pulse. Meanwhile the size and duration of the reflected pulse in the longitudinal direction remains almost unchanged. This is explained by that fact that the pulse duration τ0 exceeds the characteristic time τB = 2/ΔΩB ≃ 5.8 fs, where for a thick crystal τB = 2(Λ/c)sin2θB/|γh|. The duration τB is defined by a time delay of the waves reflected from a surface of the crystal and from an effective layer of the crystal of depth zΛ [see also expression (6) of Graeff (2004[Graeff, W. (2004). J. Synchrotron Rad. 11, 261-265.])].

[Figure 2]
Figure 2
Two-dimensional distribution of the amplitude of a Gaussian incident pulse |Ain(xs, zs)| (a) and the reflected pulse |Ah(xp, zp)| (b) in the symmetric Bragg Si(220) reflection (b = −1). Transversal pulse size r0 = 10 µm, pulse duration τ0 = 10 fs (l0 = cτ0 = 3 µm), central wavelength λ = 0.154 nm, thickness of crystal d = 50 µm, phase parameter α0 = 2. Angle between the wavevector Kh and the normal N to the reflected pulse φh = 0. Distance from the crystal to the reflected pulse R = 3 m.

All calculations whose results are shown in Figs. 2–7 are made on the basis of the general formula (9)[link]. For clarity, we shall further consider (see Figs. 3[link]–7) the size of the source rs = 75 µm, the parameter of the square-law phase of radiation of a source αs = 2, and the distance from the source to the crystal zs = 800 m. Then for an incident pulse we find that r0 = 1240 µm, α0 = 36.9 and angular divergence Δθs = 1.46 µrad.

[Figure 3]
Figure 3
Longitudinal sections of the intensity of reflected pulses Ih(zp) along the normal N at xp = 0 in cases of symmetric (a) and asymmetric (b) Bragg reflections. Long (τ01 = 10 fs) and short (τ02 = 1 fs) Gaussian pulses with amplitudes A1 = A2 = 1 [curves 1 and 2 in (a)[link]] are incident on a crystal with time interval Δt12 = 20 fs. Thickness of the crystal d = 5 µm. (a) Dashed curve 3 is the total reflected pulse, angle φh = 0. (b) Distance from the crystal R = 0 (curve 1), R = 2 m (curve 2) and R = 5 m (curve 3). The asymmetry coefficient of the reflection b = −2, critical distance RD = 1.6 m, angle of inclination of the reflected pulse φh = 23.65°.

It is clear that, starting from some duration of the incident pulse τ0 < τB, only a part ΔΩB < ΔΩ0 of the incident frequency spectrum ΔΩ0 will satisfy the diffraction conditions. This results in a sharp reduction of intensity in the case of short incident pulses in comparison with longer pulses.

As an illustration, Fig. 3[link] shows the intensity of reflection Ih(0, zp) of the long and short pulses in the cases of symmetric and asymmetric Bragg reflections. In both cases the intensity of a short pulse after reflection considerably decreases, whereas the long pulse is more weakly deformed. The small peak in the region zp = −c(Δt12 + τB) ≃ −10 µm in Fig. 3(a)[link], where τB = 2(d/c)sin2θB/|γh|, and Δt12 is the time interval between pulses, is connected by reflection of the short pulse from the bottom surface of the crystal (see also Malgrange & Graeff, 2003[Malgrange, C. & Graeff, W. (2003). J. Synchrotron Rad. 10, 248-254.]). In the symmetric Bragg case the intensity of the reflected pulse practically does not change for an increase in the distance R from the crystal to the pulse. At the same time, in the asymmetric Bragg case the maximal intensity of the reflected pulse decreases with an increase in the distance R: its width increases and the contribution of the short pulse to the total intensity becomes extremely small at R ≃ 0.5 m (see Fig. 3[link]b).

Moreover, unfortunately for practical applications, for very short femtosecond pulses the duration and shape of a reflected pulse become almost independent of the incident pulse characteristics (Figs. 4[link][link]–7) (see also Graeff, 2002[Graeff, W. (2002). J. Synchrotron Rad. 9, 82-85.], 2004[Graeff, W. (2004). J. Synchrotron Rad. 11, 261-265.]; Malgrange & Graeff, 2003[Malgrange, C. & Graeff, W. (2003). J. Synchrotron Rad. 10, 248-254.]). Formally this can be seen from equation (17)[link], as the smooth function Ain(Ω) can be taken outside of the integration. Therefore the form of a pulse on the crystal surface is determined by inverse Fourier transformation of the reflection coefficient R(Ω) [as can also be seen in the Green function in Chukhovskii & Förster (1995[Chukhovskii, F. N. & Förster, E. (1995). Acta Cryst. A51, 668-672.])].

[Figure 4]
Figure 4
Two-dimensional distribution of amplitude of the reflected pulse |Ah(xp, zp)| in the symmetric Laue case (b = 1). Distance from the crystal to the pulse R = 2 cm (a), and R = 10 cm (b). Duration of incident pulse τ0 = 1 fs, crystal thickness d = 23.19 µm, critical distance RD = 8.2 cm, angle of pulse inclination φh = −41.2°. For the incident pulse of longitudinal size l0 = 0.3 µm there appears to be a δ-function on the background of the spatial distribution of the reflected pulse and consequently it is not shown in these figures. Other parameters: size of source rs = 75 µm, phase parameter αs = 2, distance from source to crystal zs = 800 m, transversal size of incident pulse r0 = 1240 µm, phase parameter α0 = 36.9.
[Figure 5]
Figure 5
Longitudinal section of intensity of the reflected pulse Ih(zp) at xp = 0 in the symmetric Laue case (b = 1). Distance from crystal to pulse R = 0 (curve 1), R = 5 cm (curve 2) and R = 20 cm (curve 3). Duration of incident pulse τ0 = 0.1 fs (l0 = 0.03 µm), crystal thickness d = 23.19 µm, distance RD = 7.9 cm, angle φh = −41.2°. Other parameters are the same as in Fig. 4[link].

The second peak in the Bragg case at zp ≃ −VhτB, where speed Vh is defined in expression (22)[link] and τB = 2(d/c)sin2θB/|γh|, arises owing to reflection from the rear surface of the crystal [see Fig. 6(a)[link] at zp ≃ −5.6 µm, and Fig. 7[link] (curve 1) at zp ≃ −3 µm]. It is easy to see that any asymmetric Bragg case reflection and any Laue case reflection are not quite acceptable for diffraction tailoring of pulses, because already at distances R as short as 10–30 cm from the crystal the pulses become considerably diffused (Figs. 4–7). The transmitted pulse I0(x, z, t) meanwhile practically coincides in form and intensity with the incident pulse, as the transmission coefficient T(Ω) = exp(ik0χ0d/2γ0) stays constant everywhere except the very narrow spectral slot |Ω| ≤ ΔΩB [see equations (12)[link] and (14)[link]]. It is obvious that the group of super-short statistically unconnected pulses with total duration τp < τB become merged into one wide asymmetric pulse of duration of the order of τB after a reflection (see also Shastri et al., 2001a[Shastri, S. D., Zambianchi, P. & Mills, D. M. (2001a). J. Synchrotron Rad. 8, 1131-1135.]).

[Figure 6]
Figure 6
Two-dimensional distribution of amplitude of the reflected pulse |Ah(xp, zp)| in the asymmetric Bragg case (b = −2). Distance from crystal to pulse R = 5 cm (a), and R = 100 cm (b). Duration of incident pulse τ0 = 1 fs, crystal thickness d = 5 µm, critical distance RD = 64.5 cm, angle of pulse inclination φh = 23.65°. Other parameters are the same as in Fig. 4[link].
[Figure 7]
Figure 7
Longitudinal section of intensity of the reflected pulse Ih(zp) at xp = 0 in the asymmetric Bragg case (b = −0.5). Distance from crystal to pulse R = 0 (curve 1), R = 50 cm (curve 2) and R = 100 cm (curve 3). Duration of incident pulse τ0 = 0.1 fs, crystal thickness d = 5 µm, distance RD = 63.6 cm, angle φh = −12.35°. Other parameters are the same as in Fig. 4[link].

It is of interest to consider the space and time coherency of XFEL radiation, and the radiation of reflected pulses. The coherence function of an incident pulse is given by

[\Gamma_{\rm{in}}(\rho,\tau)=P^{-1}\left|\left\langle{A_{\rm{in}}(x,t)}A_{\rm{in}}^*(x+\rho,t+\tau)\right\rangle\right|,\eqno(23)]

where P = [Iin(x, t)Iin(x + ρ, t + τ)]1/2, Iin(x, t) = 〈|Ain(x, t)|2〉 (angular brackets mean the average over a sufficiently large time interval); Γin(0, τ) and Γin(ρ, 0) are the functions of time and space coherence, respectively, with Γin(0, 0) = 1. According to the calculations of Saldin et al. (2004[Saldin, E. L., Schneidmiller, E. A. & Yurkov, M. V. (2004). Report TESLA-FEL 2004-02. DESY, Hamburg, Germany.]), XFEL pulses are completely coherent over the whole cross section, and the coherence time, which is obtained after substitution of calculated pulses with amplitude and phase modulation in (23)[link], has the value τc = 0.14 fs. More convenient for the analysis is the spectral representation,

[\Gamma_{\rm{in}}(\rho,\tau)=I_{\rm{in}}^{-1}\big|\textstyle\int\!\!\int|A_{\rm{in}}(q,\Omega)|^2\exp[i(q\rho-\Omega\tau)]\,{\rm{d}}q\,{\rm{d}}\Omega\big|,\eqno(24)]

where

[I_{\rm{in}}=\textstyle\int\!\!\int|A_{\rm{in}}(q,\Omega)|^2\,{\rm{d}}q\,{\rm{d}}\Omega.]

It can be shown that the coherence functions of reflected and transmitted pulses are determined by the expression

[\eqalignno{\Gamma_g(\rho,\tau)={}&I_g^{-1}\big|\textstyle\int\!\!\int|B_g(q,\Omega)A_{\rm{in}}(q,\Omega)|^2\cr&\times\exp[i(q\rho-\Omega\tau)]\,{\rm{d}}q\,{\rm{d}}\Omega\big|.&(25)}]

If in the region of significant variation of the spectrum Ain(q, Ω) the coefficients Bg ≃ a constant, then the degree of coherence of the reflected pulse ΓgΓin, i.e. the coherence remains preserved. For a short pulse, for which the spectral width ΔΩ0 [\gg] ΔΩB, the time coherence, as follows from (25)[link], is increased, i.e. a partial monochromatization takes place; however, at the same time the pulse intensity decreases.

One of the most serious problems in diffraction of the powerful XFEL pulses will be the very high thermal load on diffracting crystals. So far there is no exact solution for X-ray diffraction taking into account thermal heating, but it is possible to make some estimations. From analysis of the Green function of the thermal conductivity equation with distributed thermal sources in a crystal subsurface layer, it follows that the time of temperature propagation over a distance Δx is Δt ≃ (Δx)2/4a2, where a2 = λT/cTρ, λT and cT are coefficients of thermal conductivity and thermal capacity, respectively, and ρ is the crystal density. For silicon at temperature T = 300 K, λT ≃ 150 W m−1 K−1, cT ≃ 700 J kg−1 K−1, ρ = 2.3 g cm−3 (Grigor'ev & Meilikhova, 1991[Grigor'ev, I. S. & Meilikhova, E. Z. (1991). Editors. Physical Values. Handbook. Moscow: Energoatomizdat. (In Russian.)]). For ΔxΛ, one obtains Δt ≃ 13 ns, and this is much longer than the duration of an X-ray pulse τ0 = 0.1–200 fs (for diamond Δt ≃ 20 ns). Thus it is quite possible that the X-ray laser pulses are simply too short to influence their own diffraction scattering through heating. An increase in temperature of a crystal by ΔT = ΔθBcotanθB/αT ≃ 10 K, where αT is the coefficient of linear expansion (for silicon αT = 2.54 × 10−6 K−1), results in displacement of the Bragg peak by an angle ΔθB, which is not essential for short pulses with ΔΩ0 [\gg] ΔΩB.

4. Conclusions

In conclusion, this paper presents a most general approach to the consideration of the diffraction of arbitrary X-ray pulses in crystals and their subsequent dispersion in space. The main attention is devoted to analysis of how space and time change the form and duration of short pulses depending on the distance from the crystal. It is shown that the unique opportunity to avoid distortion of the form and duration of a reflected femtosecond pulse is achieved by use of symmetric reflections in the Bragg diffraction geometry.

APPENDIX A

Calculation of projection kgz

Let us present the values of the wavevectors k0 and kgx in the expression for kgz in (7)[link] in the following form: k0 = K0(1 + ξ1), where ξ1 = Ω/ω0, and kgx = Kgx(1 + ξ2), where ξ2 = q/Kgx. As a result for the square-root expression in (7)[link] we find that

[k_0^2-k_{gx}^2=K_{gz}^2\left[1+2(a\xi_1-b\xi_2)+(a\xi_1^2-b\xi_2^2)\right],\eqno(26)]

where Kgz 2 = K0 2Kgx 2,

[a=(K_0/K_{gz})^2=1/\gamma_g^2,\quad\quad b=(K_{gx}/K_{gz})^2=\tan^2\theta_g.\eqno(27)]

Here γg = cos(Kg·n) = cosθg, Kh = (Khx, σh|Khz|). In the Laue case, γh > 0, and in the Bragg case, γh < 0.

We shall consider now that at ξ [\ll] 1 the following expansion takes place,

[(1+\xi)^{1/2}\simeq1+(1/2)\xi-(1/8)\xi^2.\eqno(28)]

Then with use of expression (28)[link], from equation (26)[link] we find that

[\eqalignno{\sigma_g(k_0^2-k_{gx}^2)^{1/2}={}&K_{gz}(1+a\xi_1-b\xi_2) -(1/2)K_{gz}\left[a(a-1)\xi_1^2\right.\cr&\left.-\,2ab\xi_1\xi_2+b(b+1)\xi_2^2\right],&(29)}]

where Kgz = K0γg. In view of an obvious form of a and b (27)[link] it is easy to see that in (29)[link] factors a(a − 1) = ab = b(b + 1) = tan2θg/γg2. As a result, from (29)[link] we find finally that

[\eqalignno{k_{gz}={}&K_{gz}-q\tan\theta_g+(\Omega/c\gamma_g)\cr&-[q-(\Omega/c)\sin\theta_g]^2/(2K_0\gamma_g^3).&(30)}]

If now in addition to kgz (30)[link] we consider the x-components of wavevectors kg and the item ωt, the phase in the exponential in the integral (6)[link] will have the following form,

[{\bf{k}}_g{\bf{r}}-\omega{t}=({\bf{K}}_g{\bf{r}}-\omega_0t)+S_g(q,\Omega)+D_g(q,\Omega),\eqno(31)]

where the linear (Sg) and square-law (Dg) phases are set by expressions (10)[link] and (11)[link], respectively.

APPENDIX B

X-ray pulse propagation in free space

The aim of Appendix B is to prove the validity of the form of the Gaussian incident pulse given in (15)[link]. In the plane of a source zs = 0 (the exit window of the free-electron laser) is set a field

[E_{\rm{s}}(x_{\rm{s}},t)=A_{\rm{s}}(x_{\rm{s}},t)\exp(-i\omega_0t),\eqno(32)]

where As(xs, t) is the complex slowly varying amplitude of the field, xs is the transversal coordinate in the plane of the source and ω0 is the average frequency of radiation. It is required to find the field E(xs, zs, t) in any point in space (xs, zs) at the moment in time t.

We shall present the field Es(xs, t) (32)[link] in the form of an expansion of the Fourier integral over the plane waves,

[E_{\rm{s}}(x_{\rm{s}},t)=\textstyle\int\!\!\int{E_{\rm{s}}}(q,\omega)\exp(iqx_{\rm{s}}-i\omega{t})\,{\rm{d}}q\,{\rm{d}}\omega,\eqno(33)]

where

[E_{\rm{s}}(q,\omega)=(2\pi)^{-2}\textstyle\int\!\!\int{E_{\rm{s}}}(x_{\rm{s}},t)\exp(-iqx_{\rm{s}}+i\omega{t})\,{\rm{d}}x_{\rm{s}}{\rm{d}}t.\eqno(34)]

Propagation of the pulse field E(xs, zs, t) in the region zs ≥ 0 is described by the wave equation

[\Delta{E}-(1/c^2)\,\partial^2E/\partial{t^2}=0,\eqno(35)]

where Δ = ∂2/∂xs2 + ∂2/∂zs2 is the Laplace operator. In view of the boundary condition for the field E(xs, 0, t) = Es(xs, t) in the plane zs = 0 from (33)[link] and (35)[link] it is easy to see that

[E(x_{\rm{s}},z_{\rm{s}},t)=\textstyle\int\!\!\int{E_{\rm{s}}}(q,\omega)\exp(iqx_{\rm{s}}+ik_zz_{\rm{s}}-i\omega{t})\,{\rm{d}}q\,{\rm{d}}\omega,\eqno(36)]

where kz = (k2q2)1/2, k = ω/c.

From (32)[link] and (34)[link] it follows that Es(q, ω) = As(q, Ω), where Ω = ωω0. If the characteristic size of the source rs [\gg] λ, and the pulse duration τ0 [\gg] T, where λ is the wavelength of radiation and T is the period, then q [\ll] k and Ω [\ll] ω0. In this case, with use of the expansion (28)[link] for values of kz in (36)[link], we have

[k_z\simeq K_0+\Omega/c-q^2/2K_0,\eqno(37)]

where K0 = ω0/c = 2π/λ is an average wavevector of the pulse. Substituting (37)[link] into (36)[link] we find the following expression for the pulse field in a plane zs,

[E(x_{\rm{s}},z_{\rm{s}},t)=A(x_{\rm{s}},z_{\rm{s}},t)\exp(iK_0z_{\rm{s}}-i\omega_0t),\eqno(38)]

where A(xs, zs, t) is the slowly varying amplitude of the pulse,

[\eqalignno{A(x_{\rm{s}},z_{\rm{s}},t)={}&\textstyle\int\!\!\int{A_{\rm{s}}}(q,\Omega)\exp\left[iqx_{\rm{s}}-iq^2z_{\rm{s}}/2K_0\right.\cr&\left.-\,i\Omega(t-z_{\rm{s}}/c)\right]\,{\rm{d}}q\,{\rm{d}}\Omega.&(39)}]

We shall consider now propagation in space of a Gaussian pulse for which it is possible to find simple analytical expressions. We shall present the amplitude (32)[link] of the field on the source surface zs = 0 in the following form,

[A_{\rm{s}}(x_{\rm{s}},t)=\exp\left[-(x_{\rm{s}}/r_{\rm{s}})^2+i\varphi_{\rm{s}}(x_{\rm{s}})-(t/\tau_0)^2\right],\eqno(40)]

where rs is the size of the source in the plane zs = 0, τ0 is the pulse duration and φs(xs) is the phase of the complex amplitude (40)[link]. Furthermore we shall consider that this phase is a square-law function of the coordinate xs, i.e. φs(xs) = αs(xs/rs)2, where parameter αs is equal to the phase at |xs| = rs.

For calculation of the Fourier amplitudes As(q, Ω) in (39)[link] and for calculations of other integrals the known so-called main optical integral (Gradshteyn & Ryzhik, 1980[Gradshteyn, I. S. & Ryzhik, I. M. (1980). Table of Integrals, Series and Products. New York: Academic Press.]) is used,

[\textstyle\int\exp(-i\beta{x}+i\gamma{x^2})\,{\rm{d}}x=(i\pi/\gamma)^{1/2}\exp(-i\beta^2/4\gamma),\eqno(41)]

where β and γ are arbitrary complex values.

Substituting (40)[link] into (34)[link] and (39)[link] leads to the following expression for the pulse amplitude at any plane zs,

[\eqalignno{A(x_{\rm{s}},z_{\rm{s}},t)={}&A_{\rm{s}}\exp[-(x_{\rm{s}}/r_0)^2+i\varphi_0(x_{\rm{s}})\cr&-(t-z_{\rm{s}}/c)^2/\tau_0^2+i\Phi_0],&(42)}]

where As = 1/M1/2, M = [(1 + αsW)2 + W2]1/2, W = λzs/πrs2 is the wave parameter,

[\eqalign{&r_0=r_{\rm{s}}M,\quad\quad\varphi_0(x_{\rm{s}})=\alpha_0(x_{\rm{s}}/r_0)^2,\cr&\alpha_0=\alpha_{\rm{s}}+(1+\alpha_{\rm{s}}^2)W,\cr&\Phi_0=-(1/2)\arctan[W/(1+\alpha_{\rm{s}}W)].}\eqno(43)]

From expression (42)[link] we can see that the initial Gaussian pulse keeps its form and duration in the process of propagation in space; however, the transversal size of the pulse r0(zs) increases M times in comparison with rs with increase in distance zs and with increase in phase parameter αs. This phase parameter describes an initial curvature of the wavefront. The phase of the pulse φ0(xs), also a square-law function, depends on the transversal coordinate. The parameter of this phase α0(zs) increases with increase in distance zs and with increase in parameter αs, and also increases with reduction of the source size rs. The phase parameter α0 ≠ 0, even at the initial plane wavefront, i.e. at αs = 0. The phase Φ0(zs) does not depend on the transversal coordinate xs and does not play an essential role during propagation and diffraction of the pulse. Later we shall consider for simplicity that in (42)[link] As = 1, Φ0 = 0.

The width of the angular spectrum of a pulse (42)[link] Δθs = Δqs/K0, i.e. the width of the function |A(q, Ω, zs)| ≃ exp[−(q/Δqs)2], where Δqs = 2(1 + α02)1/2/r0 = 2(1 + αs2)1/2/rs, does not depend on the distance zs and is determined by the expression

[\Delta\theta_{\rm{s}}=(\lambda/\pi{r_{\rm{s}}})(1+\alpha_{\rm{s}}^2)^{1/2}.\eqno(44)]

The angular width Δθs (44)[link] in the general case exceeds the diffraction divergence Δθd = (λ/πrs), related only to the size of the source rs.

Theoretical calculations show (Saldin et al., 2004[Saldin, E. L., Schneidmiller, E. A. & Yurkov, M. V. (2004). Report TESLA-FEL 2004-02. DESY, Hamburg, Germany.]) that on exit from undulator SASE1 (λ ≃ 0.1 nm) the full width of the pulse at half-height is equal to 90 µm, and the angular divergence of the beam is equal to 1.1 µrad. In our notation this means that rs ≃ 76.4 µm and Δθs ≃ 0.93 µrad. From here and (44)[link] it follows that the phase parameter αs ≃ 2. Then with use of formulae (43)[link] it is easy to show that at the distance zs = 800 m the transversal pulse size r0 ≃ 820 µm, and the phase parameter α0 ≃ 24.

The relation between coordinates (xs, zs) on the source of an X-ray pulse and coordinates (x, z) on the crystal surface is determined by means of the following expressions: xs = xcosθ0zsinθ0, zs = xsinθ0 + zcosθ0 + z1, where θ0 is the incident angle of the pulse on the crystal with respect to the normal n to the crystal surface, and z1 is the distance from the source to the crystal (see Fig. 8[link]). Then the amplitude of the field (42)[link] on the crystal surface z = 0 will be

[\eqalignno{A_{\rm{in}}(x,t)={}&\exp\left[-(x\gamma_0/r_0)^2(1-i\alpha_0)\right.\cr&\left.-\,(t-x\sin\theta_0/c)^2/\tau_0^2\right],&(45)}]

where γ0 = cosθ0, and the time t is counted from the moment t1 = z1/c of incidence of the pulse maximum at x = 0, z = 0 on the crystal.

[Figure 8]
Figure 8
Geometry in real space of an asymmetric Bragg case (a) and Laue case (b) diffraction. Here K0 and Kh are the average wavevectors of the incident and reflected pulses, respectively; N is the normal to the long axis of the reflected pulse; (xs, zs) is a Cartesian system of coordinates on the source surface, (x, z) is a system of coordinates on the crystal surface, the system of coordinates (xh, zh) moves together with the reflected pulse, axis zh is directed along the wavevector Kh; axes of the moving system of coordinates (xp, zp) are directed along the main axes of the pulse, φh is an angle between the normal N to the reflected pulse and the direction of pulse propagation Kh.

APPENDIX C

Reflection of a Gaussian pulse

We shall consider diffraction reflection of the incident pulse Ain(x, t) (45)[link] from the crystal. From the expression (45)[link] in view of (41)[link] we find the following expression for Fourier amplitudes of the incident pulse in (9)[link],

[\eqalignno{A_{\rm{in}}(q,\Omega)={}&A_0\exp\left[-(q-\Omega\sin\theta_0/c)^2(1+i\alpha_0)/{\Delta}q_0^2\right.\cr&\left.-\,(\Omega/\Delta\Omega_0)^2\right],&(46)}]

where ΔΩ0 = 2/τ0 is the spectral width of the incident pulse, Δq0 = 2γ0(1 + α02)1/2/r0 is the width of the angular spectrum in q-space, and amplitude A0 = (1 + iα0)1/2/(πΔq0ΔΩ0). It is easy to show that Δq0 = γ0Δqs, where Δqs = 2(1 + αs2)1/2/rs is the width of the angular spectrum of the pulse in the plane zs = 0 of the source.

For simplicity of the analysis of the form and the orientation of the reflected pulse we shall present the amplitude reflection coefficient R(q, Ω) in the integral (9)[link] in the form of a Gaussian function [see argument α(q, Ω) (13)[link] in (12)[link] and (14)[link]],

[R(q,\Omega)=\exp\left\{-\left[q-(\Omega/c)\sin\psi/\cos\theta_{\rm{B}}\right]^2/\Delta{q_{\rm{B}}^2}\right\},\eqno(47)]

where ΔqB = K0γ0ΔθB is the width of the diffraction reflection curve.

We shall now substitute Ain(q, Ω) (46)[link] and R(q, Ω) (47)[link] into the general integral equation (9)[link] for the amplitude of the reflected pulse, where g = h. As a result, using (41)[link] we find that

[A_h(x,z,t)=A_R\exp\left[-\Phi_1^2(1-i\alpha_0)-\Phi_2^2(1-i\beta_0)\right],\eqno(48)]

where

[\Phi_1=(x-z\tan\theta_h)\gamma_0/r_0,\eqno(49)]

[\Phi_2=\left[t-z/c\gamma_h-(x-z\tan\theta_h)\sin\theta_0/c\right]/\tau_R.\eqno(50)]

Here, τR is the duration of the reflected pulse, which is determined by the following formula,

[\tau_R=\tau_E(1+\beta_0^2)^{1/2},\eqno(51)]

where

[\eqalign{&\tau_E=(\tau_0^2+\tau_{\rm{B}}^2)^{1/2},\quad\quad\tau_{\rm{B}}=2/\Delta\Omega_{\rm{B}},\cr&\Delta\Omega_{\rm{B}}=\omega_0\Delta\theta_{\rm{B}}\,{\rm{cotan}}\,\theta_{\rm{B}},}\eqno(51a)]

[\eqalign{&\beta_0=(\tau_Z/\tau_E)^2,\quad\quad\tau_Z=2(F_hz)^{1/2},\cr&F_h=(\sin\theta_0-\sin\theta_h)^2/(2K_0c^2\gamma_h^3),}\eqno(51b)]

[A_R=(\tau_0/\tau_R)(1-i\beta_0)^{1/2}.\eqno(51c)]

In order to obtain expression (48)[link] it is considered that the ratio Δq0/ΔqB = Δθs/ΔθB [\ll] 1, where Δθs = (λ/πrs)(1 + αs2)1/2 is the width of the angular spectrum of the source radiation, and distance |z| [\ll] zF, where zF = πr02/[λb2(1 + α02)].

From equations (51)[link] it follows that the pulse duration τR after reflection from the crystal increases in comparison with τ0 for two reasons. The first reason is related to the finite quantity of the spectral width ΔΩB of the Bragg reflections. For sufficiently short pulses the spectral width ΔΩ0 > ΔΩB, and therefore τB > τ0. The second reason is connected to diffusion broadening of the part of the pulse duration τZ on increasing the distance from the crystal to the reflected pulse along the wavevector Kh. From the characteristic distance RD, at which the intensity of the pulse |AR|2 will decrease twice, it is possible to estimate from the equation β0 = 1,

[R_D=(\tau_0^2+\tau_{\rm{B}}^2)/4|F_h\gamma_h|.\eqno(52)]

From equation (52)[link] it follows that diffusion broadening of the pulse is absent only in the case of symmetric Bragg reflection (b = −1), at which sinθ0 = sinθh, Fh = 0 and RD → ∞. The critical distance RD (52)[link] increases with increase in τ0, if τ0 > τB (see Fig. 9[link]). The dependence of the value τB from the asymmetry coefficient of reflection b is shown in the insert of Fig. 9[link]. For short pulses with τ0 [\ll] τB the distance RD does not depend on the duration of the incident pulse τ0. Unfortunately in practice the distance RD does not exceed several tens of centimetres in the Laue case at τ0 ≤ 1–10 fs (see Fig. 9[link]).

[Figure 9]
Figure 9
Dependence of the critical distance RD on the asymmetry coefficient of reflection b at various durations of the incident pulse τ0: curve 1, 1 fs; curve 2, 10 fs; curve 3, 20 fs; σ-polarization.

From the form of arguments Φ1 and Φ2 in (48)[link] we can see that the reflected pulse Ah(x, z, t) propagates with the speed of light c in a vacuum along the direction of the wavevector Kh. The orientation of the pulse in space is determined mainly by the angles θ0 and θh, i.e. by the asymmetry coefficient of the reflection b.

For analysis of this, we shall pass to the system of coordinates (xh, zh), which moves together with the pulse, and the axis zh is directed along the wavevector Kh, i.e. this coordinate system is turned in relation to the laboratory system of coordinates (x, z) by angle θh (see Fig. 8[link]),

[\eqalign{&x=ct\sin\theta_h+x_h\cos\theta_h+z_h\sin\theta_h,\cr&z=ct\cos\theta_h+z_h\cos\theta_h-x_h\sin\theta_h.}\eqno(53)]

Then the functions Φ1,2 in (48)[link] will have the following form,

[\Phi_1=bx_h/r_0,\quad\quad\Phi_2=(a_{0h}x_h+z_h)/(c\tau_R),\eqno(54)]

where a0h = (sinθ0 − sinθh)/γh.

From equations (54)[link] it can be seen that only in the symmetric Bragg geometry when sinθ0 = sinθh and a0h = 0 do the axes of the reflected pulse coincide with axes xh and zh. In all other cases the pulse propagates so that its axes are inclined by some angle φh relative to the direction of propagation Kh (see Fig. 8[link]).

We shall consider the most interesting case of a wide and short pulse, whose transversal size r0 is much larger than its longitudinal size l0 = cτ0, incident on a crystal. In order to find the angle φh between the normal N to the long axis of the reflected pulse and the direction of distribution Kh we shall pass to a new system of coordinates (xp, zp) by means of equations

[\eqalign{&x_h=x_p\cos\varphi_h+z_p\sin\varphi_h,\cr&z_h=z_p\cos\varphi_h-x_p\sin\varphi_h.}\eqno(55)]

The angle φh is obtained from the condition that the coefficient of product xpzp in the expression Φ12 + Φ22 in (48)[link] equals zero. As a result we find that

[\tan\varphi_h=a_{0h}(1+\delta),\eqno(56)]

where δ = (bcτR/r0)2/(1 + a0h2) [\ll] 1. The dependence of the angle of inclination φh of the reflected pulse from the asymmetry coefficient of the reflection b is shown in Fig. 10[link].

[Figure 10]
Figure 10
Dependence of the inclination angle of the reflected pulse φh (1), of the incident angle θ0 (2), and of the reflected angle θh (3) on the asymmetry coefficient of the reflection b.

So, the modulus of the amplitude of the reflected pulse in a moving system of coordinates (xp, zp) is

[|A_h(x_p,z_p)|=(\tau_0/\tau_R)\exp\left[-(x_p/r_{\rm{T}})^2-(z_p/r_{\rm{L}})^2\right],\eqno(57)]

where rT = r0/(|b|cosφh) is the transversal size of the pulse and rL = VLτR is the longitudinal size of the pulse. Here VL = ccosφh is a projection of the pulse speed c on the axis zp. Expression (57)[link] represents, figuratively speaking, an instant photo-picture of the reflected pulse in the moment of time t = z/cγh. From equation (56)[link] it is easy to see that

[\cos\varphi_h\simeq|\gamma_h|/(1-2\sin\theta_0\sin\theta_h+\sin^2\theta_0)^{1/2}.\eqno(58)]

This expression can be found also from the condition of equality of the optical paths LABC = LDEF (see Fig. 8[link]).

Earlier (Malgrange & Graeff, 2003[Malgrange, C. & Graeff, W. (2003). J. Synchrotron Rad. 10, 248-254.]) the speed VL was not quite correctly referred to as the group velocity. Such a discrepancy has arisen because in this work the incident and reflected pulses were considered infinite in the transversal direction along the axis xp (i.e. r0 → ∞). For this reason propagation of the reflected pulse with projection speed VT = csinφh along this transversal direction could not be considered in principle.

The most important, and rather unfavourable for practical applications, feature of short X-ray pulse diffraction is the inclination of the reflected pulse relative to the propagation direction (see Figs. 8[link] and 10[link]). The real pulse duration τtot, i.e. the time of its passage through a plane, perpendicular to wavevector Kh, will be determined both by angle of inclination φh and by projection r0/|b| onto this plane of the transversal size of the pulse: τtot = (r0/c)|sinφh/b|. If, for example, r0 = 800 µm, b = 1, then the angle of inclination φh = −41.2° and the total duration of the reflected pulse τtot = 2 × 103 fs, which exceeds the duration of the incident femtosecond pulse by some orders.

Acknowledgements

The author is very grateful to D. Novikov for helpful discussions. The work was supported by the Russian Foundation Base Research (Grants No. 06-02-17249, No. 07-02-00324) and the International Science and Technology Center (Project No. 3124).

References

First citationAfanas'ev, A. M. & Kohn, V. G. (1971). Acta Cryst. A27, 421–430.  CrossRef CAS IUCr Journals Web of Science
First citationAltarelli, M. et al. (2006). Editors. Report DESY 2006–097. DESY, Hamburg, Germany. (http://xfel.desy.de/tdr/index_eng.html .)
First citationArthur, J. et al. (2002). LCLS Conceptual Design Report. LCLS, USA. (http://www-ssrl.slac.stanford.edu/lcls/cdr/ .)
First citationBauspiess, W., Bonse, U. & Graeff, W. (1976). J. Appl. Cryst. 9, 68–80.  CrossRef IUCr Journals Web of Science
First citationChukhovskii, F. N. & Förster, E. (1995). Acta Cryst. A51, 668–672.  CrossRef CAS Web of Science IUCr Journals
First citationGradshteyn, I. S. & Ryzhik, I. M. (1980). Table of Integrals, Series and Products. New York: Academic Press.
First citationGraeff, W. (2002). J. Synchrotron Rad. 9, 82–85.  Web of Science CrossRef CAS IUCr Journals
First citationGraeff, W. (2004). J. Synchrotron Rad. 11, 261–265.  Web of Science CrossRef CAS IUCr Journals
First citationGrigor'ev, I. S. & Meilikhova, E. Z. (1991). Editors. Physical Values. Handbook. Moscow: Energoatomizdat. (In Russian.)
First citationMalgrange, C. & Graeff, W. (2003). J. Synchrotron Rad. 10, 248–254.  Web of Science CrossRef CAS IUCr Journals
First citationSaldin, E. L., Schneidmiller, E. A. & Yurkov, M. V. (2004). Report TESLA-FEL 2004–02. DESY, Hamburg, Germany.
First citationShastri, S. D., Zambianchi, P. & Mills, D. M. (2001a). J. Synchrotron Rad. 8, 1131–1135.  Web of Science CrossRef CAS IUCr Journals
First citationShastri, S. D., Zambianchi, P. & Mills, D. M. (2001b). Proc. SPIE, 4143, 69–77.  CrossRef CAS
First citationSlobodetzkii, I. Sh. & Chukhovskii, F. N. (1970). Kristallografiya, 15, 1101–1107. (In Russian.)
First citationTanaka, T. & Shintake, T. (2005). SCSS X-FEL Conceptual Design Report, edited by Takashi Tanaka and Tsumoru Shintake. SCSS XFEL, R&D Group, RIKEN Harima Institute/SPring-8, Japan. (http://www-xfel.spring8.or.jp/SCSSCDR.pdf .)
First citationWark, J. S. & He, H. (1994). Laser Part. Beams, 12, 507–513.  CrossRef CAS
First citationWark, J. S. & Lee, R. W. (1999). J. Appl. Cryst. 32, 692–703.  Web of Science CrossRef CAS IUCr Journals

© International Union of Crystallography. Prior permission is not required to reproduce short quotations, tables and figures from this article, provided the original authors and source are cited. For more information, click here.

Journal logoJOURNAL OF
SYNCHROTRON
RADIATION
ISSN: 1600-5775
Follow J. Synchrotron Rad.
Sign up for e-alerts
Follow J. Synchrotron Rad. on Twitter
Follow us on facebook
Sign up for RSS feeds