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

Journal logoJOURNAL OF
SYNCHROTRON
RADIATION
ISSN: 1600-5775

Analytic expressions for the angular and the spectral fluxes at Compton X-ray sources

crossmark logo

aLaboratoire de l'Accélérateur Linéaire, Orsay, France
*Correspondence e-mail: mjacquet@lal.in2p3.fr

Edited by M. Yamamoto, RIKEN SPring-8 Center, Japan (Received 21 April 2016; accepted 26 October 2016)

The goal of this paper is to express simply the number of photons impinging on a target in the framework of accelerator-based Compton X-ray sources. From the basic kinematics of Compton sources, analytic formulas for the angular and the spectral fluxes are established as functions of the energy spread or/and the angular divergence of the electron and the laser beams. Their detailed predictions are compared with Monte Carlo simulations. These analytic expressions allow one to compute in a simple and precise way the X-ray flux in a given angular acceptance and a given energy bandwidth, knowing the characteristics of the incoming beams.

1. Introduction

Today, Compton X-ray sources are in full development thanks to the exceptional improvement of high-power lasers over the last 15 years. The principle is based on the production of X-ray pulses of a few tens of keV in energy by Compton backscattering of intense laser light of micrometric wavelength against an electron bunch of tens of MeV in energy (i.e. about 100 times less energetic than electrons used to produce synchrotron radiation in the X-ray range). Such Compton sources are compact installations (with areas of ∼100 m2) which provide high-intensity, high-quality X-ray beams with a tunable energy. The present most ambitious projects aim at producing a total flux of 1012 1014 photons s−1 (Jacquet, 2014[Jacquet, M. (2014). Nucl. Instrum. Methods Phys. Res. B, 331, 1-5.]) that gives access to experimental methods currently used at synchrotron beamlines. For various applications, the development of these sources will allow the use of powerful analysis techniques in such environments as hospitals, laboratories or museums. The diversity of the possible applications of these sources increases the demand for a simple formulation of the X-ray flux available at a target sample with an intuitive understanding of its spectral and spatial properties. Our motivation is to guide the users when evaluating the performances of such sources in their domain of competence (medical, materials, art history, etc.). Indeed, to our knowledge, there is no analytic formulas in the literature describing angular and spectral X-ray fluxes expected from a given Compton source. There are some analogies between synchrotron radiation sources and Compton sources, but the spectral and the spatial properties are different. Thus, Compton backscattering sources need a specific description. Our analytic formulas provide the flux within a finite polar angle around the electron beam direction and within a finite bandwidth around the Compton edge. The energy spread and the angular divergence of the electron beam and of the laser affect the X-ray flux in the chosen kinematic region. The estimation of such effects is the purpose of this article. The analytic expressions are obtained by convolving purely kinematic effects with expressions that characterize the laser and the electron beam properties.

The basic principles of a Compton source, the kinematics and our calculation framework are described in §2[link]. The performances of current laser systems are such that the main impact on the Compton source characteristics comes from the angular divergence [\sigma_{\rm{e}}^{\,\prime}] and the energy spread [\sigma_{\rm{e}}] of the incident electron beam. The latter may vary widely depending on the chosen accelerator complex (see §2.2[link]). Analytic expressions for the total X-ray flux and for its angular and spectral dependences are established in §3[link] and §4[link] as functions of these two electron beam parameters and compared with Monte Carlo simulations.

2. Key parameters for the X-ray flux

This section introduces the key parameters needed to understand the analytic developments that follow.

2.1. The Compton process

Fig. 1[link] describes the Compton scattering between a free relativistic electron of Lorentz factor γ and a photon of energy [E_{\rm{L}}] which is the basic process exploited in an accelerator-based X-ray source.

[Figure 1]
Figure 1
Schematic drawing of the scattering process between an electron and a laser photon. xyz is the laboratory frame.

Electrons of a few tens of MeV are used in X-ray sources whereas the laser is usually in the infrared domain. In the laboratory frame, the scattered photon energy [E_{\rm{X}}] varies quadratically with the electron energy and linearly with the laser photon energy [E_{\rm{L}}]. Assuming [\gamma\gg1] and the laser photon energy is small compared with the electron rest mass energy, we can derive from the kinematics of the process

[\eqalignno{ E_{\rm{X}} & = {{E_{\rm{m}}}\over{1+\gamma^{\,2}\theta_{\rm{X}}^{\,2}}}_{\vphantom{\big|}}, & (1a) \cr E_{\rm{m}} & = 2\gamma^{\,2} E_{\rm{L}} \left(1 + \cos\theta_{\rm{c}}\right), & (1b)}]

where [\theta_{\rm{c}}] is the collision angle and [\theta_{\rm{X}}] the scattering polar angle of the Compton photon with respect to the incoming electron momentum. Equation (1a)[link] implies an univocal dependence between the energy of the backscattered photon and its emission angle [\theta_{\rm{X}}]. Photons of maximum energy [E_{\rm{m}}] are those emitted on-axis ([\theta_{\rm{X}}] = 0°). For an X-ray which is produced by an electron of Lorentz factor γ, let ɛ be the ratio of [E_{\rm{X}}] and [E_{\rm{m}}]. The dependence on ɛ of the normalized differential cross section [p_{E_{\rm{X}}}] can be expressed as (Telnov, 2000[Telnov, V. (2000). Nucl. Instrum. Methods Phys. Res. A, 455, 80-89.])

[\eqalignno{ p_{E_{\rm{X}}}(\varepsilon) & = {{{\rm{d}}p}\over{{\rm{d}}\varepsilon_{\vphantom{\big|}}}} = {{3}\over{2}} \left(1-2\varepsilon + 2\varepsilon^2 \right), & (2a) \cr \varepsilon & \equiv {{E_{\rm{X}}}/{E_{\rm{m}}}}. & (2b) }]

Equations (1a), (1b)[link], (2a) and (2b)[link] will be the key relations used to establish the analytic expressions for the X-ray flux in a given spectral bandwidth or/and within a given angular acceptance, and to quantify the reduction of the angular and spectral fluxes due to the energy spread and the angular divergence of the two colliding beams.

The Compton X-ray sources involve multi-collisions between a laser pulse and an electron bunch. Thus, to establish the total X-ray flux, we need to take into account the laser pulse and the electron bunch sizes. Let [n_{\rm{e}}] and [n_{\rm{L}}] be, respectively, the number of electrons per bunch and the number of photons per laser pulse, and let [x_{\rm e,L}], [y_{\rm e,L}] and [z_{\rm e,L}] be the transverse and longitudinal dimensions (r.m.s.) of the electron bunch and laser pulse at the interaction point. Assuming the electron beam and the laser have Gaussian distributions in the three dimensions, for electron–photon collisions taking place with a crossing angle [\theta_{\rm{c}}] in the xz plane, the number of X-ray produced per second is (Suzuki, 1976[Suzuki, T. (1976). KEK Technical Report 76-3, July 1976. KEK, Tsukuba, Japan.])

[F(\theta_{\rm{c}}) = {{ \Sigma_{\rm th} \, n_{\rm{e}} n_{\rm{L}} \, f_{\rm rep} }\over{ 2\pi \left[ (x_{\rm{e}}^2 + x_{\rm{L}}^2) + \tan^2({{\theta_{\rm{c}}}/{2}}) (z_{\rm{e}}^2+z_{\rm{L}}^2) \right]^{1/2} \, \left(y_{\rm{e}}^2 + y_{\rm{L}}^2\right)^{1/2} }}. \eqno(3)]

[F(\theta_{\rm{c}})] is proportional to the Thomson cross section [\Sigma_{\rm th}], the luminosity and the repetition frequency of the interactions [f_{\rm rep}]. High X-ray fluxes are foreseen in various projects. For example, ThomX (Variola et al., 2014[Variola, A., Haissinkski, J., Loulergue, A. & Zomer, F. (2014). ThomX Technical Design Report. LAL/RT 14-21. SOLEIL/SOU-RA-3629.])) foresees F(0) ≃ 1013 photons s−1 with [E_{\rm{m}}] = 45 keV for a 50 MeV electron bunch ([\gamma] = 100) of 1 nC, micrometric wavelength laser pulses of 10 mJ, [f_{\rm rep}] = 20 MHz, and [x_{\rm e,L},y_{\rm e,L}] ≃ 40 µm.

Expression (3)[link] will be our reference flux hereafter referred to as [F_0 \equiv F(\theta_{\rm{c}})].

2.2. X-ray spectral bandwidth

Relations (1a) and (1b)[link] imply that the X-ray energy spectrum produced by electrons and photons whose energies are single-valued is strictly monochromatic at a given emission angle. But the energy dispersion and the angular divergence of the stored electrons and, to a lesser extent, of the photons at the interaction point lead to a significant broadening of [E_{\rm{X}}] at a given scattering angle [\theta_{\rm{X}}].

The relative energy spreads (r.m.s.) of an electron bunch and of a laser pulse will be denoted by [\sigma_{\rm{e}}] and [\sigma_{\rm{L}}], while their angular divergences (r.m.s.) will be denoted by [\sigma_{\rm{e}}^{\,\prime}] and [\sigma_{\rm{L}}^{\,\prime}] (assuming that the horizontal and the vertical divergences are equal). The order of magnitude of the broadening [\Delta E_{\rm{X}}/E_{\rm{X}}] of the on-axis Compton spectrum due to each one of these four parameters can be estimated from equations (1a) and (1b)[link] by replacing γ, [E_{\rm{L}}], [\theta_{\rm{X}}^{\,2}] and [\theta_{\rm{c}}] with [\gamma(1+\sigma_{\rm{e}})], [E_{\rm{L}}(1+\sigma_{\rm{L}})], [\sigma_{\rm{e}}^{\,\prime\,2}] and [\theta_{\rm{c}} + \sigma_{\rm{L}}^{\,\prime}], respectively, and performing four Taylor series expansions of equation (1a)[link] in [\sigma_{\rm{e}}], [\sigma_{\rm{L}}], [\sigma_{\rm{e}}^{\,\prime\,2}] and [\sigma_{\rm{L}}^{\,\prime}], resulting in: [\Delta E_{\rm{X}}/E_{\rm{X}}] [\approx] [2\sigma_{\rm{e}}], [\Delta E_{\rm{X}}/E_{\rm{X}}] [\approx] [\sigma_{\rm{L}}], [\Delta E_{\rm{X}}/E_{\rm{X}}] [\approx] [\gamma^{\,2}\sigma_{\rm{e}}^{\,\prime\,2}] and [\Delta E_{\rm{X}}/E_{\rm{X}}] [\approx] [(\cos\theta_{\rm{c}} \ \sigma_{\rm{L}}^{\,\prime\,2} /2 + \sin\theta_{\rm{c}} \ \sigma_{\rm{L}}^{\,\prime}) /(1+ \cos\theta_{\rm{c}})] due to [\sigma_{\rm{e}}], [\sigma_{\rm{L}}], [\sigma_{\rm{e}}^{\,\prime}] and [\sigma_{\rm{L}}^{\,\prime}], respectively.

We will give now orders of magnitude for these [\sigma_{\rm{e}}], [\sigma_{\rm{L}}], [\sigma_{\rm{e}}^{\,\prime}] and [\sigma_{\rm{L}}^{\,\prime}] parameters for current electron guns and pulsed lasers:

(i) The normalized horizontal and vertical emittances [\gamma x_{\rm{e}} \sigma_{\rm{e}}^{\,\prime}] and [\gamma y_{\rm{e}} \sigma_{\rm{e}}^{\,\prime}] of ∼1 nC bunches delivered by good quality current electron guns are 1–5 mm mrad (Arnold & Teichert, 2011[Arnold, A. & Teichert, J. (2011). Phys. Rev. ST Accel. Beams 14, 024801.]; Rao & Dowell, 2013[Rao, T. & Dowell, D. H. (2013). An Engineering Guide to Photoinjectors, https://arxiv.org/ftp/arxiv/papers/1403/1403.7539.pdf.]). With these values, transverse sizes and normalized divergences are typically [x_{\rm{e}},y_{\rm{e}}] ≃ 20–100 µm and [\gamma \sigma_{\rm{e}}^{\,\prime}] ≃ 0.01–0.25 rad, respectively.

(ii) Concerning [\sigma_{\rm{e}}], the electron beam relative energy spread, current values at electron accelerators are of a few [10^{-3}] up to a few percent.

(iii) On the laser side, the product of [\Delta E_{\rm{L}}], the r.m.s. energy bandwidth of the pulse, and [\Delta t_{\rm{L}}], its r.m.s. temporal duration, is constrained by the uncertainty principle (Donnelly & Grossman, 1998[Donnelly, T. D. & Grossman, C. (1998). Am. J. Phys. 66, 677.]) [\Delta E_{\rm{L}} \Delta t_{\rm{L}}] [\geq] [\hbar/2]. Unchirped pulses have the minimum time-bandwidth product, i.e. close to [\hbar/2], whereas larger values prevail for chirped pulses. Then, let us consider a laser with a wavelength λ, and an r.m.s. pulse duration [\Delta t_{\rm{L}}]. Assuming unchirped pulses, one derives from the uncertainty principle: [\sigma_{\rm{L}}] [\equiv] [\Delta E_{\rm{L}}/E_{\rm{L}}] = [\lambda/(4 \pi c \Delta t_{\rm{L}})], where c is the speed of light. It follows that infrared fs–ps pulses have a relative energy spread [\sigma_{\rm{L}}] of a few [10^{-5}] (for 10 ps pulses) up to a few [10^{-3}] (for 100 fs pulses).

(iv) In good quality lasers with Gaussian pulses, the relation between the r.m.s. transverse pulse sizes [x_{\rm{L}}], [y_{\rm{L}}] and the angular divergence [\sigma_{\rm{L}}^{\,\prime}] is [x_{\rm{L}} \sigma_{\rm{L}}^{\,\prime}] = [y_{\rm{L}} \sigma_{\rm{L}}^{\,\prime}] = [\lambda/(4\pi)] (Schmüser et al., 2008[Schmüser, P., Dohlus, M. & Rossbach, J. (2008). Ultraviolet and Soft X-ray Free-Electron Lasers. Berlin: Springer.]), leading to the following typical values: [x_{\rm{L}},y_{\rm{L}}] ≃ 10–100 µm and [\sigma_{\rm{L}}^{\,\prime}] ≃ 1–10 mrad (infrared laser).

Table 1[link] summarizes these orders of magnitude and the corresponding broadening of the Compton spectrum [\Delta E_{\rm{X}}/E_{\rm{X}}]. The spectrum broadening due to the laser bandwidth [\sigma_{\rm{L}}] is negligible, as well as the broadening due to the laser divergence if the product [\sin\theta_{\rm{c}} \sigma_{\rm{L}}^{\,\prime}] is smaller than ∼1 mrad, which is the case in almost all current Compton projects where electrons and laser photons collide head-on (Jacquet, 2014[Jacquet, M. (2014). Nucl. Instrum. Methods Phys. Res. B, 331, 1-5.]). We conclude that [\Delta E_{\rm{X}}/E_{\rm{X}}] is mainly governed by the angular divergence and the relative energy spread of the electron beam.

Table 1
Orders of magnitude of the relative energy spread and divergence of electron and laser beams at the interaction point, and associated broadening of the Compton spectrum

Variable Typical values [\Delta E_{\rm{X}}/E_{\rm{X}}]
[\sigma_{\rm{e}}] [10^{-3}][10^{-2}] [2\times10^{-3}][2\times10^{-2}]
[\sigma_{\rm{L}}] [10^{-5}][10^{-3}] [10^{-5}][10^{-3}]
[\gamma\sigma_{\rm{e}}^{\,\prime}\,\,{\rm{(rad)}}] 0.01–0.25 [10^{-4}][10^{-1}]
[\sigma_{\rm{L}}^{\,\prime}\,\,{\rm{(mrad)}}] 1–10 [10^{-7}][10^{-3}] ([\sin\theta_{\rm{c}}\sigma_{\rm{L}}^{\,\prime}\,\,\lesssim1\,\,{\rm{mrad)}}]
[10^{-3}][10^{-2}] ([\sin\theta_{\rm{c}}\sigma_{\rm{L}}^{\,\prime}] ≃ 1–10 mrad)

2.3. Input and notations

In this section, we introduce the probability distributions used to establish the analytic expressions of §3[link] and §4[link]. The nominal Lorentz factor of the electron beam will be referred to as [\gamma_0]. Then the Lorentz factor of a given electron is written as [\gamma] = [(1 + \delta\gamma)\gamma_0], where [\delta\gamma] is the difference in energy of this electron with respect to the nominal beam energy ( E0) divided by E0. With [\theta_{\rm{c}}] being the nominal value of the crossing angle between the electron beam and the laser in the xz plane (see Fig. 1[link]), we define [\zeta_{{\rm{e}}x}], [\zeta_{{\rm{e}}y}], [\theta_{\rm{c}} + \zeta_{{\rm{L}}x}] and [\zeta_{{\rm{L}}y}] as the angles to z axis of the projections of the momentum of a given electron and of a given laser photon in the xz and yz planes. The probability distributions of [\delta\gamma], [\zeta_{{\rm{e}}x,{\rm{e}}y}] and [\zeta_{{\rm{L}}x,{\rm{L}}y}] are assumed to be Gaussian with standard deviations [\sigma_{\rm{e}}], [\sigma_{{\rm{e}}x,{\rm{e}}y}^{\,\prime}] and [\sigma_{{\rm{L}}x,{\rm{L}}y}^{\,\prime}], respectively. We further assume [\sigma_{{\rm{e}}x}^{\,\prime}] = [\sigma_{{\rm{e}}y}^{\,\prime}] = [\sigma_{\rm{e}}^{\,\prime}] and [\sigma_{{\rm{L}}x}^{\,\prime}] = [\sigma_{{\rm{L}}y}^{\,\prime}] = [\sigma_{\rm{L}}^{\,\prime}]. [\zeta_{{\rm{eL}}x}] ([\zeta_{{\rm{eL}}y}], respectively) is defined as the convolution of the [\zeta_{{\rm{e}}x}] and [\zeta_{{\rm{L}}x}] distributions (of [\zeta_{{\rm{e}}y}] and [\zeta_{{\rm{L}}y}], respectively) and follows a Gaussian distribution whose standard deviation is [\sigma_{\rm{eL}}^{\,\prime}] = [(\sigma_{\rm{e}}^{\,\prime\,2} + \sigma_{\rm{L}}^{\,\prime\,2})^{1/2}].

We introduce also the polar and azimuthal angles with respect to the z-axis direction (see Fig. 2[link]), [\zeta_{\rm{e}}] and [\varphi_{\rm{e}}], of an electron momentum ([\zeta_{{\rm{e}}x}] = [\zeta_{\rm{e}} \cos\varphi_{\rm{e}}], [\zeta_{{\rm{e}}y}] = [\zeta_{\rm{e}} \sin\varphi_{\rm{e}}], [\zeta_{\rm{e}}^2] = [\zeta_{{\rm{e}}x}^2 + \zeta_{{\rm{e}}y}^2]), and similarly [\zeta_{\rm{X}}] and [\varphi_{\rm{X}}] for a photon scattered by this electron. For small [\zeta_{\rm{e}}] and [\zeta_{\rm{X}}] (i.e. when [\sin\zeta_{\rm{e}}][\zeta_{\rm{e}}] and [\sin\zeta_{\rm{X}}][\zeta_{\rm{X}}]), the polar angle [\theta_{\rm{X}}] of the backscattered photon with respect to the electron momentum is [\theta_{\rm{X}}^{\,2}] = [\zeta_{\rm{e}}^2] + [\zeta_{\rm{X}}^2] [-] [2\zeta_{\rm{e}}\zeta_{\rm{X}}\cos(\varphi_{\rm{e}}-\varphi_{\rm{X}})]. Summing over azimuthal angles we obtain

[\theta_{\rm{X}}^{\,2} = \zeta_{\rm{e}}^2 + \zeta_{\rm{X}}^2. \eqno(4)]

The [\theta_{\rm{X}}] angle averaged over [\varphi_{\rm{e}}] and [\varphi_{\rm{X}}] expressed in (4)[link] is the one that appears in equation (1a)[link].

[Figure 2]
Figure 2
Polar angles of an electron momentum ([\zeta_{\rm{e}}]) and of a Compton photon ([\zeta_{\rm{X}}]) scattered by this electron with respect to the z-axis, and polar angle [\theta_{\rm{X}}] of the Compton photon with respect to the electron momentum. Note that the electron momentum is practically not affected by the collision.

Explicitly, the probability distributions of [\delta\gamma], [\zeta_{{\rm{eL}}x,{\rm{eL}}y}] and [\zeta_{\rm{e}}^2] defined in the ranges []\!-\infty,+\infty[\,], []\!-\infty,+\infty[] and [[0+\infty[\,], respectively, are

[\eqalignno{ {{ {\rm{d}}p }\over{ {\rm{d}}(\delta\gamma) }} & = {{1}\over{\sqrt{2\pi}\sigma_{\rm{e}}}} \, \exp\left[{-{(\delta\gamma)^2}/{2\sigma_{\rm{e}}^2}}\right], & (5a) \cr {{ {\rm{d}}p }\over{ {\rm{d}}\zeta_{{\rm{eL}}x,{\rm{eL}}y} }} & = {{1}\over{\sqrt{2\pi}\sigma_{\rm{eL}}^{\,\prime}}} \, \exp\left[ -(\zeta_{{\rm{eL}}x,{\rm{eL}}y})^2/2\sigma_{\rm{eL}}^{\,\prime\,2} \right], &(5b) \cr {{ {\rm{d}}p}\over{{\rm{d}}(\zeta_{\rm{e}}^2)}} & = {{1}\over{2\sigma_{\rm{e}}^{\prime\,2}}} \, \exp\left(-{\zeta_{\rm{e}}^2}/{2\sigma_{\rm{e}}^{\,\prime\,2}}\right). & (5c) }]

In the following sections we will obtain analytic expressions for the ratios [R_{\rm{tot}}] = [F_{\rm{tot}}/F_0], [R_{\alpha}] = [F_{\alpha}/F_0], [R_{\rm{bw}}] = [F_{\rm{bw}}/F_0] and [R_{\alpha,{\rm{bw}}}] = [F_{\alpha,{\rm{bw}}}/F_0], where [F_{\rm{tot}}] is the total flux, [F_{\alpha}] the flux in a given angular acceptance α (of a few milliradians) around the z-axis direction, [F_{\rm{bw}}] the flux in a given energy bandwidth (bw) centered at the on-axis X-ray energy [E_{\rm{m}}^{\,0}] = [2\gamma_0^2\,E_{\rm{L}} (1 + \cos \theta_{\rm{c}})], and [F_{\alpha,{\rm{bw}}}] the flux in a given angular acceptance and a given energy bandwidth. We will study the dependence of these ratios on the [\sigma_{\rm{e}}^{\,\prime}] and [\sigma_{\rm{e}}] parameters which govern the broadening of the Compton angular spectrum as discussed in §2.2[link], and we will confront our formulas with simulations performed with the CAIN code (Yokoya, 2003[Yokoya, K. (2003). User's Manual of CAIN, https://lcdev.kek.jp/~yokoya/CAIN/cain235/CainMan235.pdf. ]) generated with several [\sigma_{\rm{e}}^{\,\prime}] and [\sigma_{\rm{e}}] values, all other parameters ([n_{\rm{e}}], [n_{\rm{L}}], [f_{\rm{rep}}] [x_{\rm{e}}], [y_{\rm{e}}], [x_{\rm{L}}], [y_{\rm{L}}], [z_{\rm{e}}] and [z_{\rm{L}}]) remaining fixed. In the CAIN simulations, we use [\gamma_0] = 100 (i.e. the order of magnitude to produce Compton backscattered photons in the X-ray domain), a laser wavelength [\lambda] = 1 µm, and we assume a pulse waist of 40 µm. For a full range coverage of the electron beam parameters in current accelerators (see Table 1[link]), the [\sigma_{\rm{e}}^{\,\prime}] and [\sigma_{\rm{e}}] ranges used for this work are [\gamma\sigma_{\rm{e}}^{\,\prime}] ≃ 0.01–0.25 rad (i.e. [\sigma_{\rm{e}}^{\,\prime}] ≃ 0–2.5 mrad for [\gamma] ≃ 100) and [\sigma_{\rm{e}}] ≃ 0–2%.

3. Spectral or angular selection

In this section, we will obtain the expression of the X-ray flux when one selects either the energy or the emission angle of the X-rays.

3.1. Dependence of the total flux Ftot on the electron and the laser beam angular divergences

Since we assume here that no cut is applied to the X-ray energy, the electron beam energy spread does not come into this calculation. Thus we treat here only the electron beam and the laser beam divergences. We consider the collision of an electron with a photon at a nominal crossing angle [\theta_{\rm{c}}] in the horizontal plane xz. The effect of the two [\zeta_{{\rm{eL}}x}] and [\zeta_{{\rm{eL}}y}] variables (defined in §2.3[link]) can be treated as an additional contribution to [\theta_{\rm{c}}]. Indeed, for small [\zeta_{{\rm{eL}}x}] and [\zeta_{{\rm{eL}}y}] (i.e. when [\sin\zeta_{{\rm{eL}}x}][\zeta_{{\rm{eL}}x}] and [\sin\zeta_{{\rm{eL}}y}][\zeta_{{\rm{eL}}y}]) and by making the approximation that [\zeta_{{\rm{eL}}x}] and [\zeta_{{\rm{eL}}y}] are awarded only to the electron (see Fig. 3[link]), a geometrical calculation leads to an effective electron–photon crossing angle [\Theta_{\rm{c}}] such that [\Theta_{\rm{c}}] = [{\rm{acos}}\{ \zeta_{{\rm{eL}}x} \sin\theta_{\rm{c}}] + [\cos\theta_{\rm{c}} [1 - (\zeta_{{\rm{eL}}x}^2 +\zeta_{{\rm{eL}}y}^2)]^{1/2}\}.]

[Figure 3]
Figure 3
Nominal ([\theta_{\rm{c}}]) and effective ([\Theta_{\rm{c}}]) electron–photon crossing angles assuming that [\zeta_{{\rm{eL}}x}] and [\zeta_{{\rm{eL}}y}] are awarded only to the electron.

Then, for an electron bunch and a laser pulse of r.m.s. angular divergence [\sigma_{\rm{e}}^{\,\prime}] and [\sigma_{\rm{L}}^{\,\prime}], respectively, the total flux [F_{\rm{tot}}] [\equiv] [F(\theta_{\rm{c}})] is obtained by integrating [F(\Theta_{\rm{c}})], weighted by the Gaussian probabilities of [\zeta_{{\rm{eL}}x}] and [\zeta_{{\rm{eL}}y}] [see equation (5b)[link]]. Keeping up to the second-order terms in [\zeta_{{\rm{eL}}x}] and [\zeta_{{\rm{eL}}y}] results in the following expression for the ratio [R_{\rm{tot}}] [\equiv] [F_{\rm{tot}}/F_0]:

[\eqalignno{ R_{\rm{tot}} = {}& 1 - {{ \left[1 + \tan^2({{\theta_{\rm{c}}}/{2}})\right]^2 \,\left[1 - {{1} \over {2}} r_{xz} \tan^2({{\theta_{\rm{c}}}/{2}})\right] }\over{ 4 \left[1+r_{xz} \tan^2({{\theta_{\rm{c}}}/{2}})\right]^2_{\vphantom{\big|}} }} \cr& \times r_{xz}\left(\sigma_{\rm{e}}^{\,\prime\,2}+\sigma_{\rm{L}}^{\,\prime\,2} \right), &(6)}]

where rxz = [(z_{\rm{e}}^2+z_{\rm{L}}^2)/(x_{\rm{e}}^2+x_{\rm{L}}^2)]. For [x_{\rm{e}}] = 72 µm, [x_{\rm{L}}] = 40 µm, [z_{\rm{e}}] = 4.8 mm, [z_{\rm{L}}] = 3 mm and [\sigma_{\rm{L}}^{\,\prime}] = 2 mrad (i.e. the laser beam divergence value corresponding to a 40 µm waist size), [R_{\rm{tot}}] is shown in Fig. 4[link] as a function of [\sigma_{\rm{e}}^{\,\prime}], for [\theta_{\rm{c}}] = 0°, 1° and 2°, and compared with results obtained with the CAIN simulation program. The analytic expression reproduces well the differences between the various collision angle cases. A reduced collision angle amplifies the divergence impact on the total flux but only to a small extent, less than 1% for head-on collisions.

[Figure 4]
Figure 4
The ratio [R_{\rm{tot}}] = [F_{\rm{tot}}/F_0] as a function of [\sigma_{\rm{e}}^{\,\prime}] for a laser divergence [\sigma_{\rm{L}}^{\,\prime}] = 2 mrad, for [\theta_{\rm{c}}] = 0° (black bullets, black line), 1° (red squares, red line) and 2° (blue open circle, blue line). Points are CAIN simulations. For each point, the error bar indicates the statistical uncertainty calculated from the number of generated events (namely ∼4.7 × 105, 4.1 × 105 and 3.1 × 105 events for each of the eight points generated with [\theta_{\rm{c}}] = 0°, 1° and 2°, respectively). Lines illustrate the analytic formula [equation (6)[link]] results.

3.2. X-ray flux in a given energy bandwidth (Fbw)

In this section we look at the number of X-rays produced when an energy cut is applied. For this, the electron energy spread is decisive whereas [F_{\rm{bw}}/F_{\rm{tot}}] is independent of the electron beam angular divergence since no angular selection is assumed here. Also, we do not take into account the broadening of the Compton spectrum ([\Delta E_{\rm{X}}/E_{\rm{X}}][10^{-3}][10^{-2}]) that occurs in the particular case where both the laser divergence and the collision angle are large (see Table 1[link]).

Let us consider an electron of Lorentz factor [\gamma] = [\gamma_0 (1+\delta\gamma)]. The energy of a photon which is backscattered by this electron is [E_{\rm{X}}] and, according to equation (1b)[link], the maximum value of [E_{\rm{X}}] is [E_{\rm{m}}] = [E_{\rm{m}}^{\,0} (1+\delta\gamma)^2] where [E_{\rm{m}}^{\,0}] = [2\gamma_0^2 E_{\rm{L}} (1 + \cos \theta_{\rm{c}})]. The ratio [E_{\rm{X}}/E_{\rm{m}}^{\,0}] is denoted by [\varepsilon_0]. [F_{\rm{bw}}] is the flux of photons whose [\varepsilon_0] lies in the interval between 1 − bw and 1 + bw. The energy probability distribution (2a)[link] expressed as a function of [\varepsilon_0] becomes

[p_{E_{\rm{X}}}(\varepsilon_0) = {{3} \over {2}} \left[ 1-{{2\varepsilon_0} \over {(1 + \delta\gamma)^2 }} + {{2 \varepsilon_0^2} \over {(1 + \delta\gamma)^4 }} \right] {{1} \over {(1 + \delta\gamma)^2 }}. \eqno(7)]

Fig. 5[link] shows this distribution for three different cases: [\delta\gamma] < 0, [\delta\gamma] = 0 and [\delta\gamma] > 0. Since [p_{E_{\rm{X}}}] is a quasi-linear function for [\varepsilon_0] larger than ∼0.65, the integral of [p_{E_{\rm{X}}}] between [\varepsilon_{0_1}] and [\varepsilon_{0_2}] can be approximated by [p_{E_{\rm{X}}}[(\varepsilon_{0_1}+\varepsilon_{0_2})/2](\varepsilon_{0_2}-\varepsilon_{0_1})]. Thus [F_{\rm{bw}}] is calculated in the following way (see Fig. 5[link]):

[Figure 5]
Figure 5
Differential cross section [p_{E_{\rm{X}}}] for three electron populations: electrons with positive [\delta\gamma] (dotted line), with [\delta\gamma] = 0 (solid line), and with negative [\delta\gamma] (dashed line), as a function of the scattered photon energy [\varepsilon_0] = [E_{\rm{X}}/E_{\rm{m}}^{\,0}]. Also represented are [f_{\rm{bw}}^{\,{\rm{s}}}] (green area) and [f_{\rm{bw}}^{\,{\rm{L}}}] (area hatched in green).

(i) Electrons with [\delta\gamma] such that [(1 + \delta\gamma)^2] < 1 − bw do not contribute to [F_{\rm{bw}}].

(ii) The contribution to [F_{\rm{bw}}] of electrons having a relative energy difference [\delta\gamma] such that 1 − bw < [(1+\delta\gamma)^2] < 1 + bw is denoted by [f_{\rm{bw}}^{\,{\rm{s}}}(\delta\gamma)] and is equal to the integral of [p_{E_{\rm{X}}}] between [\varepsilon_{0_1}] = 1 − bw and [\varepsilon_{0_2}] = [(1+\delta\gamma)^2] (green area): [f_{\rm{bw}}^{\,{\rm{s}}}(\delta\gamma)] = [p_{E_{\rm{X}}} \{[(1-{\rm{bw}})+(1+\delta\gamma)^2]/2\}[(1+\delta\gamma)^2 - (1-{\rm{bw}})]]. These electrons are denoted by [\delta\gamma_{\rm{s}}] in Fig. 5[link].

(iii) The contribution of electrons whose [\delta\gamma] satisfies 1 + bw < [(1 + \delta\gamma)^2] is denoted by [f_{\rm{bw}}^{\,{\rm{L}}}(\delta\gamma)] and is equal to the integral of [p_{E_{\rm{X}}}] between [\varepsilon_{0_1}] = 1 − bw and [\varepsilon_{0_2}] = 1 + bw (area hatched in green): [f_{\rm{bw}}^{\,{\rm{L}}}(\delta\gamma)] = [p_{E_{\rm{X}}} (1)\ 2{\rm{bw}}]. These electrons are denoted by [\delta\gamma_{\rm{L}}] in Fig. 5[link].

Assuming that the [\delta\gamma] variable follows the Gaussian distribution (5a)[link] where [\sigma_{\rm{e}}] ≃ 0–2%, one can make the approximation [(1+\delta\gamma)^2][1+2\delta\gamma] and [R_{\rm{bw}}] can be written

[\eqalign{ R_{\rm{bw}} = {}& R_{\rm{tot}} \Bigg[ \int_{-{\rm{bw}}/2}^{{\rm{bw}}/2} {{1}\over{\sqrt{2\pi}\sigma_{\rm{e}}}} \exp\left(-\delta\gamma^{\,2}/2\sigma_{\rm{e}}^2\right) \, f_{\rm{bw}}^{\,{\rm{s}}}(\delta\gamma) \,{\rm{d}}(\delta\gamma) \cr& + \int_{{\rm{bw}}/2}^{+\infty} {{1}\over{\sqrt{2\pi}\sigma_{\rm{e}}}} \exp\left(-\delta\gamma^{\,2}/2\sigma_{\rm{e}}^2\right) \, f_{\rm{bw}}^{\,{\rm{L}}}(\delta\gamma) \,{\rm{d}}(\delta\gamma) \Bigg].}]

Performing a first-order Taylor expansion in [\delta\gamma], the calculation of the above integrals leads to

[\eqalignno{ R_{\rm{bw}} = {}& {{3}\over{2}} \, R_{\rm{tot}} \, {\rm{bw}} \Big\{ 1- \left({\rm{bw}}-{{{\rm{bw}}^2}/{2}}\right) \, {\rm{ERF}}\left[{{{\rm{bw}}}/({2\sqrt{2}\sigma_{\rm{e}}})}\right] \cr& -6 \sqrt{{{2}/{\pi}}} \, \sigma_{\rm{e}} \exp\left(-{\rm{bw}}^2/8\sigma_{\rm{e}}\right) \Big\} ,&(8)}]

and is shown in Fig. 6[link] as a function of [\sigma_{\rm{e}}], for three energy bandwidths: (a) bw = 0.1%, (b) bw = 0.7% and (c) bw = 3%.

[Figure 6]
Figure 6
The ratio [R_{\rm{bw}}] = [F_{\rm{bw}}/F_0] as a function of [\sigma_{\rm{e}}] for three energy bandwidths: (a) bw = 0.1%, (b) bw = 0.7% and (c) bw = 3%. Points are CAIN simulations and the error bars indicate the statistical uncertainties. Lines are obtained using the analytic formula (8)[link].

First of all, in the absence of any energy spread, one notes a factor 3/2 compared with the classical formula used for synchrotron radiation, showing the need for a proper analytic development when dealing with Compton sources. This factor, which comes from the Compton cross section, remains whatever the selected energy acceptance. Fig. 6[link] shows that a linear behavior of the flux reduction with [\sigma_{\rm{e}}] prevails if the selected bandwidth is small compared with [\sigma_{\rm{e}}] (as in the 0.1% and 0.7% cases and in the right-hand region of the 3% case), whereas the flux remains globally constant when the energy acceptance is large with respect to [\sigma_{\rm{e}}] (as in the left-hand region of the 3% case). In the three bandwidth cases, the small systematic difference between the CAIN simulation values and the analytic results for [\sigma_{\rm{e}}] ≃ 2% shows that higher-order terms in [\delta\gamma] have to be taken into account for [\sigma_{\rm{e}}] values larger than 2–3%. In any event, the reduction due to the energy spread is less than 10% in the parameter range considered here.1

3.3. X-ray flux in a given angular acceptance

We now consider the photon flux [F_{\alpha}] within a given angular acceptance α. In this case, the effect of an energy spread is completely negligible since the mean value of [\delta\gamma] vanishes.

For any value of the polar angle θ, the [\gamma_0^2\theta^2] product is denoted as [\mu_{\theta}]. Let us consider an electron whose momentum makes a polar angle [\zeta_{\rm{e}}] with respect to the z axis (see Fig. 2[link]) while the photon backscattered by this electron has a polar angle [\zeta_{\rm{X}}]. Using equations (1a)[link], (2a)[link] and (4)[link], the dependence of the normalized differential Compton cross section on [\mu_{\zeta_{\rm{X}}}] is given by

[p_{\mu}(\mu_{\zeta_{\rm{X}}}) = {{{\rm{d}}}\over{{\rm{d}}\mu_{\zeta_{\rm{X}}}}} = {{3}\over{2}} \,\, {{1+(\mu_{\zeta_{\rm{X}}}+\mu_{\zeta_{\rm{e}}})^2 }\over{ (1+\mu_{\zeta_{\rm{X}}}+\mu_{\zeta_{\rm{e}}})^4 }}. \eqno(9)]

This distribution is shown in Fig. 7[link] for [\mu_{\zeta_{\rm{e}}}] = 0 and [\mu_{\zeta_{\rm{e}}}] = 0.04. Here again, with [p_{\mu}] being a quasi-linear function of [\mu_{\zeta_{\rm{X}}}] for [\mu_{\zeta_{\rm{X}}}] smaller than ∼0.15, its integral between [\mu_{\zeta_{X_1}}] and [\mu_{\zeta_{X_2}}] can be approximated by [p_{\mu}\{[\mu_{\zeta_{X_1}}+\mu_{\zeta_{X_2}}]/2\}(\mu_{\zeta_{X_1}}-\mu_{\zeta_{X_2}})]. Then, as illustrated by Fig. 7[link], an electron contribution to [F_{\alpha}] is equal to [\mu_{\alpha}\,p_{\mu}(\mu_{\alpha}/2)] and, according to the exponential distribution (5c)[link] of [\zeta_{\rm{e}}^2], the ratio [R_{\alpha}] for an electron bunch of r.m.s. divergence [\sigma_{{\rm{e}}x}^{\,\prime}] = [\sigma_{{\rm{e}}y}^{\,\prime}] = [\sigma_{\rm{e}}^{\,\prime}] can be simply expressed as

[R_{\alpha} = R_{\rm{tot}}\, \mu_{\alpha} \int\limits_{0}^{{+\infty}} {{1}\over{2\sigma_{\rm{e}}^{\,\prime\,2}}} \, \exp\left(-\zeta_{\rm{e}}^2/2\sigma_{\rm{e}}^{\,\prime\,2}\right) \, p_{\mu}(\mu_{\alpha}/2) \, {\rm{d}}(\zeta_{\rm{e}}^2). \eqno(10)]

The calculation of equation (10)[link] using a first-order Taylor series expansion in [\zeta_{\rm{e}}^2] leads to

[\eqalignno{ R_{\alpha} = {}& {{3}\over{2}}\, R_{\rm{tot}}\, {{\mu_{\alpha}}\over{(1+{\mu_{\alpha}}/2)^2}} \cr& \times \left[1 + {{\mu_{\alpha}^2}\over{4}} - 4\gamma_0^{\,2}\sigma_{\rm{e}}^{\,\prime\,2} \,\, {{ 2-\mu_\alpha/2+ \mu_\alpha^2/4 }\over{ 1+\mu_\alpha/2 }} \right]. &(11)}]

Expression (11)[link] is valid as long as [\mu_{\alpha}] and [\gamma_0^{\,2}\sigma_{\rm{e}}^{\,\prime\,2}] are less than ∼0.15 and ∼0.03, respectively. For larger values of [\mu_{\alpha}], the linear approximation of [p_{\mu}] is no longer valid and the integral between 0 and [\mu_{\alpha}] must be explicitly calculated. For [\gamma_0^{\,2}\sigma_{\rm{e}}^{\,\prime\,2}] [\gtrsim] 0.03, higher-order terms in [\zeta_{\rm{e}}^2] have to be taken into account in the series expansion of equation (10)[link]. [R_{\alpha}] and points obtained with CAIN are shown in Fig. 8[link] as a function of [\sigma_{\rm{e}}^{\,\prime}] (for [\gamma_0] = 100) for an anglular acceptance α of 1 mrad and 2 mrad. Equation (11)[link] is represented by the solid lines and is in very good agreement with CAIN simulations up to [\sigma_{\rm{e}}^{\,\prime}] ≃ 1.5 mrad. The need to take into account higher-order terms in [\zeta_{\rm{e}}^2] in the modeling of [R_{\alpha}] for large values of [\sigma_{\rm{e}}^{\,\prime}] is highlighted for [\sigma_{\rm{e}}^{\,\prime}] ≃ 2 mrad. The dotted lines indicate the analytic expression results when the third-order term in [\zeta_{\rm{e}}^2] is taken into account.2

[Figure 7]
Figure 7
The angular differential cross section [p_{\mu}] as a function of [\mu_{\zeta_{\rm{X}}}] for [\mu_{\zeta_{\rm{e}}}] = 0 (solid line) and for [\mu_{\zeta_{\rm{e}}}] = 0.04 (dashed line). The shaded area shows the contribution to [F_{\alpha}] of an electron whose momentum has a polar angle [\zeta_{\rm{e}}].
[Figure 8]
Figure 8
The ratio [R_{\alpha}] = [F_{\alpha}/F_0] as a function of [\sigma_{\rm{e}}^{\,\prime}] for an angular acceptance [\alpha] of 1 mrad (open red circles and red line) and 2 mrad (black bullets and black line). Points are CAIN simulations (the error bars are too small to be visible). Solid lines are obtained with equation (11)[link], dotted lines show [R_{\alpha}] calculated to third order in [\zeta_{\rm{e}}^2].

4. X-ray flux within a given angular acceptance and a given energy bandwidth

In this section we assume that both an angular cut and an energy selection are applied to the backscattered photons and we calculate the consequent reduction of the X-ray flux. In the first two subsections, the electron beam divergence and its energy spread will be treated independently to understand the impact of each one of these parameters. Then we will take into account both of these beam characteristics.

4.1. Dependence on the electron beam angular spread

Our starting point is equation (8)[link] which gives the X-ray yield in a given energy bandwidth, and we will modify this expression to take into account the angular acceptance defined by a maximum value α of the X-ray emission direction with respect to the z axis and the electron beam angular spread. In order to achieve a single formulation, we associate an angle to the energy bandwidth, namely [\theta_{\rm{bw}}] = [\sqrt{\rm bw} / \gamma_0]. At this stage, we have to treat separately the [\alpha] < [\theta_{\rm{bw}}] case and the [\alpha] > [\theta_{\rm{bw}}] case, as illustrated in Fig. 9[link]. The dashed red line represents the electron momentum direction which makes an angle [\zeta_{\rm{e}}] with respect to the z direction. α defines the angular cut and [\theta_{\rm{bw}}] the energy acceptance translated into an angle. More precisely, to the energy bandwidth (or energy acceptance) bw corresponds an angular acceptance which is a cone whose axis is the momentum of the electron which backscatters the incoming photon and whose half aperture is [\theta_{\rm{bw}}]. In Fig. 9[link], those X-rays which pass both the angular and the energy cuts are the ones emitted in the green domain. The hatched regions show emission directions which are outside the X-ray angular acceptance cone of aperture α. The accepted region depends on the three angles involved in the following way:

[Figure 9]
Figure 9
Top: the [\alpha] > [\theta_{\rm{bw}}] case where [\theta_{\rm{bw}}] = [\sqrt{\rm{bw}}/\gamma_0]. Three configurations are shown: (a) [\zeta_{\rm{e}}] < [\alpha-\theta_{\rm{bw}}], (b) [\alpha-\theta_{\rm{bw}}] < [\zeta_{\rm{e}}] < [\alpha], (c) [\alpha] < [\zeta_{\rm{e}}] < [\alpha+\theta_{\rm{bw}}]. Bottom: the [\alpha] < [\theta_{\rm{bw}}] case. (d) [\zeta_{\rm{e}}] < [\theta_{\rm{bw}}- \alpha], (e) [\theta_{\rm{bw}}- \alpha] < [\zeta_{\rm{e}}] < [\theta_{\rm{bw}}], (f) [\theta_{\rm{bw}}] < [\zeta_{\rm{e}}] < [\alpha+\theta_{\rm{bw}}].

(i) For [\alpha] > [\theta_{\rm{bw}}] (see upper panels in Fig. 9[link]) and values of [\zeta_{\rm{e}}] such that [\zeta_{\rm{e}}] < [\alpha-\theta_{\rm{bw}}], all the X-rays whose energy belong to the accepted bandwidth pass the angular cut [case (a)]; when [\zeta_{\rm{e}}] increases, the cone intersection diminishes and depends on whether [\zeta_{\rm{e}}] is larger than α or not [(b) and (c) cases];

(ii) For [\alpha] < [\theta_{\rm{bw}}] (see lower panels in Fig. 9[link]), the cone intersection is reduced as the electron angular spread increases and depends on whether [\zeta_{\rm{e}}] is larger than [\theta_{\rm{bw}}] or not [(e) and (f) cases].

To take into account the various cases enumerated above, we introduce the following two angles: [\theta_{\rm{m}}] = [\min(\alpha,\theta_{\rm{bw}})] and [\theta_{\rm{M}}] = [\max(\alpha,\theta_{\rm{bw}})]. Then, using the [\zeta_{\rm{e}}^2] probability distribution of equation (5c)[link], an integration is performed for each of the three cases and the number of X-rays backscattered in the angular acceptance α and within a given energy bandwidth bw is given by the following expression:

[\eqalignno{ & R_{\alpha,{\rm{bw}}}(\sigma_{\rm{e}}^{\,\prime}) = R_{\rm{bw}} \left[ \int\limits_0^{(\alpha-\theta_{\rm{bw}})^2} {{1}\over{2\sigma_{\rm{e}}^{\prime\,2}}}\, \exp\left(-\zeta_{\rm{e}}^2/2\sigma_{\rm{e}}^{\,\prime\,2}\right) \, {{\theta^{\,2}_{\rm{m}}}\over{\theta^{\,2}_{\rm{bw}}}} \, {\rm{d}}(\zeta_{\rm{e}}^2) \right. \cr& +\int\limits_{(\alpha-\theta_{\rm{bw}})^{2}}^{\theta_{\rm{M}}^{\,2}} {{1}\over{2\sigma_{\rm{e}}^{\,\prime\,2}}} \, \exp\left(-\zeta_{\rm{e}}^2/2\sigma_{\rm{e}}^{\,\prime\,2}\right) \, \left[{{\theta_{\rm{m}}^{\,2}} \over {\theta_{\rm{bw}}^{\,2}}} - {{\theta_{\rm{m}}^{\,2}-(\theta_{\rm{M}}-\zeta_{\rm{e}})^2} \over {2 \theta_{\rm{bw}}^{\,2}}} \right] \,{\rm{d}}(\zeta_{\rm{e}}^2) \cr& \left.\,+ \int\limits_{\theta_{\rm{M}}^{\,2}}^{(\alpha+\theta_{\rm{bw}})^2} {{1}\over{2\sigma_{\rm{e}}^{\,\prime\,2}}} \, \exp\left(-\zeta_{\rm{e}}^2/2\sigma_{\rm{e}}^{\,\prime\,2}\right) \, {{(\theta_{\rm{bw}}+\alpha-\zeta_{\rm{e}})^2}\over{2\theta_{\rm{bw}}^{\,2}}} \,{\rm{d}}(\zeta_{\rm{e}}^2) \right], &(12)}]

where [R_{\rm{bw}}] is given by equation (8)[link] with [\sigma_{\rm{e}}] ≃ 0, i.e. [R_{\rm{bw}}] = [({{3}/{2}})\ R_{\rm{tot}}\,{\rm{bw}}[1-({\rm{bw}}-{{{\rm{bw}}^2}\!/{2}})]]. Equation (12)[link] leads to the following analytic expression for [R_{\alpha,{\rm{bw}}}]:

[\eqalignno{ R_{\alpha,{\rm{{\rm{bw}}}}}(\sigma_{\rm{e}}^{\,\prime}) = {}& {{R_{\rm{bw}}}\over{\theta^2_{\rm{bw}}}} \Bigg( \theta_{\rm{m}}^{\,2} + \sigma_{\rm{e}}^{\,\prime\,2} \bigg\{ \exp\left[-(\alpha-\theta_{\rm{bw}})^2/2\sigma_{\rm{e}}^{\,\prime\,2}\right] \cr& - \exp\left[-(\alpha+\theta_{\rm{bw}})^2/2\sigma_{\rm{e}}^{\,\prime\,2}\right] \bigg\} \cr& + {{\sqrt{\pi}}\over{\sqrt{2}}} \ \sigma_{\rm{e}}^{\,\prime} \ \theta_{\rm{M}} \ {\rm{ERF}}\left({{\theta_{\rm{M}}-\theta_{\rm{m}}} \over {\sqrt{2}\sigma_{\rm{e}}^{\,\prime}}} \right) \cr& + {{\sqrt{\pi}}\over{\sqrt{2}}} \ \sigma_{\rm{e}}^{\,\prime} \ \theta_{\rm{m}} \ {\rm{ERF}}\left({{\theta_{\rm{M}}} \over {\sqrt{2}\sigma_{\rm{e}}^{\,\prime}}} \right) \cr& - {{\sqrt{\pi}}\over{\sqrt{2}}} \ \sigma_{\rm{e}}^{\,\prime} \ (\alpha+\theta_{\rm{bw}}) \,{\rm{ERF}}\left({{\alpha+\theta_{\rm{bw}}} \over {\sqrt{2}\sigma_{\rm{e}}^{\,\prime}}} \right) \Bigg). &(13)}]

CAIN simulations and results from equation (13)[link] are shown in Fig. 10[link] where [R_{\alpha,{\rm{bw}}}] is plotted against the angular divergence [\sigma_{\rm{e}}^{\,\prime}] for several values of the acceptance angle α and several values of the energy bandwidth bw.

[Figure 10]
Figure 10
The ratio [R_{\alpha,{\rm{bw}}}] = [F_{\alpha,{\rm{bw}}}/F_0] plotted as a function of [\sigma_{\rm{e}}^{\,\prime}] (assuming [\sigma_{\rm{e}}] = 0) for [\alpha] = 1 mrad (open circles) and 2 mrad (black bullets) and for (a) bw = 0.1%, (b) bw = 0.7% and (c) bw = 3%. Lines show the results of equation (13)[link] while points illustrate CAIN simulations (the error bars are too small to be visible).

4.2. Dependence on the electron beam energy spread

We now assume [\sigma_{\rm{e}}^{\,\prime}] = 0 mrad and we focus on the dependence of [R_{\alpha,{\rm{bw}}}] on the energy spread of the electrons. We start again with the ratio [R_{\rm{bw}}] of the X-ray flux within a given energy bandwidth bw [see equation (8)[link]] and F0, the X-ray flux in the `ideal' case where [\sigma_{\rm{e}}] = [\sigma_{\rm{e}}^{\,\prime}] = 0 [see equation (3)[link]]. We assume that an angular selection α is applied. Only the first-order term in [\delta\gamma] will be kept (where [\delta\gamma] is defined in §3.2[link]). The quantities [2|\delta\gamma|] and [\gamma_0^{\,2}\alpha^2] are denoted by [\mu_{\gamma}] and [\mu_{\alpha}], respectively. Then, depending on the relative values of [\mu_{\alpha}] and bw, three configurations have to be distinguished: (i) bw < [\mu_{\alpha}] < 2bw, (ii) 2bw < [\mu_{\alpha}] and (iii) [\mu_{\alpha}] < bw. Case (i) is illustrated in Fig. 11[link]. In this figure, the energy bandwidth is assumed to be centered at [\varepsilon_0] = 1. The left panels correspond to [\delta\gamma] negative, the right ones to [\delta\gamma] positive. Hatched areas represent X-rays whose energy is within the selected energy bandwidth but which are rejected by the angular cut. As shown in this figure, the size of the overlap between the energy acceptance domain and the angular acceptance one depends on bw, α and [\delta\gamma] in the following way:

[Figure 11]
Figure 11
X-ray energy spectrum. The fraction of this spectrum which falls within the energy acceptance band is shown in green. The hatched bands represent X-rays which are eliminated because of the angular cutoff. Left panels correspond to [\delta\gamma] negative, right ones to [\delta\gamma] positive. (a) and (b) illustrate the cases where 0 < [\mu_{\gamma}] < [\mu_{\alpha}-{\rm{bw}}], (c) and (d) where [\mu_{\alpha}-{\rm{bw}}] < [\mu_{\gamma}] < bw, and (e) and (f) where bw < [\mu_{\gamma}] < bw + [\mu_{\alpha}].

(i) For increasing values of [|\delta\gamma|], as long as [\mu_{\gamma}] + bw < [\mu_{\alpha}], all X-rays accepted in the [F_{\rm{bw}}] energy band are emitted within the angular acceptance α as shown in panels (a) and (b).

(ii) For [\mu_{\alpha}] − bw < [\mu_{\gamma}] < bw, the fraction of X-rays remaining in the [\mu_{\alpha}] angular acceptance band is (bw − [\mu_{\gamma}] + [\mu_{\alpha})/(2{\rm{bw}})] [see panels (c) and (d)].

(iii) The same fraction of X-rays pass the angular cut when bw < [\mu_{\gamma}] < bw + [\mu_{\alpha}] [panels (e) and (f)].

Similarly, the fraction of X-rays remaining in the [\mu_{\alpha}] band in the configuration (ii) defined above is 1 and [({\rm{bw}}-\mu_{\gamma}+\mu_{\alpha})/(2{\rm{bw}})] for [\mu_{\gamma}] < [\mu_{\alpha}] − bw and [\mu_{\alpha}] − bw < [\mu_{\gamma}] < bw + [\mu_{\alpha}], respectively. In the (iii) configuration, this fraction is [\mu_{\alpha}/{\rm{bw}}] for [\mu_{\gamma}] < bw − [\mu_{\alpha}] and (bw − [\mu_{\gamma}] + [\mu_{\alpha})/(2{\rm{bw}})] for bw − [\mu_{\alpha}] < [\mu_{\gamma}] < bw + [\mu_{\alpha}].

From this and from equation (5a)[link] which implies the following probability distribution,

[{{{\rm{d}}p}\over{{\rm{d}}(\mu_{\gamma})}} = {{1}\over{\sigma_{\rm{e}}\sqrt{2\pi}}} \, \exp\left(-\mu_{\gamma}^2/8\sigma_{\rm{e}}^{\,2}\right),]

the following expression of [R_{\alpha,{\rm{bw}}}] as a function of [\sigma_{\rm{e}}] can be derived:

[\eqalignno{ & R_{\alpha,{\rm{bw}}}(\sigma_{\rm{e}})_{\vphantom{\big|}} = \cr& \,\,\,R_{\rm{bw}} \left[\int\limits_0^{|\mu_{\alpha}-{\rm{bw}}|} {{1}\over{\sigma_{\rm{e}}\sqrt{2\pi}}} \, \exp\left(-\mu_{\gamma}^2/8\sigma_{\rm{e}}^{\,2}\right) \, {{ \min(\mu_{\alpha},{\rm{bw}}) }\over{ {\rm{bw}} }} \,{\rm{d}}(\mu_{\gamma}) \right. \cr& \,\,\,\left. +\int\limits_{|\mu_{\alpha}-{\rm{bw}}|}^{\mu_{\alpha}+{\rm{bw}}} {{1}\over{\sigma_{\rm{e}}\sqrt{2\pi}}} \, \exp\left(-\mu_{\gamma}^2/8\sigma_{\rm{e}}^{\,2}\right) \, {{ \mu_{\alpha}+{\rm{bw}}-\mu_{\gamma} }\over{ 2 {\rm{bw}} }} \,{\rm{d}}(\mu_{\gamma}) \right], &(14)}]

where [R_{\rm{bw}}] is given by equation (8)[link]. In the calculation of equation (14)[link], a simplification occurs which makes the dependence on min[(\mu_{\alpha},{\rm{bw}})] disappear, resulting in

[\eqalignno{ R_{\alpha,{\rm{{\rm{bw}}}}}(\sigma_{\rm{e}}) = {}& {{R_{\rm{bw}}} \over {2{\rm{bw}}}} \left( {{ 2\sqrt{2}\sigma_{\rm{e}} }\over{ \sqrt{\pi} }} \left\{ \exp\left[-{{(\mu_{\alpha}+{\rm{bw}})^2}\over{8\sigma_{\rm{e}}^2}}\right] \right.\right. \cr& \left. - \exp\left[-{{(\mu_{\alpha}-{\rm{bw}})^2}\over{8\sigma_{\rm{e}}^2}}\right] \right\} \cr& + (\mu_{\alpha}+{\rm{bw}}) \,{\rm{ERF}}\left({{\mu_{\alpha}+{\rm{bw}}}\over{2\sqrt{2}\sigma_{\rm{e}}}}\right) \cr& \left. - (\mu_{\alpha}-{\rm{bw}}) \,{\rm{ERF}}\left({{\mu_{\alpha}-{\rm{bw}}}\over{2\sqrt{2}\sigma_{\rm{e}}}}\right) \right). &(15)}]

CAIN simulations and results from equation (15)[link] are displayed in Fig. 12[link] where [R_{\alpha,{\rm{bw}}}] is plotted as a function of [\sigma_{\rm{e}}] for several values of the angular and the energy acceptances.

[Figure 12]
Figure 12
The ratio [R_{\alpha,{\rm{bw}}}] = [F_{\alpha,{\rm{bw}}}/F_0] versus [\sigma_{\rm{e}}], for [\alpha] = 1 mrad (open circles) and 2 mrad (black bullets), and for (a) bw = 0.1%, (b) bw = 0.7% and (c) bw = 3%. [\sigma_{\rm{e}}^{\,\prime}] = 0 is assumed here. Lines are obtained by using the analytic expression (15)[link] while points are CAIN simulations (the error bars are too small to be visible).

4.3. Combined effect of the energy spread and the angular divergence of the electron beam

When both [\sigma_{\rm{e}}] and [\sigma_{\rm{e}}^{\,\prime}] must be taken into account to calculate the ratio [R_{\alpha,{\rm{bw}}}(\sigma_{\rm{e}},\sigma_{\rm{e}}^{\,\prime})], this analytic formulation encounters a severe limitation for the following reason. When the incoming photons are backscattered by electrons whose energies are distributed according to equation (5a)[link], one may try to compute the X-ray flux in a way similar to that carried out in §4.1[link]. The bw term which comes in the [\theta_{\rm{bw}}] definition would have to be replaced by bweffective = bw + [2\delta\gamma], where [\delta\gamma] is defined in §3.2[link]. This would lead to an equation similar to equation (12)[link] with the [\delta\gamma] parameter occurring in the integrants and in the integrals' limits. The integration over [\zeta_{\rm{e}}^2] could still be performed, but the next step, namely an integral over [\delta\gamma] that takes into account the [\delta\gamma] distribution [equation (5a)[link]], cannot be performed analytically. Nevertheless, very good approximations can be obtained by employing equations previously established in this paper.

Indeed, if the electron beam angular divergence plays a dominant role with respect to the energy spread in the calculation of the X-ray flux within some acceptance cuts, a good approximation of [R_{\alpha,{\rm{bw}}}(\sigma_{\rm{e}},\sigma_{\rm{e}}^{\,\prime})] is obtained by using equation (13)[link] and assuming [\theta_{\rm{bw}}] = [({{\rm{bw}}_{\rm{eff}}})^{1/2}\!/\gamma_0], where [{\rm{bw}}_{\rm{eff}}] is the quadratic sum of bw and the standard deviation of [\mu_{\gamma}]: bweff = [[{\rm{bw}}^2+(2\sigma_{\rm{e}})^2]^{1/2}]. On the contrary, when the electron beam energy spread is dominant with respect to the angular divergence, equation (15)[link] must be used wherein the quantity [\sqrt{2\sigma_{\rm{e}}}] is replaced by the quadratic sum of [\sqrt{2\gamma_0^2\sigma_{\rm{e}}^{\,\prime\,2}}] and [\sqrt{2\sigma_{\rm{e}}}], resulting in the substitution [2\sigma_{\rm{e}}] [\rightarrow] [2\sigma_{\rm{e}}] + [2\gamma_0^2\sigma_{\rm{e}}^{\,\prime\,2}].

CAIN simulations and results of the modified equations (13)[link] or (15)[link] are shown in Fig. 13[link] for several values of α and bw. The discontinuities seen in the lines mark the use of equations (15)[link] or (13)[link] with the appropriate adaptation described above. When [2\gamma_0^2\sigma_{\rm{e}}^{\,\prime\,2}] + bw < [[{(2\sigma_{\rm{e}})^2 + {\rm{bw}}^2}]^{1/2}], equation (15)[link] is used wherein [2\sigma_{\rm{e}}] is replaced by [2\sigma_{\rm{e}}] + [2\gamma_0^2\sigma_{\rm{e}}^{\,\prime\,2}]. Otherwise, equation (13)[link] is used where bw is replaced by [[{\rm{bw}}^2+(2\sigma_{\rm{e}})^2]^{1/2}].

[Figure 13]
Figure 13
The ratio [R_{\alpha,{\rm{bw}}}] = [F_{\alpha,{\rm{bw}}}/F_0] as a function of [\sigma_{\rm{e}}^{\,\prime}], for an electron energy spread [\sigma_{\rm{e}}] of 0.5% (circles) and 1.5% (squares) and for a selection angle [\alpha] of 1 mrad (open symbols) and 2 mrad (full symbols). (a) bw = 0.1%. (b) bw = 0.7%. (c) bw = 3%. Lines are the analytic calculations (see text), points are CAIN simulations (the error bars are too small to be visible).

5. Additional comments and conclusion

We have derived analytic expressions for the total, the spectral and the angular X-ray fluxes of a Compton source according to the Compton kinematics and to the characteristics of the incoming electron and laser beams. The two dominant parameters governing the quality of such a source are the electron beam energy spread and its divergence. We point out a limitation of the analytic calculation of the flux [R_{\alpha,{\rm{bw}}}] in a given energy bandwidth and a specified solid angle when both the energy spread and the divergence of the electron beam must be taken into account in the calculation. Nevertheless, the expressions of [R_{\alpha,{\rm{bw}}}] obtained when either the energy spread or the angular divergence of the electron beam is taken into account can still be used with an appropriate modification.

Analytic expressions (11)[link], (13)[link] and (15)[link] provide the photon yield [R_{\alpha}] within a given angular acceptance [equation (11)[link]] and the photon yield [R_{\alpha,{\rm{bw}}}] within both a given angular acceptance and a given energy bandwidth [equations (13)[link] or (15)[link]]. Table 2[link] summarizes the validity domains of these equations in terms of the selected Compton kinematic region and according to the energy spread and the angular divergence of the two incoming beams as long as they are within the parameters ranges of current electron and laser beams (see Table 1[link]).

Table 2
Validity of equations (11)[link], (13)[link] and (15)[link] in terms of the parameters

The typical maximum values for current electron and laser beams shown in Table 1[link] are recalled. Each value indicates the upper limit of the parameter for which the equation is valid.

Parameters [\gamma\alpha] (rad) bw [\gamma\sigma^{\,\prime}_{\rm{e}}] (rad) [\sin\theta_{\rm{c}} \,\sigma_{\rm{L}}^{\,\prime}] (mrad) [\sigma_{\rm{e}}] [\sigma_{\rm{L}}]
e and typical laser values     0.25 10 A few % A few ‰
 
Equation (11)[link] 0.4   0.18 10 A few % A few ‰
Equations (13)[link] and (15)[link] Any 0.35 0.25 1 2–3%
†For the particular cases where [\sin\theta_{\rm{c}}\,\sigma_{\rm{L}}^{\,\prime}] ≃ 1–10 mrad, [R_{\alpha,{\rm{bw}}}] is valid but is overestimated by ∼0.5–5% via the term [R_{\rm{bw}}] (see §3.2[link]).

Within the large spectrum of X-ray applications in bio­medical, cultural heritage or material science, specific X-ray beams are required depending on the analysis technique and the sample to be used. For instance, a narrow energy bandwidth is required in most of the X-ray diffraction experiments (Dik et al., 2008[Dik, J., Janssens, K., Van Der Snickt, G., van der Loeff, L., Rickers, K. & Cotte, M. (2008). Anal. Chem. 80, 6436-6442.]) whereas therapy techniques such as stereotactic radiation therapy (Jacquet & Suortti, 2015[Jacquet, M. & Suortti, P. (2015). Phys. Med. 31, 596-600.]) can relax the constraint on the spectral width. Thus, for an incoming electron beam with a given emittance, it may be beneficial to minimize its transverse size even at the expense of its divergence. This could be the case for instance in experiments where a large X-ray number is the main requirement, or when an as small as possible source size is needed to carry out some analysis technique (such as phase-contrast imaging techniques). On the other hand, the electron beam divergence may be minimized at the expense of the transverse size when a quasi-monochromatic X-ray beam is required.

Analytic formulas (11)[link], (13)[link] and (15)[link] allow one to compute easily the spectral and spatial properties of an X-ray source based on photon backscattering, and highlight the beam parameters that play a key role for a given application. They should help when studying the feasibility of a particular experiment envisaged at some X-ray source.

Footnotes

1In the particular case of a Compton machine design such that [\sin\theta_{\rm{c}} \sigma_{\rm{L}}^{\,\prime}] ≃ 1–10 mrad, the effect of the laser divergence leads to a reduction of [R_{\rm{bw}}] from a few per million to a few percent depending on [\theta_{\rm{c}}], [\sigma_{\rm{L}}^{\,\prime}] and bw. The largest effect occurs for [\theta_{\rm{c}}] = 90° and [\sigma_{\rm{L}}^{\,\prime}] ≃ 10 mrad and would result for instance in a decrease of R0.1% and R3% by about 5% and 1%, respectively.

2We do not give here the expression of this third-order analytic formula since it is relatively heavy and since equation (11)[link] reproduces sufficiently well the simulations in our parameter range.

Acknowledgements

The authors are warmly grateful to Professor Jacques Haissinski for the many discussions and for his careful reading of the paper.

References

First citationArnold, A. & Teichert, J. (2011). Phys. Rev. ST Accel. Beams 14, 024801.  CrossRef Google Scholar
First citationDik, J., Janssens, K., Van Der Snickt, G., van der Loeff, L., Rickers, K. & Cotte, M. (2008). Anal. Chem. 80, 6436–6442.  Web of Science CrossRef PubMed CAS Google Scholar
First citationDonnelly, T. D. & Grossman, C. (1998). Am. J. Phys. 66, 677.  CrossRef Google Scholar
First citationJacquet, M. (2014). Nucl. Instrum. Methods Phys. Res. B, 331, 1–5.  Web of Science CrossRef CAS Google Scholar
First citationJacquet, M. & Suortti, P. (2015). Phys. Med. 31, 596–600.  Web of Science CrossRef PubMed Google Scholar
First citationRao, T. & Dowell, D. H. (2013). An Engineering Guide to Photoinjectors, https://arxiv.org/ftp/arxiv/papers/1403/1403.7539.pdfGoogle Scholar
First citationSchmüser, P., Dohlus, M. & Rossbach, J. (2008). Ultraviolet and Soft X-ray Free-Electron Lasers. Berlin: Springer.  Google Scholar
First citationSuzuki, T. (1976). KEK Technical Report 76–3, July 1976. KEK, Tsukuba, Japan.  Google Scholar
First citationTelnov, V. (2000). Nucl. Instrum. Methods Phys. Res. A, 455, 80–89.  CrossRef CAS Google Scholar
First citationVariola, A., Haissinkski, J., Loulergue, A. & Zomer, F. (2014). ThomX Technical Design Report. LAL/RT 14–21. SOLEIL/SOU-RA-3629.  Google Scholar
First citationYokoya, K. (2003). User's Manual of CAIN, https://lcdev.kek.jp/~yokoya/CAIN/cain235/CainMan235.pdf.   Google Scholar

© 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