Expand this Topic clickable element to expand a topic
Skip to content
Optica Publishing Group

Research on the beam structures observed from X-ray optics in the far field

Open Access Open Access

Abstract

For advanced X-ray sources such as synchrotron radiation facilities and X-ray free electron lasers, a smooth, structure-free beam on the far-field plane is usually strongly desired. The formation of the fine structures in far-field images downstream from imperfect optics must be understood. Although numerous studies have discussed the impacts on focused beams, there are still few quantitative theories for the impacts on beams in the far field. This article is an advance on our previous work, which discussed the uniformity of the intensity distribution in the far field. Here, a new theoretical approach is presented. It not only eases the assumptions needed to relate the fine structures to the wavefront curvature, but it also provides a quantitative estimation of the impacts of optical errors. The theoretical result is also verified by X-ray experiments.

Published by Optica Publishing Group under the terms of the Creative Commons Attribution 4.0 License. Further distribution of this work must maintain attribution to the author(s) and the published article's title, journal citation, and DOI.

1. Introduction

Current developments in modern X-ray synchrotron radiation and free electron laser facilities enable a wide range of applications. X-ray beamlines in these facilities must deliver high-quality X-ray beams to their users. The effects of imperfect beamline optics have been widely studied both on and off the focal plane [111]. However, there is still a lack of a satisfactory physical explanation for the impacts on the far-field images. In this paper, “far-field” means far from the focal plane of the tested optics. For optics that have no focal plane, such as a plane mirror, “far-field” means far from the source.

For many X-ray imaging experiments, a clean and uniform beam of variable size in the far field is strongly desired. However, due to the limitations of the present manufacturing technology for modern X-ray optical elements, some fine structures always appear in such beams, especially after reflective optics [1218]. An understanding of the origin of these structures is a necessary step towards their mitigation. Past works have shown them to be closely related to the curvature (second derivative) of the wavefront [1927]. Recently, the authors of this work measured fine structures in X-ray beams reflected by flat mirrors [28] and confirmed the structures’ relationship to the curvature of the wavefront by using rigorous wave optics theory.

A new theoretical approach is presented in this work. First, with our new approach, the far-field beam structures can be quantitatively estimated from known optical imperfections. With such information, manufacturers of X-ray optics could be given clear specifications of permissible fabrication errors. Second, the theoretical approach used in this work extends to general X-ray optics, rather than only to reflective optics as in our previous work. Third, this approach can be applied to partially as well as fully coherent beams.

In this paper, we will show how our theory can be related to the transport of intensity equation (TIE), which is based on fundamental physical principles and is widely used in X-ray phase imaging [2933]. Thus, they have the same loose assumptions about the source. The local variation of the intensity in the far-field images is shown to be simply proportional to the local variation of the wavefront curvature, which can be known from in-situ measurement using the speckle-based technique [28,34,35] or other in-situ wavefront sensing techniques [3641]. It can also be calculated using the more routine ex-situ measurement [42]. The new theory has also been tested experimentally. Although some of the results have been shown in our previous paper [28], they were only used to verify the previous theory qualitatively. Here, using the new theory, we interpret them quantitatively.

2. Theory

2.1 Propagation of intensity of light

Our derivation starts from the conservation of energy for the electromagnetic field [43]. The conservation law tells us that the time-averaged Poynting vector of the field < P > in a homogeneous free space with no free charges or currents must obey the following equation:

$$\oint_S { < {\mathbf P} > \cdot d{\mathbf S}} = 0\,,$$
where S is the closed surface of a bounded volume. The above equation shows that in homogeneous free space, under the assumptions of no absorption, scattering, or any other dissipation, the net energy that goes into or out of the closed volume is equal to 0. Its differential form is:
$$\nabla \cdot < {\mathbf P} > = 0\,,$$
where “∇” is the nabla operator in 3D space. Equation (2) means that the divergence of < P > is 0 under the above assumptions. We use s to represent the propagation direction of the time-averaged Poynting vector and I(r) to represent the intensity of the light. The Poynting vector can be represented as [43]:
$$\mathrm{\ < }{\mathbf P}\mathrm{\ > =\ }I\textrm{(}{\mathbf r}\textrm{)}{\mathbf s}\,,$$
r represents the coordinates in 3D space. The light propagation direction is parallel to the normal vector of its wavefront, which is a surface with an arbitrary constant phase. If we use ϕ(r) to represent the phase of the light in 3D space, the light propagation direction s is:
$${\mathbf s} = \frac{1}{k}\nabla \phi ({\mathbf r})\,,$$
where k is the wave number of the electromagnetic field, k = 2π/λ, and λ is the wavelength. Under the paraxial assumption, the well-known transport of intensity equation (TIE) [2933] can be derived using the above equations (see AppendixA.1 for details):
$$- k\frac{{\partial I({\boldsymbol x},z)}}{{\partial z}} = {\nabla _{\boldsymbol x}} \cdot [I({\boldsymbol x},z){\nabla _{\boldsymbol x}}\phi ({\boldsymbol x},z = 0)]\,.$$

Whereas in X-ray imaging TIE is mainly used to reconstruct the phase information, in this work, the phase information is always assumed to be known. We focus on the propagation of the intensity. By substituting Eqs. (3) and (4) into Eq. (2) and then solving Eq. (2), we can represent the light intensity at a point (x, y, z) in the space as [43] (see Appendix A.2 for detailed derivation):

$$I(x,y,z) = I({x_0},{y_0},{z_0}){e^{ - \int_{{s_0}}^s {\frac{1}{k}} {\nabla ^2}\phi ({\mathbf r})ds}}\,.$$

The integration is along the light ray, which is parallel to the local normal vector of the wavefront. In a homogeneous medium, it is a straight line. The above equation provides us with an analytical expression for the propagation of the intensity of light. It indicates that the propagation of the light intensity is modulated by the second derivative of the wavefront. We use it to discuss the impacts of the imperfections of the optics on the far-field images. We only need to notice that the phase term ϕ(r) can be written as the addition of an ideal wavefront and a curvature error caused by the imperfections from the optics.

For the trivial case of a plane wave, Eq. (6) shows that the intensity distribution remains unchanged by propagation as expected. See Appendix A.3 for the details.

2.2 Application to an arbitrary wavefront

In a curved wavefront, the local propagation direction is no longer constant, but varies across the initial plane. For example, the local wavefront propagation directions on a perfect spherical wave all point away from the origin if the wave is divergent, or toward the focus if the wave is convergent.

It can be proved [43,44] that for any curved wavefront, the integral term in Eq. (6) can be represented as:

$$\begin{aligned}{l} \frac{1}{k}{\nabla ^2}\phi ({\mathbf r}) &= \nabla \cdot {\mathbf s} \\& = \frac{1}{{{R_1}(s)}} + \frac{1}{{{R_2}(s)}}\,, \end{aligned}$$
where R1 and R2 are the principal radii of curvature of the wavefront and are functions of arc length s. Thus, Eq. (6) can be written as:
$$I({\mathbf s}) = I({x_0},{y_0},{z_0}){e^{ - \int_{{s_0}}^s {\left( {\frac{1}{{{R_1}(s)}} + \frac{1}{{{R_2}(s)}}} \right)} ds}}\,.$$

If the wavefront is a plane, the integrand in Eq. (8) is zero and the integration is trivial. If the wavefront is cylindrical, only one of the terms of the integrand in Eq. (8) is nonzero. Otherwise, to evaluate the integral term in the above equation, one notes that along the propagation direction, the wavefronts at two close positions are parallel and therefore [44]:

$$\frac{{d{R_1}(s)}}{{ds}} = 1,\;\,\,\frac{{d{R_2}(s)}}{{ds}} = 1\,.$$

So, we have:

$$\begin{array}{l} \int_{{s_0}}^s {\left( {\frac{1}{{{R_1}(s)}} + \frac{1}{{{R_2}(s)}}} \right)} ds\\ = \int_{{s_0}}^s {\left( {\frac{{\frac{{d{R_1}(s)}}{{ds}}}}{{{R_1}(s)}} + \frac{{\frac{{d{R_2}(s)}}{{ds}}}}{{{R_2}(s)}}} \right)} ds\\ = \ln ({{R_1}(s){R_2}(s)} )- \ln ({{R_1}({s_0}){R_2}({s_0})} )\,. \end{array}$$

Using the above equation, Eq. (8) becomes [43]:

$$I({\mathbf s}) = I(x,y,z) = I({x_0},{y_0},{z_0})\frac{{{R_1}({s_0}){R_2}({s_0})}}{{{R_1}(s){R_2}(s)}}\,.$$

R1 and R2 are in the orthogonal directions.

Equation (11) obeys the energy conservation law. Figure 1 illustrates two small wavefront surface elements at two different positions along the propagation direction. The small surface elements are bounded by lines of curvature. A0B0C0D0 is at the position s0, and ABCD is at the position s. R1(s0) and R2(s0) are the principal radii of curvature of the line segments A0B0 and B0C0, respectively. We have

$${A_0}{B_0} = {R_1}({s_0})d\theta ,\;\;\;\;{B_0}{C_0} = {R_2}({s_0})d\psi \,,$$
where and are the small angles corresponding to A0B0 and B0C0. OH and OV are the centers of these small arcs. The areas of these small surface elements are:
$$d{S_0} = {A_0}{B_0} \cdot {B_0}{C_0} = {R_1}({s_0}){R_2}({s_0})d\theta d\psi \,.$$

 figure: Fig. 1.

Fig. 1. A0B0C0D0 and ABCD are two small wavefront surface elements at two different positions along the propagation direction. The mesh represents the orthogonal curvilinear coordinate system. A0B0, B0C0, AB and BC are lines of curvature.

Download Full Size | PDF

Similarly, we have:

$$dS = AB \cdot BC = {R_1}(s){R_2}(s)d\theta d\psi \,.$$

The energy conservation law requires [43]:

$$I({s_0})d{S_0} = I(s)dS\,.$$

Combining Eq. (12)-(15), we get Eq. (11). This proves that Eq. (6) is also valid for any curved wavefront. Especially, for a perfect spherical wavefront, R1 = R2 = R and the Gaussian curvature is 1/R2. The energy conservation law for a perfect spherical wavefront is I(s0)R2(s0)=I(s)R2(s).

2.3 Application to imperfect X-ray optics

In a practical optical system, the optical elements are not always perfect. The wavefront at the exit plane of the system is usually not uniform. This is particularly true in X-ray optical systems due to the extremely small wavelength and small grazing angle of X-ray reflective optics. Thus, very high-quality optics are needed to deliver high-quality X-ray beams. However, at present, the available X-ray optical elements are far from perfect. As a result, it is very common to see fine structures in the far-field images.

We only consider the wavefront error in this work. The optical element is assumed not to absorb or scatter any of the incident radiation. Thus, only the phase differs in the planes immediately before and after the optical element, while the intensity is the same. We use I0(s0) to represent the ideal intensity distribution at the plane immediately after the optical element, and I’(s0) to represent the non-ideal intensity distribution at that plane. Here “ideal” means that the incident beam impinges on an ideal optic. Since the error of the optic is assumed to disturb only the phase of the beam, the intensity distribution for both ideal and non-ideal cases at the planes immediately after the optic is identical, I0(s0)=I’(s0). We apply Eq. (11) to the non-ideal wavefront, using R’1, 2(s0) to represent the local wavefront curvature at the plane immediately after the optics, R’1, 2(s) to represent the local wavefront curvature at the detector plane:

$${I^{\prime}}({\mathbf s}) = {I^{\prime}}(x,y,z) = {I_0}({x_0},{y_0},{z_0})\frac{{R_1^{\prime}({s_0})R_2^{\prime}({s_0})}}{{R_1^{\prime}(s)R_2^{\prime}(s)}}\,.$$

It is easy to find that the intensity distribution in the far field changes according to the local magnification factor R’1, 2(s0)/R’1, 2(s), also see Fig. 2. As a result, the contrast of the fine structures that appeared in the far-field image is proportional to the variations of the local magnification factor. Letting d = s-s0, the above equation can be written as:

$${I^{\prime}}({\mathbf s}) = {I_0}({s_0})\left[ {\frac{d}{{R_1^{\prime}(s)}} - 1} \right]\left[ {\frac{d}{{R_2^{\prime}(s)}} - 1} \right]\,,$$
where I0(s0) represents the incident beam. The above equation should be modified for other geometries different from that of Fig. 2. Note that we set the values of the radius of curvature to be always positive. Figure 2 only shows the case when the focus sits between the optical element and the detector. Appendix A.4 shows how the signs change in different geometries.

 figure: Fig. 2.

Fig. 2. A typical experimental layout for X-ray optics.

Download Full Size | PDF

Equation (17) relates the contrast of the fine structures in the intensity distribution to the local change of the wavefront curvature at the detector plane. Similar results can be also found in Peterzol et al [45].

We can also rewrite Eq. (16) in terms of the local wavefront’s radii of curvature at the plane immediately after the optical element:

$${I^{\prime}}({\mathbf s}) = {I_0}({s_0})\frac{{R_1^{\prime}({s_0})}}{{d - R_1^{\prime}({s_0})}}\frac{{R_2^{\prime}({s_0})}}{{d - R_2^{\prime}({s_0})}}\,.$$

Under the assumption that the perturbation of the phase is small, we use the first terms of the Taylor expansion of the fractions in the above equation:

$${I^{\prime}}({\mathbf s}) \approx I({\mathbf s})\left[ {1 - \frac{{\Delta \left( {\frac{1}{{{R_1}({s_0})}}} \right)}}{{\left( {\frac{1}{{{R_1}({s_0})}} - \frac{1}{d}} \right)}}} \right]\left[ {1 - \frac{{\Delta \left( {\frac{1}{{{R_2}({s_0})}}} \right)}}{{\left( {\frac{1}{{{R_2}({s_0})}} - \frac{1}{d}} \right)}}} \right]\,\,,$$
where I(s) represents the intensity distribution for the ideal wavefront at the detector plane. It can be obtained using Eq. (18) by replacing the disturbed (primed) radii of curvature with the ideal (unprimed) ones. Note that the Taylor expansion is made in terms of the local curvature rather than the local radius of curvature. Equation (16), (17) and (19) relate the local variations of the intensity distribution for a non-ideal wavefront to the variations of three interchangeable physical quantities, namely the local magnification factor, the wavefront local curvature at the detector plane and the plane immediately after the optical element, respectively.

Unlike our previous work [28], the theory proposed in this work imposes no assumptions on the type of X-ray optical elements. It can be applied to refractive (such as CRL), reflective (such as mirror) as well as diffractive (such as Fresnel zone plate) optics, as long as the relationship between the error from the optical element and the wavefront has been set. For instance, because of the very small grazing angle of the X-ray mirror, the wavefront coordinates should be projected back to the mirror surface through division by the small grazing angle, usually several milliradians. For strongly focusing optics in the far field, such as the CRL stack, multilayer Laue lens or focusing mirror, for which dR1,2(s0) according to Eq. (19), the relative change of the wavefront curvature determines the contrast of the fine structures appearing in the far-field images. If the variation of the wavefront curvature is within 5%, then the expected intensity variation is also within 5%. For weakly or non-focusing optics, on the other hand, the contrast of the fine structures also depends on the distance from the optic to the detector as predicted by Eq. (17) or (19). We will give experimental evidence for both cases later in this paper. For 2D focusing optics such as most CRLs, one needs to note that the principal directions of the perturbed wavefront will be different from the principal directions of the individual components. This may add more difficulties in predicting the far-field intensity accurately. However, this does not arise for the 1D focusing optics used as examples in this paper.

It is easy to show that Eq. (6) can lead to the foundations of geometrical optics [43]. Under certain circumstances, the ray-optical approach is sufficient to explain the formation of the fine structures in the far-field intensity distribution. From our derivation of the TIE at the beginning of this section, the TIE can be applied to any type of light source, as long as the beam intensity and the phase are properly defined. For a partially coherent source, the phase can be defined statistically [33]. Peterzol et al. [45] have pointed out that with some restrictions on high-spatial-frequency features, the TIE can be derived from general diffraction theory. Our theory has the same assumptions as TIE. Thus, the same argument can also be applied to our approach. The change of the wavefront local curvature determines how much beam energy can be pulled out from the ideal intensity distribution. For a highly coherent beam, the energy removed from the ideal distribution can become concentrated in even finer structures in the far-field image. Equation (16), (17) and (19) still provide a good way to estimate the uniformity of the intensity distribution in the far-field image.

3. Experiments

X-ray mirrors are simple yet still suitable optical elements for the verification of the theory proposed in the previous section because the small grazing angle makes the effects of surface errors much less in the sagittal direction than in the tangential one. The fine structures that appear on the observation plane thus come mainly from the surface error along the mirror length, effectively reducing the problem from two dimensions to just one. The relationship between the local wavefront curvature and the far-field intensity distribution can be easily observed. The features would be much more difficult to match for 2D optics since contributions from both orthogonal directions will be added to the intensity distribution. No one-to-one correspondence can be easily found.

Two mirrors have been measured to verify the theory. The experiments using X-rays were conducted at the Diamond Light Source B16 Test beamline [46]. Two types of monochromatic beam with different energy bandwidths can be provided by either a Ru/B4C double multilayer monochromator (DMM) or a Si (111) double crystal monochromator (DCM). Except for the monochromators, there are only windows between the tested mirror and the source. Thus, the ideal intensity image in the far field will have a relatively large area with gradual spatial variations in the intensity. The measurements using visible light were conducted at the Optical Metrology Laboratory [42], which is also at the Diamond Light Source.

3.1 Measurement of a plane mirror

A plane mirror with no coating was measured using the at-wavelength wavefront sensing technique. The rms slope error of this plane mirror is about 0.17 µrad. The slope error and the local wavefront curvature of this mirror were obtained using the speckle-based in-situ at-wavelength techniques [28,34,35]. A diffuser is used to generate the speckle pattern for X-rays. By comparing the local shifts of the speckle pattern with the reference pattern, the local wavefront curvature or slope can be reconstructed depending on the choice of references. The X-ray beam came from the double crystal monochromator. Except for the monochromator, there are only windows between the tested mirror and the source. Table 1 shows the basic experimental parameters for this mirror.

Tables Icon

Table 1. Experimental parameters for the plane mirror

Although this mirror has been shown in our previous work [28], there we only show the correspondence between the wavefront curvature and the fine structures in the far-field images qualitatively. Here, using the theory proposed in this paper, we can discuss the contrast of the fine structures quantitatively.

Figure 3(a) shows the reflected intensity image on the detector plane. Figure 3(b) shows the measured wavefront local curvature on the same plane. Firstly, although this mirror is quite smooth in terms of its slope error, those fine structures can still be seen from the intensity image. If we have a closer look at Fig. 3(a) and (b), we can find that fine structures shown in (b) can find their matches in (a). Secondly, some other features that appear on the intensity image are not caused by the wavefront curvature error in the measured direction, which is vertical in this case. Detailed inspection reveals that they are mainly from the incident beam. They are formed by other factors, such as the wavefront curvature error in the orthogonal direction or the absorption. Thirdly, for a plane mirror, we can use Eq. (17) to estimate the contrast of the fine structures that appeared in the far-field intensity image. Figure 3(c) shows a strip of the extracted and averaged intensity data and the intensity distribution estimated using Eq. (17) from the corresponding strip of wavefront curvature data. Observing the two curves, we can see that all the features match well with each other. It is known that the result from the speckle-based technique is smeared a little bit by the convolution with several detector pixels due to the choice of finite window size. That can explain the discrepancies between the intensity curve and the theoretical results, especially on the sharp jumps. We define the contrast of the fine structures here as σI/I, where σI is the standard deviation of the beam intensity. According to Eq. (17), the standard deviation of the wavefront curvature at the detector plane multiplied by d can also represent the contrast of the fine structures, where d is the distance between the mirror center and the detector plane (1.705 m). From Fig. 3(c), according to our definition, the standard deviation of the extracted intensity is around 138. The mean value is around 3300. Thus, σI/I ≈ 0.042. The standard deviation of the extracted wavefront curvature is 0.0278 m-1. Multiplying it by the distance d also represents the contrast of the fine structures on the far-field image. The value is around 0.0474, which is close to σI/I. These results verify Eq. (17).

 figure: Fig. 3.

Fig. 3. (a) is the far-field intensity image after the reflective mirror. (b) is the wavefront curvature on the detector plane. A strip of intensity data is extracted, averaged, and showed in (c) in a blue curve. The average is along the horizontal axis. The corresponding strip of wavefront curvature data is also extracted and averaged. It is then used for Eq. (17). The result is shown in a red curve in (c).

Download Full Size | PDF

3.2 Measurement of an elliptical mirror

An elliptical mirror with multiple coated stripes was also measured. These stripes are designed to have parabolic sections with different amplitudes to modulate the exit beam [47]. The stripe measured here has the highest modulated amplitude. The parameters of the mirror are shown in Table 2. The slope of the parabolic modulation was measured offline using the Diamond-NOM [48] and is shown in Fig. 4(a).

 figure: Fig. 4.

Fig. 4. (a) Mirror slope modulation measured using Diamond-NOM. (b) The intensity distribution in the far field. The mean intensity curve (averaged along the vertical axis) within the red rectangular box is extracted and shown in the red curve in (c). The blue curve in (c) is the estimated intensity distribution from the wavefront curvature error using Eq. (19). The horizontal coordinate of the blue curve is along the mirror surface.

Download Full Size | PDF

Tables Icon

Table 2. Main parameters of the elliptical mirror

Since the distance from the mirror centre to the detector is much larger than to the mirror focus, Eq. (19) is appropriate for the estimation of the far-field intensity variation. To do that, we need to convert the mirror surface slope modulation to the wavefront curvature modulation, which is the derivative of the wavefront slope modulation. Note that the wavefront slope modulation is twice the mirror surface slope modulation. Considering the elliptical shape of the mirror surface, we cannot just project the mirror surface coordinates to the wavefront coordinates by simply multiplying the angular factor of sin(θ) [49]. To be accurate, one needs to do ray tracing. We use the xrt python package [50] for our calculation. The converted wavefront curvature can be used to estimate the intensity distribution. This is shown in Fig. 4(c) with a blue curve. This curve is calculated using Eq, (19).

The measured far-field X-ray intensity distribution after this elliptical mirror is shown in Fig. 4(b). The horizontal stripes shown in Fig. 4(b) are introduced by the double multilayer monochromator. The mirror was placed facing sideways to decouple the impact of the incident beam. The intensity data within a red rectangular box is extracted and averaged. This is shown in Fig. 4(c) with a red curve. The two curves in Fig. 4(c) show clearly the correlation between the wavefront curvature modulation and the intensity distribution. They both have step-like jumps. For the measured intensity distribution, we define the contrast as (Imax-Imin)/(Imax + Imin). The contrast of the first jump (at around pixel 100) is (3650-2750)/(3650 + 2750)≈0.14, the contrast of the second jump (at around pixel 390) is (3850-3000)/(3850 + 3000) ≈0.124, and the contrast from the lowest step to the highest is (3850-2750)/(3850 + 2750)≈0.17. The corresponding contrast of the estimated intensity distribution using Eq. (19) is around 0.24 for the first jump, 0.19 for the second jump and 0.27 from the lowest step to the highest. These values are compatible with those from the measured intensity distribution. The discrepancies come from several reasons. First, the mirror pitch angle may not be at the exact nominal one due to the error coming from the installation. Second, at the position with a large jump, our theory may not be effective because the nonzero window size chosen for the speckle technique creates a convolution effect on the data processing procedure. This can further impact the correspondence between the theory and the experiment. In any case, we can still use the theory proposed in Section 2 to estimate the variation of the intensity image in the far field.

4. Conclusions

In this paper, we have confirmed with theoretical derivations the correlation between the wavefront curvature and the fine structures that appear in far-field images produced by imperfect X-ray optics. We have also provided equations that allow quantitative estimation of the contrast of the fine structures if the imperfections of the optic are known. Our new theoretical approach has a wider range of applications than our previous work. It can be used for general X-ray optics and for all sources, including partially coherent ones, whose radiation has a defined intensity and phase. The X-ray intensity variations in the far field after a plane mirror and an elliptical mirror have been investigated experimentally. The measured contrast of the fine structures is in good agreement with the theory proposed here. Thus, the proposed theoretical approach can help us to set manufacturing tolerances on the optical imperfections when smooth, structure-free beams in the far field are required.

Appendix

A.1 Derivation of the TIE

We derive the TIE from Eq. (2). Replacing the wavefront propagation direction s in Eq. (3) with Eq. (4), we obtain:

$$< {\mathbf P} > = I({\mathbf r})\frac{1}{k}\nabla \phi ({\mathbf r})\,.$$

We now write r as (x, z). x = (x,y) is a 2D vector of transverse coordinates that is perpendicular to the longitudinal direction z. The operator ∇ can be decomposed as:

$$\nabla = {\nabla _x} + \frac{\partial }{{\partial z}}{{\mathbf e}_z}\,,$$
where
$${\nabla _x} = \frac{\partial }{{\partial x}}{{\mathbf e}_x} + \frac{\partial }{{\partial y}}{{\mathbf e}_y}\,.$$
ex, eyare the unit vectors along the x and y direction, respectively. Under the paraxial assumption, the gradient of the phase term can be represented as:
$$\nabla \phi ({\mathbf r}) = {\nabla _{\boldsymbol x}}\phi ({\boldsymbol x},z = 0) + k{{\mathbf e}_z}\,,$$
where ezis the unit vector along the longitudinal direction z. Substituting Eq. (20) and Eq. (23) into Eq. (2), we have:
$$- k\frac{{\partial I({\boldsymbol x},z)}}{{\partial z}} = {\nabla _{\boldsymbol x}} \cdot [I({\boldsymbol x},z){\nabla _{\boldsymbol x}}\phi ({\boldsymbol x},z = 0)]\,.$$

Equation (24) is Eq. (5) in the main text. It is the well-known transport of intensity equation (TIE).

A.2 Derivation of Eq. (6) [43]

To do this, we introduce the derivative along the light path:

$$\frac{d}{{ds}} = {\mathbf s} \cdot \nabla \,,$$
where s is the length of the arc along the ray. The propagation direction of the light s is defined in Eq. (4). We have |ds/ds|=1.

Substituting for < P > from Eq. (20) into Eq. (2), and using the identities ∇·uv = u(∇·v)+v·∇u and ∇·∇=∇2, Eq. (2) yields:

$$\frac{1}{k}I({\mathbf r}){\nabla ^2}\phi ({\mathbf r}) + \frac{1}{k}\nabla \phi ({\mathbf r}) \cdot \nabla I({\mathbf r}) = 0\,.$$

The above equation can also be written as:

$$\frac{1}{k}{\nabla ^2}\phi ({\mathbf r}) + \frac{1}{k}\nabla \phi ({\mathbf r}) \cdot \nabla \ln I({\mathbf r}) = 0\,.$$

Using Eq. (4) and the directional derivative defined in Eq. (25), the above equation can be further expressed as:

$$\frac{d}{{ds}}\ln I({\mathbf s}) ={-} \frac{1}{k}{\nabla ^2}\phi ({\mathbf r})\,.$$

The solution of Eq. (28) is:

$$I({\mathbf s}) = I({x_0},{y_0},{z_0}){e^{ - \int_{{s_0}}^s {\frac{1}{k}} {\nabla ^2}\phi ({\mathbf r})ds}}\,.$$

Equation (29) provides us with an analytical expression for the propagation of the intensity of light. It is Eq. (6) in the main text.

A.3 Application of Eq. (6) to a plane wave

Equation (6) or (29) indicates that the propagation of the light intensity is modulated by the second derivative of the wavefront. For the trivial case of a plane wave, the phase term ϕ(r) is:

$$\phi ({\mathbf r}) = {\mathbf k} \cdot {\mathbf r} = {k_x}x + {k_y}y + {k_z}z\,.$$

For any point (x0, y0, z0) in the initial plane, the propagation direction of the light is a constant unit vector, s = (kx/k, ky/k, kz/k). As a result, the integration path for Eq. (6) or (29) is a straight line along the direction of s, and the equation becomes:

$$I({\mathbf s}) = I[{x(s),y(s),\,z(s)} ]= I({x_0},{y_0},{z_0})\,,$$
thus showing that as expected the intensity distribution remains unchanged by propagation.

A.4 Eq. (17), (18), (19) in different geometries

Figure 2 shows one case when the focus sits between the optical element and the detector. There are other cases. Figure 5 shows the corresponding sign changes for Eq. (17), (18) and (19) in different geometries. The radii of curvature in these cases are always assumed to be positive.

 figure: Fig. 5.

Fig. 5. Different forms of Eq. (17), (18) and (19) in different geometries. s0 represents the position of the tested optic, s represents the detector. d always represents the distance between the optic and the detector. The radii of curvature in the diagram are always positive.

Download Full Size | PDF

Funding

Diamond Light Source (beamtime NR26501 , NT31201).

Acknowledgments

We thank Dr Oliver Fox, Dr Vishal Dhamgaye and Andrew Malandain for their support on the X-ray experiments. We also thank Dr Simon Alcock and Dr Ioana Nistea for providing the visible light measurement results.

Disclosures

The authors declare no conflicts of interest.

Data availability

Data underlying the results presented in this paper are not publicly available at this time but may be obtained from the authors upon reasonable request.

References

1. E. L. Church and P. Z. Takacs, “Specification of surface figure and finish in terms of system performance,” Appl. Opt. 32(19), 3344–3353 (1993). [CrossRef]  

2. J. E. Harvey, K. L. Lewotsky, and A. Kotha, “Effects of surface scatter on the optical performance of x-ray synchrotron beam-line mirrors,” Appl. Opt. 34(16), 3024–3032 (1995). [CrossRef]  

3. D. Cocco, “Recent developments in UV optics for ultra-short, ultra-intense coherent light sources,” Photonics 2(1), 40–49 (2015). [CrossRef]  

4. T. Pardini, D. Cocco, and S. P. Hau-Riege, “Effect of slope errors on the performance of mirrors for x-ray free electron laser applications,” Opt. Express 23(25), 31889–31895 (2015). [CrossRef]  

5. L. Raimondi and D. Spiga, “Mirrors for X-ray telescopes: Fresnel diffraction-based computation of point spread functions from metrology,” Astron. Astrophys. 573, A22 (2015). [CrossRef]  

6. V. V. Yashchuk, L. V. Samoylova, and I. V. Kozhevnikov, “Specification of x-ray mirrors in terms of system performance: new twist to an old plot,” Opt. Eng. 54(2), 025108 (2015). [CrossRef]  

7. F. Seiboth, A. Schropp, M. Scholz, F. Wittwer, C. Rödel, M. Wünsche, T. Ullsperger, S. Nolte, J. Rahomäki, and K. Parfeniukas, “Perfect X-ray focusing via fitting corrective glasses to aberrated optics,” Nat. Commun. 8(1), 14623 (2017). [CrossRef]  

8. S. Matsuyama, T. Inoue, J. Yamada, J. Kim, H. Yumoto, Y. Inubushi, T. Osaka, I. Inoue, T. Koyama, and K. Tono, “Nanofocusing of X-ray free-electron laser using wavefront-corrected multilayer focusing mirrors,” Sci. Rep. 8(1), 17440 (2018). [CrossRef]  

9. R. Celestre, S. Berujon, T. Roth, M. Sanchez del Rio, and R. Barrett, “Modelling phase imperfections in compound refractive lenses,” J. Synchrotron Radiat. 27(2), 305–318 (2020). [CrossRef]  

10. L. Hu, J. P. Sutter, and H. Wang, “Fast convolution-based performance estimation method for diffraction-limited source with imperfect X-ray optics,” J. Synchrotron Radiat. 27(6), 1539–1552 (2020). [CrossRef]  

11. F. Seiboth, A. Kubec, A. Schropp, S. Niese, P. Gawlitza, J. Garrevoet, V. Galbierz, S. Achilles, S. Patjens, and M. E. Stuckelberger, “Rapid aberration correction for diffractive X-ray optics by additive manufacturing,” Opt. Express 30(18), 31519–31529 (2022). [CrossRef]  

12. A. Rack, T. Weitkamp, M. Riotte, D. Grigoriev, T. Rack, L. Helfen, T. Baumbach, R. Dietsch, T. Holz, and M. Krämer, “Comparative study of multilayers used in monochromators for synchrotron-based coherent hard X-ray imaging,” J. Synchrotron Rad. 17(4), 496–510 (2010). [CrossRef]  

13. C. Morawe, R. Barrett, K. Friedrich, R. Klünder, and A. Vivo, “Spatial coherence studies on x-ray multilayers,” Proc. SPIE 8139, 813909 (2011). [CrossRef]  

14. A. Rack, C. Morawe, L. Mancini, D. Dreossi, D. Parkinson, A. MacDowell, F. Siewert, T. Rack, T. Holz, and M. Krämer, “Reflection on multilayer mirrors: beam profile and coherence properties,” Proc. SPIE 9207, 92070V (2014). [CrossRef]  

15. C. Morawe, R. Barrett, P. Cloetens, B. Lantelme, J.-C. Peffen, and A. Vivo, “Graded multilayers for figured Kirkpatrick-Baez mirrors on the new ESRF end station ID16A,” Proc. SPIE 9588, 958803 (2015). [CrossRef]  

16. S. Matsuyama, S. Yasuda, J. Yamada, H. Okada, Y. Kohmura, M. Yabashi, T. Ishikawa, and K. Yamauchi, “50-nm-resolution full-field X-ray microscope without chromatic aberration using total-reflection imaging mirrors,” Sci. Rep. 7(1), 46358 (2017). [CrossRef]  

17. M. N. Boone, F. Van Assche, S. Vanheule, S. Cipiccia, H. Wang, L. Vincze, and L. Van Hoorebeke, “Full-field spectroscopic measurement of the X-ray beam from a multilayer monochromator using a hyperspectral X-ray camera,” J. Synchrotron Radiat. 27(1), 110–118 (2020). [CrossRef]  

18. T. Koyama, Y. Senba, H. Yamazaki, T. Takeuchi, M. Tanaka, Y. Shimizu, K. Tsubota, Y. Matsuzaki, H. Kishimoto, and T. Miura, “Double-multilayer monochromators for high-energy and large-field X-ray imaging applications with intense pink beams at SPring-8 BL20B2,” J. Synchrotron Radiat. 29(5), 1265–1272 (2022). [CrossRef]  

19. F. Roddier, “Curvature sensing and compensation: a new concept in adaptive optics,” Appl. Opt. 27(7), 1223–1225 (1988). [CrossRef]  

20. I. A. Schelokov, O. Hignette, C. Raven, A. A. Snigirev, I. Snigireva, and A. Suvorov, “X-ray interferometry technique for mirror and multilayer characterization,” Proc. SPIE 2805, 282–292 (1996). [CrossRef]  

21. A. Rommeveaux and A. Souvorov, “Flat x-ray mirrors as optical elements for coherent synchrotron radiation conditioning,” Proc. SPIE 3773, 70–77 (1999). [CrossRef]  

22. A. Souvorov, M. Yabashi, K. Tamasaku, T. Ishikawa, Y. Mori, K. Yamauchi, K. Yamamura, and A. Saito, “Deterministic retrieval of surface waviness by means of topography with coherent X-rays,” J. Synchrotron Radiat. 9(4), 223–228 (2002). [CrossRef]  

23. A. Suvorov, H. Ohashi, S. Goto, K. Yamauchi, and T. Ishikawa, “One-dimensional surface profile retrieval from grazing incidence images under coherent X-ray illumination,” Nucl. Instrum. Methods Phys. Res., Sect. A 616(2-3), 277–280 (2010). [CrossRef]  

24. J. Nicolas and G. García, “Modulation of intensity in defocused beams,” Proc. SPIE 8848, 884810 (2013). [CrossRef]  

25. D. Spiga, S. Basso, M. Bavdaz, V. Burwitz, M. Civitani, O. Citterio, M. Ghigo, G. Hartner, B. Menz, and G. Pareschi, “Profile reconstruction of grazing-incidence X-ray mirrors from intra-focal X-ray full imaging,” Proc. SPIE 8861, 88611F (2013). [CrossRef]  

26. D. Spiga, L. Raimondi, C. Svetina, and M. Zangrando, “X-ray beam-shaping via deformable mirrors: Analytical computation of the required mirror profile,” Nucl. Instrum. Methods Phys. Res., Sect. A 710, 125–130 (2013). [CrossRef]  

27. J. P. Sutter, S. G. Alcock, F. Rust, H. Wang, and K. Sawhney, “Structure in defocused beams of X-ray mirrors: causes and possible solutions,” Proc. SPIE 9208, 92080G (2014). [CrossRef]  

28. L. Hu, H. Wang, J. P. Sutter, and K. Sawhney, “Investigation of the stripe patterns from X-ray reflection optics,” Opt. Express 29(3), 4270–4286 (2021). [CrossRef]  

29. M. R. Teague, “Deterministic phase retrieval: a Green’s function solution,” J. Opt. Soc. Am. 73(11), 1434–1441 (1983). [CrossRef]  

30. T. Gureyev, A. Roberts, and K. Nugent, “Partially coherent fields, the transport-of-intensity equation, and phase uniqueness,” J. Opt. Soc. Am. A 12(9), 1942–1946 (1995). [CrossRef]  

31. T. E. Gureyev and K. A. Nugent, “Rapid quantitative phase imaging using the transport of intensity equation,” Opt. Commun. 133(1-6), 339–346 (1997). [CrossRef]  

32. D. Paganin, Coherent X-ray optics (Oxford University, 2006).

33. C. Zuo, J. Li, J. Sun, Y. Fan, J. Zhang, L. Lu, R. Zhang, B. Wang, L. Huang, and Q. Chen, “Transport of intensity equation: a tutorial,” Opt. Lasers. Eng. 135, 106187 (2020). [CrossRef]  

34. H. Wang, J. Sutter, and K. Sawhney, “Advanced in situ metrology for x-ray beam shaping with super precision,” Opt. Express 23(2), 1605–1614 (2015). [CrossRef]  

35. L. Hu, H. Wang, O. Fox, and K. Sawhney, “Two-dimensional speckle technique for slope error measurements of weakly focusing reflective X-ray optics,” J. Synchrotron Radiat. 29(6), 1385–1393 (2022). [CrossRef]  

36. O. Hignette, A. K. Freund, and E. Chinchio, “Incoherent X-ray mirror surface metrology,” Proc. SPIE 3152, 188–199 (1997). [CrossRef]  

37. H. Yumoto, H. Mimura, S. Matsuyama, S. Handa, Y. Sano, M. Yabashi, Y. Nishino, K. Tamasaku, T. Ishikawa, and K. Yamauchi, “At-wavelength figure metrology of hard x-ray focusing mirrors,” Rev. Sci. Instrum. 77(6), 063712 (2006). [CrossRef]  

38. M. Idir, P. Mercere, M. H. Modi, G. Dovillaire, X. Levecq, S. Bucourt, L. Escolano, and P. Sauvageot, “X-ray active mirror coupled with a Hartmann wavefront sensor,” Nucl. Instrum. Methods Phys. Res., Sect. A 616(2-3), 162–171 (2010). [CrossRef]  

39. C. M. Kewish, M. Guizar-Sicairos, C. Liu, J. Qian, B. Shi, C. Benson, A. M. Khounsary, J. Vila-Comamala, O. Bunk, and J. R. Fienup, “Reconstruction of an astigmatic hard X-ray beam and alignment of KB mirrors from ptychographic coherent diffraction data,” Opt. Express 18(22), 23420–23427 (2010). [CrossRef]  

40. J. Sutter, S. Alcock, and K. Sawhney, “In situ beamline analysis and correction of active optics,” J. Synchrotron Radiat. 19(6), 960–968 (2012). [CrossRef]  

41. L. Assoufid, X. Shi, S. Marathe, E. Benda, M. J. Wojcik, K. Lang, R. Xu, W. Liu, A. T. Macrander, and J. Z. Tischler, “Development and implementation of a portable grating interferometer system as a standard tool for testing optics at the Advanced Photon Source beamline 1-BM,” Rev. Sci. Instrum. 87(5), 052004 (2016). [CrossRef]  

42. I. T. Nistea, S. Alcock, M. Bazan da Silva, and K. Sawhney, “The Optical Metrology Laboratory at Diamond: pushing the limits of nano-metrology,” Proc. SPIE 11109, 1110906 (2019). [CrossRef]  

43. M. Born and E. Wolf, Principles of Optics: 60th Anniversary Edition (Cambridge University, 2019).

44. M. Kline, “A note on the expansion coefficient of geometrical optics,” Comm. Pure Appl. Math. 14, 473 (1961). [CrossRef]  

45. A. Peterzol, A. Olivo, L. Rigon, S. Pani, and D. Dreossi, “The effects of the imaging system on the validity limits of the ray-optical approach to phase contrast imaging,” Med. Phys. 32(12), 3617–3627 (2005). [CrossRef]  

46. K. J. S. Sawhney, I. P. Dolbnya, M. K. Tiwari, L. Alianelli, S. M. Scott, G. M. Preece, U. K. Pedersen, and R. D. Walton, “A Test Beamline on Diamond Light Source,” AIP Conf. Proc. 1234, 387–390 (2010). [CrossRef]  

47. D. Laundy, K. Sawhney, I. Nistea, S. Alcock, I. Pape, J. Sutter, L. Alianelli, and G. Evans, “Development of a multi-lane X-ray mirror providing variable beam sizes,” Rev. Sci. Instrum. 87(5), 051802 (2016). [CrossRef]  

48. S. Alcock, K. Sawhney, S. Scott, U. Pedersen, R. Walton, F. Siewert, T. Zeschke, F. Senf, T. Noll, and H. Lammert, “The Diamond-NOM: A non-contact profiler capable of characterizing optical figure error with sub-nanometre repeatability,” Nucl. Instrum. Methods Phys. Res., Sect. A 616(2-3), 224–228 (2010). [CrossRef]  

49. X. Shi, R. Reininger, M. Sanchez del Rio, and L. Assoufid, “A hybrid method for X-ray optics simulation: combining geometric ray-tracing and wavefront propagation,” J. Synchrotron Radiat. 21(4), 669–678 (2014). [CrossRef]  

50. K. Klementiev and R. Chernikov, “Powerful scriptable ray tracing package xrt,” Proc. SPIE 9209, 92090A (2014). [CrossRef]  

Data availability

Data underlying the results presented in this paper are not publicly available at this time but may be obtained from the authors upon reasonable request.

Cited By

Optica participates in Crossref's Cited-By Linking service. Citing articles from Optica Publishing Group journals and other participating publishers are listed here.

Alert me when this article is cited.


Figures (5)

Fig. 1.
Fig. 1. A0B0C0D0 and ABCD are two small wavefront surface elements at two different positions along the propagation direction. The mesh represents the orthogonal curvilinear coordinate system. A0B0, B0C0, AB and BC are lines of curvature.
Fig. 2.
Fig. 2. A typical experimental layout for X-ray optics.
Fig. 3.
Fig. 3. (a) is the far-field intensity image after the reflective mirror. (b) is the wavefront curvature on the detector plane. A strip of intensity data is extracted, averaged, and showed in (c) in a blue curve. The average is along the horizontal axis. The corresponding strip of wavefront curvature data is also extracted and averaged. It is then used for Eq. (17). The result is shown in a red curve in (c).
Fig. 4.
Fig. 4. (a) Mirror slope modulation measured using Diamond-NOM. (b) The intensity distribution in the far field. The mean intensity curve (averaged along the vertical axis) within the red rectangular box is extracted and shown in the red curve in (c). The blue curve in (c) is the estimated intensity distribution from the wavefront curvature error using Eq. (19). The horizontal coordinate of the blue curve is along the mirror surface.
Fig. 5.
Fig. 5. Different forms of Eq. (17), (18) and (19) in different geometries. s0 represents the position of the tested optic, s represents the detector. d always represents the distance between the optic and the detector. The radii of curvature in the diagram are always positive.

Tables (2)

Tables Icon

Table 1. Experimental parameters for the plane mirror

Tables Icon

Table 2. Main parameters of the elliptical mirror

Equations (31)

Equations on this page are rendered with MathJax. Learn more.

S < P > d S = 0 ,
< P >= 0 ,
  < P   >=   I ( r ) s ,
s = 1 k ϕ ( r ) ,
k I ( x , z ) z = x [ I ( x , z ) x ϕ ( x , z = 0 ) ] .
I ( x , y , z ) = I ( x 0 , y 0 , z 0 ) e s 0 s 1 k 2 ϕ ( r ) d s .
l 1 k 2 ϕ ( r ) = s = 1 R 1 ( s ) + 1 R 2 ( s ) ,
I ( s ) = I ( x 0 , y 0 , z 0 ) e s 0 s ( 1 R 1 ( s ) + 1 R 2 ( s ) ) d s .
d R 1 ( s ) d s = 1 , d R 2 ( s ) d s = 1 .
s 0 s ( 1 R 1 ( s ) + 1 R 2 ( s ) ) d s = s 0 s ( d R 1 ( s ) d s R 1 ( s ) + d R 2 ( s ) d s R 2 ( s ) ) d s = ln ( R 1 ( s ) R 2 ( s ) ) ln ( R 1 ( s 0 ) R 2 ( s 0 ) ) .
I ( s ) = I ( x , y , z ) = I ( x 0 , y 0 , z 0 ) R 1 ( s 0 ) R 2 ( s 0 ) R 1 ( s ) R 2 ( s ) .
A 0 B 0 = R 1 ( s 0 ) d θ , B 0 C 0 = R 2 ( s 0 ) d ψ ,
d S 0 = A 0 B 0 B 0 C 0 = R 1 ( s 0 ) R 2 ( s 0 ) d θ d ψ .
d S = A B B C = R 1 ( s ) R 2 ( s ) d θ d ψ .
I ( s 0 ) d S 0 = I ( s ) d S .
I ( s ) = I ( x , y , z ) = I 0 ( x 0 , y 0 , z 0 ) R 1 ( s 0 ) R 2 ( s 0 ) R 1 ( s ) R 2 ( s ) .
I ( s ) = I 0 ( s 0 ) [ d R 1 ( s ) 1 ] [ d R 2 ( s ) 1 ] ,
I ( s ) = I 0 ( s 0 ) R 1 ( s 0 ) d R 1 ( s 0 ) R 2 ( s 0 ) d R 2 ( s 0 ) .
I ( s ) I ( s ) [ 1 Δ ( 1 R 1 ( s 0 ) ) ( 1 R 1 ( s 0 ) 1 d ) ] [ 1 Δ ( 1 R 2 ( s 0 ) ) ( 1 R 2 ( s 0 ) 1 d ) ] ,
< P >= I ( r ) 1 k ϕ ( r ) .
= x + z e z ,
x = x e x + y e y .
ϕ ( r ) = x ϕ ( x , z = 0 ) + k e z ,
k I ( x , z ) z = x [ I ( x , z ) x ϕ ( x , z = 0 ) ] .
d d s = s ,
1 k I ( r ) 2 ϕ ( r ) + 1 k ϕ ( r ) I ( r ) = 0 .
1 k 2 ϕ ( r ) + 1 k ϕ ( r ) ln I ( r ) = 0 .
d d s ln I ( s ) = 1 k 2 ϕ ( r ) .
I ( s ) = I ( x 0 , y 0 , z 0 ) e s 0 s 1 k 2 ϕ ( r ) d s .
ϕ ( r ) = k r = k x x + k y y + k z z .
I ( s ) = I [ x ( s ) , y ( s ) , z ( s ) ] = I ( x 0 , y 0 , z 0 ) ,
Select as filters


Select Topics Cancel
© Copyright 2024 | Optica Publishing Group. All rights reserved, including rights for text and data mining and training of artificial technologies or similar technologies.