0801.4523/T2.tex
1: % Template article for preprint document class `elsart'
2: % with harvard style bibliographic references
3: % SP 2001/01/05
4: 
5: \documentclass{elsart}
6: 
7: % Use the option doublespacing or reviewcopy to obtain double line spacing
8: %\documentclass[doublespacing]{elsart}
9: 
10: % the natbib package allows both number and author-year (Harvard)
11: % style referencing;
12: \usepackage[square,comma]{natbib}
13: % if you use PostScript figures in your article
14: % use the graphics package for simple commands
15: % \usepackage{graphics}
16: % or use the graphicx package for more complicated commands
17: \usepackage{graphicx}
18: \usepackage{pxfonts}
19: % or use the epsfig package if you prefer to use the old commands
20: % \usepackage{epsfig}
21: 
22: % The amssymb package provides various useful mathematical symbols
23: \usepackage{amssymb}
24: 
25: 
26: 
27: \begin{document}
28: 
29: 
30: 
31: 
32: \begin{frontmatter}
33: 
34: % Title, authors and addresses
35: 
36: % use the thanksref command within \title, \author or \address for footnotes;
37: % use the corauthref command within \author for corresponding author footnotes;
38: % use the ead command for the email address,
39: % and the form \ead[url] for the home page:
40: % \title{Title\thanksref{label1}}
41: % \thanks[label1]{}
42: % \author{Name\corauthref{cor1}\thanksref{label2}}
43: % \ead{email address}
44: % \ead[url]{home page}
45: % \thanks[label2]{}
46: % \corauth[cor1]{}
47: % \address{Address\thanksref{label3}}
48: % \thanks[label3]{}
49: 
50: \title{Transverse coherence properties of X-ray beams in third-generation synchrotron radiation sources}
51: 
52: % use optional labels to link authors explicitly to addresses:
53: % \author[label1,label2]{}
54: % \address[label1]{}
55: % \address[label2]{}
56: 
57: \author{Gianluca Geloni}
58: \author{Evgeni Saldin}
59: \author{Evgeni Schneidmiller}
60: \author{and Mikhail Yurkov}
61: 
62: \address{Deutsches Elektronen-Synchrotron (DESY), Hamburg,
63: Germany}
64: 
65: 
66: 
67: \begin{abstract}
68: This article describes a complete theory of spatial coherence for
69: undulator radiation sources. Current estimations of coherence
70: properties often assume that undulator sources are
71: quasi-homogeneous, like thermal sources, and rely on the
72: application of the van Cittert-Zernike theorem for calculating the
73: degree of transverse coherence. Such assumption is not adequate
74: when treating third generation light sources, because the vertical
75: (geometrical) emittance of the electron beam is comparable or even
76: much smaller than the radiation wavelength in a very wide spectral
77: interval that spans over four orders of magnitude (from $0.1
78: \mathrm{\AA}$ up to $10^3 \mathrm{\AA}$). Sometimes, the so-called
79: Gaussian-Schell model, that is widely used in statistical optics
80: in the description of partially-coherent sources, is applied as an
81: alternative to the quasi-homogeneous model. However, as we will
82: demonstrate, this model fails to properly describe coherent
83: properties of X-ray beams from non-homogeneous undulator sources.
84: As a result, a more rigorous analysis is required. We propose a
85: technique, based on statistical optics and Fourier optics, to
86: explicitly calculate the cross-spectral density of an undulator
87: source in the most general case, at any position after the
88: undulator. Our theory, that makes consistent use of dimensionless
89: analysis, allows relatively easy treatment and physical
90: understanding of many asymptotes of the parameter space, together
91: with their region of applicability. Particular emphasis is given
92: to the asymptotic situation when the horizontal emittance is much
93: larger than the radiation wavelength, and the vertical emittance
94: is arbitrary. This case is practically relevant for third
95: generation synchrotron radiation sources.
96: \end{abstract}
97: 
98: \begin{keyword}
99: 
100: % keywords here, in the form: keyword \sep keyword
101: X-ray beams \sep Undulator radiation \sep Transverse coherence
102: \sep Van Cittert-Zernike theorem \sep Emittance effects
103: 
104: % PACS codes here, in the form: \PACS code \sep code
105: \PACS 41.60.m \sep 41.60.Ap \sep 41.50 + h \sep 42.50.Ar
106: 
107: \end{keyword}
108: 
109: \end{frontmatter}
110: 
111: % main text
112: 
113: \clearpage
114: \section{\label{sec:intro} Introduction}
115: 
116: In recent years, continuous evolution of synchrotron radiation
117: (SR) sources has resulted in a dramatic increase of brilliance
118: with respect to older designs. Among the most exciting properties
119: of third generation facilities of today is a high flux of coherent
120: X-rays \cite{EDGA}. The availability of intense coherent X-ray
121: beams has triggered the development of a number of new
122: experimental techniques based on coherence properties of light
123: such as scattering of coherent X-ray radiation, X-ray photon
124: correlation spectroscopy and phase contrast imaging. The
125: interested reader may find an extensive reference list to these
126: applications in \cite{PETR}.
127: 
128: Characterization of transverse coherence properties of SR at a
129: specimen position is fundamental for properly planning, conducting
130: and analyzing experiments involving above-mentioned techniques.
131: These objectives can only be met once coherence properties of
132: undulator sources are characterized and then propagated along the
133: photon beamline up to the specimen.
134: 
135: This paper is mainly dedicated to a description of transverse
136: coherence properties of SR sources. Therefore, it constitutes the
137: first step for tracking coherence properties through optical
138: elements, which can be only done when the source is quantitatively
139: described. Solution of the problem of evolution of radiation
140: properties in free-space is also given. This can be used to
141: characterize radiation at the specimen position when there are no
142: optical elements between the source and the specimen.
143: 
144: Since SR is a random process, the description of transverse
145: coherence properties of the source and its evolution should be
146: treated in terms of probabilistic statements. Statistical optics
147: \cite{GOOD, MAND, NEIL} presents most convenient tools to deal
148: with fluctuating electromagnetic fields. However, it was mainly
149: developed in connection with polarized thermal light that is
150: characterized by quite specific properties. Besides obeying
151: Gaussian statistics, polarized thermal light has two other
152: specific characteristics allowing for major simplifications of the
153: theory: stationarity and homogeneity (i.e. perfect incoherence) of
154: the source \cite{GOOD, MAND}.
155: 
156: As we will see, SR obeys Gaussian statistics, but stationarity and
157: homogeneity do not belong to SR fields. Thus, although the
158: language of statistical optics must be used to describe SR
159: sources, one should avoid a-priori introduction of an incorrect
160: model. In contrast to this, up to now it has been a widespread
161: practice to assume that undulator sources are perfectly incoherent
162: (homogeneous) and to draw conclusions about transverse coherence
163: properties of undulator light based on these assumptions
164: \cite{PETR}.
165: 
166: In particular, in the case of thermal light, transverse coherence
167: properties of the radiation can be found with the help of the
168: well-known van Citter-Zernike (VCZ) theorem. However, for third
169: generation light sources either planned or in operation, the
170: horizontal electron beam (geometrical)
171: emittance\footnote{Emittances are measured in m $\cdot$ rad.
172: However, radians are a dimensionless unit. Therefore, we can
173: present emittances measured in meters. This is particularly useful
174: in the present paper, since we need to compare emittances with
175: radiation wavelength.} is of order of $1 \div 3$ nm while the
176: vertical emittance is bound to the horizontal through a coupling
177: factor $\zeta \sim 10^{-2}$, corresponding to vertical emittances
178: of order $0.1 \div 0.3~ \mathrm{\AA}$. As we will show, these
179: facts imply that the VCZ theorem cannot be applied in the vertical
180: direction, aside for the hard X-ray limit at wavelengths shorter
181: than $0.1 ~\mathrm{\AA}$. Similar remarks hold for future sources,
182: like Energy Recovery Linac (ERL)-based spontaneous radiators. ERL
183: technology is expected to constitute the natural evolution of
184: today SR sources and to provide nearly fully diffraction-limited
185: sources in the $1 \AA$-range, capable of three order of magnitudes
186: larger coherent flux compared to third generation light sources
187: \cite{EDGA}. Horizontal and vertical emittance will be of order of
188: $0.3 \AA$, which rules out the applicability of the VCZ theorem.
189: 
190: The need for characterization of partially coherent undulator
191: sources is emphasized very clearly in reference \cite{HOWE}.
192: However, studying spatial-coherence properties of undulator
193: sources is not a trivial task. Difficulties arise when one tries
194: to simultaneously include the effect of intrinsic divergence of
195: undulator radiation, and of electron beam size and divergence.
196: 
197: An attempt to find the region of applicability of the VCZ theorem
198: for third generation light sources is reported in \cite{TAKA}.
199: Authors of \cite{TAKA} conclude that the VCZ theorem can only be
200: applied when SR sources are close to the diffraction limit. We
201: will see that the VCZ theorem can only be applied in the opposite
202: case.
203: 
204: A model to describe partially-coherent sources, the
205: Gaussian-Schell model, is widely used in statistical optics. Its
206: application to SR is described in \cite{COI1,COI2,COI3}. The
207: Gaussian-Schell model includes non-homogeneous sources, because
208: the typical transverse dimension of the source can be comparable
209: or smaller than the transverse coherence length. While
210: \cite{COI1,COI2,COI3} are of general theoretical interest, they do
211: not provide a satisfactory approximation of third generation SR
212: sources. In fact, as we will see, undulator radiation has very
213: specific properties that cannot be described in terms of
214: Gaussian-Schell model.
215: 
216: The first treatment of transverse coherence properties from SR
217: sources accounting for the specific nature of undulator radiation
218: and anticipating operation of third generation SR facilities is
219: given, in terms of Wigner distribution, in \cite{KIM2,KIM3}.
220: References \cite{KIM2,KIM3} present the most general algorithm for
221: calculating properties of undulator radiation sources.
222: 
223: In the present paper we base our theory on the characterization of
224: the cross-spectral density of the system. The cross-spectral
225: density is merely the statistical correlation function of the
226: radiation field at two different positions on the observation
227: plane at a given frequency. It is equivalent to the Wigner
228: distribution, the two quantities being related by Fourier
229: transformation. Based on cross-spectral density we developed a
230: comprehensive theory of third generation SR sources in the
231: space-frequency domain, where we exploited the presence of small
232: or large parameters intrinsic in the description of the system.
233: First, we took advantage of the particular but important situation
234: of perfect resonance, when the field from the undulator source can
235: be presented in terms of analytical functions. Second, we
236: exploited the practical case of a Gaussian electron beam, allowing
237: further analytical simplifications. Third, we took advantage of
238: the large horizontal emittance (compared with the radiation
239: wavelength) of the electron beam, allowing separate treatments of
240: horizontal and vertical directions. Finally, we studied asymptotic
241: cases of our theory for third generation light sources with their
242: region of applicability. In particular, we considered both
243: asymptotes of a small and a large vertical emittance compared with
244: the radiation wavelength, finding surprising results. In the case
245: of a small vertical emittance (smaller than the radiation
246: wavelength) we found that the radiation is not diffraction limited
247: in the vertical direction, an effect that can be ascribed to the
248: influence of the large horizontal emittance on the vertical
249: coherence properties of radiation. In the case of a large vertical
250: emittance we described the quasi-homogeneous source in terms of a
251: non-Gaussian model, which has higher accuracy and wider region of
252: applicability with respect to a simpler Gaussian model. In fact,
253: the non-Gaussian model proposed in our work accounts for
254: diffraction effects. Nonetheless, in order to solve the
255: propagation problem in free-space, a geometrical optics approach
256: can still be used.
257: 
258: We organize our work as follows. First, a general algorithm for
259: the calculation of the field correlation function of an undulator
260: source is given in Section \ref{sec:due}. Such algorithm is not
261: limited to the description of third generation light sources, but
262: can also be applied, for example, to ERL sources. In Section
263: \ref{sec:main} we specialize our treatment to third generation
264: light sources exploiting a large horizontal emittance compared to
265: the radiation wavelength, and studying the asymptotic limit for a
266: small vertical emittance of the electron beam. In the following
267: Section \ref{sec:quas} we consider the particular case of
268: quasi-homogeneous undulator sources, and discuss the applicability
269: range of the van VCZ theorem. Finally, in Section \ref{sec:conc}
270: we come to conclusions.
271: 
272: \section{\label{sec:due} Cross-spectral density of an undulator source}
273: 
274: \subsection{\label{sub:def} Thermal light and Synchrotron Radiation: some concepts and definitions}
275: 
276: The great majority of optical sources emits thermal light. Such is
277: the case of the sun and the other stars, as well as of
278: incandescent lamps. This kind of radiation consists of a large
279: number of independent contributions (radiating atoms) and is
280: characterized by random amplitudes and phases in space and time.
281: Its electromagnetic field can be conveniently described in terms
282: of statistical optics that has been intensively developed during
283: the last few decades \cite{GOOD, MAND}.
284: 
285: \begin{figure}
286: \begin{center}
287: \includegraphics*[width=140mm]{figure1.eps}% Here is how to import EPS art
288: \caption{\label{thermal} Light intensity from an incandescent lamp
289: driven by a constant electric current. A statistically stationary
290: wave has an average that does not vary with time.}
291: \end{center}
292: \end{figure}
293: %
294: Consider the light emitted by a thermal source passing through a
295: polarization analyzer, as in Fig. \ref{thermal}. Properties of
296: polarized thermal light are well-known in statistical optics, and
297: are referred to as properties of completely chaotic, polarized
298: light \cite{GOOD, MAND}. Thermal light is a statistical random
299: process and statements about such process are probabilistic
300: statements. Statistical processes are handled using the concept of
301: statistical ensemble, drawn from statistical mechanics, and
302: statistical averages are performed over an ensemble, or many
303: realizations, or outcomes of the statistical process under study.
304: Polarized thermal sources obey a very particular kind of random
305: process in that it is Gaussian and stationary\footnote{Ergodic
306: too. From a qualitative viewpoint, a given random process is
307: ergodic when all ensemble averages can be substituted by time
308: averages. Ergodicity is a stronger requirement than stationarity
309: \cite{GOOD,MAND}.}. Moreover, they are homogeneous.
310: 
311: The properties of Gaussian random processes are well-known in
312: statistical optics. For instance, the real and imaginary part of
313: the complex amplitudes of the electric field from a polarized
314: thermal source have Gaussian distribution, while the instantaneous
315: radiation power fluctuates in accordance with the negative
316: exponential distribution. It can be shown \cite{GOOD} that a
317: linearly filtered Gaussian process is also a Gaussian random
318: process. As a result, the presence of a monochromator and a
319: spatial filter as in the system depicted in Fig. \ref{thermal} do
320: not change the statistics of the signal. Finally, higher order
321: field correlation functions can be found in terms of the second
322: order correlation. This dramatically simplifies the description of
323: the random process.
324: 
325: Stationarity is a subtle concept. There are different kinds of
326: stationarity. However, for Gaussian processes different kinds of
327: stationarity coincide \cite{GOOD,MAND}. In this case, stationarity
328: means that all ensemble averages are independent of time. It
329: follows that a necessary condition for a certain process to be
330: stationary is that the signal last forever. Yet, if a signal lasts
331: much longer than the short-scale duration of the field
332: fluctuations (its coherence time $\tau_c$)  and it is observed for
333: a time much shorter than its duration $\sigma_T$, but much longer
334: than $\tau_c$ it can be reasonably considered as everlasting and
335: it has a chance to be stationary as well, as in the case of
336: thermal light.
337: 
338: Finally, thermal sources are homogeneous. The field is correlated
339: on the transverse direction over the possible minimal distance,
340: which is of order of a wavelength. This means that the radiation
341: intensity at the source remains practically unvaried on the scale
342: of a correlation length. One can say, equivalently, that the
343: source is homogeneous. Homogeneity is the equivalent, in the
344: transverse direction, of stationarity and implies a constant
345: ensemble-averaged intensity along the transverse direction.
346: 
347: Consider now a SR source, as depicted\footnote{Radiation at the
348: detector consists of a carrier modulation of frequency $\omega$
349: subjected to random amplitude and phase modulation. The Fourier
350: decomposition of the radiation contains frequencies spread about
351: the monochromator bandwidth: it is not possible, in practice, to
352: resolve the oscillations of the radiation fields which occur at
353: the frequency of the carrier modulation. Therefore, for comparison
354: with experimental results, we average the theoretical results over
355: a cycle of oscillations of the carrier modulation.} in Fig.
356: \ref{SR}. Like thermal light, also SR is a random process. In
357: fact, relativistic electrons in a storage ring emit SR passing
358: through bending magnets or undulators. The electron beam shot
359: noise causes fluctuations of the beam density which are random in
360: time and space from bunch to bunch. As a result, the radiation
361: produced has random amplitudes and phases. Moreover, in Section
362: \ref{sub:seco} we will demonstrate that SR fields obey Gaussian
363: statistics. Statistical properties satisfied by single-electron
364: contributions (elementary phasors) to the total SR field are
365: weaker than those satisfied by single-atom contributions in
366: thermal sources. Thus, the demonstration that thermal light obeys
367: Gaussian statistics cannot be directly applied to the SR case, and
368: some condition should be formulated on the parameter space to
369: define the region where SR is indeed a Gaussian random process. We
370: will show this fact with the help of Appendix A of \cite{OURU}.
371: 
372: In contrast with thermal light, SR is intrinsically
373: non-stationary, because it presents a time-varying
374: ensemble-averaged intensity on the temporal scale of the duration
375: of the X-ray pulse generated by a single electron bunch. For this
376: reason, in what follows the averaging brackets $\langle ...
377: \rangle$ will always indicate the ensemble average over bunches
378: (and not a time average).
379: %
380: \begin{figure}
381: \begin{center}
382: \includegraphics*[width=140mm]{figure2.eps}% Here is how to import EPS art
383: \caption{\label{SR} The intensity of an X-ray beam from a SR
384: source. A statistically non-stationary wave has a time-varying
385: intensity averaged over an ensemble of bunches.}
386: \end{center}
387: \end{figure}
388: %
389: Finally, SR sources are not completely incoherent, or homogeneous.
390: In fact, there is a close connection between the state of
391: coherence of the source and the angular distribution of the
392: radiant intensity. A thermal source that is correlated over the
393: minimal possible distance (which is of order of the wavelength) is
394: characterized by a radiant intensity distributed over a solid
395: angle of order $2 \pi$. This is not the case of SR light, that is
396: confined within a narrow cone in the forward direction. The high
397: directionality of SR rules out the possibility of description in
398: terms of thermal light. As we will see, depending on the
399: situation, SR may or may not be described by a quasi-homogeneous
400: model, where sources are only locally coherent over a distance of
401: many wavelengths but the linear dimension of the source is much
402: larger than the correlation distance\footnote{Note that the high
403: directionality of SR is not in contrast with the poor coherence
404: which characterizes the quasi-homogeneous limit.}.
405: 
406: In spite of differences with respect to the simpler case of
407: thermal light, SR fields can be described in terms of statistical
408: optics. However, statistical optics was developed in relation with
409: thermal light emission. Major assumptions typical of this kind of
410: radiation like, for example, stationarity, are retained in
411: textbooks \cite{GOOD,MAND}. Thus, usual statistical optics
412: treatment must be modified in order to deal with SR problems. We
413: will reduce the problem of characterization of transverse
414: coherence properties of undulator sources in the space-frequency
415: domain to the calculation of the correlation of the field produced
416: by a single electron with itself. This correlation is known as
417: cross-spectral density. The non-stationarity of the process
418: imposes some (practically non restrictive) condition on the
419: parameter space region where a treatment based on cross-spectral
420: density can be applied to SR phenomena. Such condition involves
421: the length of the electron bunch, the number of undulator periods
422: and the radiation wavelength. Given the fact that the electron
423: bunch length can vary from about $10$ ps for third generation SR
424: sources to about $100$ fs  for ERL sources, understanding of the
425: region of applicability this condition is a-priori not obvious.
426: All these subjects will be treated in the next Section
427: \ref{sub:seco}.
428: 
429: 
430: 
431: 
432: 
433: 
434: 
435: \subsection{\label{sub:seco} Second-order correlations in
436: space-frequency domain}
437: 
438: 
439: 
440: 
441: In SR experiments with third generation light sources, detectors
442: are limited to about $100$ ps time resolution. Therefore, they
443: cannot resolve a single X-ray pulse in time domain, whose duration
444: is about $30$ ps. They work, instead, by counting the number of
445: photons at a certain frequency over an integration time longer
446: than the pulse. It is therefore quite natural to consider signals
447: in the frequency domain. With this in mind, let ${\bar{E}}_b(z,
448: \vec{r}_{{}}, \omega)$ be a fixed polarization component of the
449: Fourier transform of the electric field at location $(z,
450: \vec{r}_{{}})$, in some cartesian coordinate system, and frequency
451: $\omega$ by a given collection of electromagnetic sources. We will
452: often name it, slightly improperly, "the field". Subscript "b"
453: indicates that the field is generated by the entire bunch.
454: ${\bar{E}}_b(z, \vec{r}_{{}}, \omega)$ is linked to the time
455: domain field ${{E}}_b(z, \vec{r}_{{}}, t)$ through the Fourier
456: transform
457: 
458: \begin{equation}
459: \bar{{E}}_b(\omega) = \int_{-\infty}^{\infty} dt {{E}}_b(t) \exp(i
460: \omega t)~, ~~~~~{{E}}_b(t) = \frac{1}{2\pi}
461: \int_{-\infty}^{\infty} d\omega \bar{{E}}_b(\omega) \exp(-i \omega
462: t)~. \label{ftran}
463: \end{equation}
464: %
465: We will be interested in the case of an ultra relativistic
466: electron beam going through a certain magnetic system, an
467: undulator in particular. In this case $z$ is the observation
468: distance along the optical axis of the undulator and
469: $\vec{r}_{{}}$ fixes the transverse position of the observer. The
470: contribution of the $k$-th electron to the field depends on the
471: transverse offset $\vec{l}_{k}$ and deflection angles
472: $\vec{\eta}_{k}$ that the electron has at some reference point on
473: the optical axis $z$, e.g. the center of the undulator. Moreover,
474: the arrival time $t_k$ at the center of the undulator has the
475: effect of multiplying the field by a phase factor $\exp{(i\omega
476: t_k)}$, i.e. the time-domain electric field is retarded by a time
477: $t_k$. At this point we do not need to explicitly specify the
478: dependence on offset and deflection. The total field can be
479: written as
480: 
481: \begin{equation}
482: {\bar{E}}_b\left(z,\vec{r}_{{}},\omega\right)=\sum_{k=1}^{N_e}
483: \bar{E}\left(\vec{\eta}_k,\vec{l}_k,z,\vec{r}_{{}},\omega\right)
484: \exp{(i\omega t_k)} ~, \label{total}
485: \end{equation}
486: %
487: where $\vec{\eta}_k,\vec{l}_k$ and $t_k$ are random variables and
488: $N_e$ is the number of electrons in the bunch. Note that $\bar{E}$
489: in Eq. (\ref{total}) is a complex quantity, that can be written as
490: $\bar{E} = \mid\bar{E}_k\mid \exp(i \phi_k)$. It follows that the
491: SR field pulse at fixed frequency and position is a sum of a many
492: phasors, one for each electron, of the form $\bar{E} \exp{(i\omega
493: t_k)} = \mid\bar{E}_{k}\mid \exp(i \phi_k + i \omega t_k)$.
494: 
495: Elementary phasors composing the sum obey three statistical
496: properties, that are satisfied in SR problems of interest. First,
497: random variables $t_k$ are statistically independent of each
498: other, and of variables $\vec{\eta}_k$ and $\vec{l}_k$. Second,
499: random variables $\mid\bar{E}_{k}\mid$ (at fixed frequency
500: $\omega$), are identically distributed for all values of $k$, with
501: a finite mean $\langle \mid\bar{E}_{k}\mid \rangle$ and a finite
502: second moment $\langle \mid\bar{E}_{k}\mid^2 \rangle $. These two
503: assumptions follows from the properties of shot noise in a storage
504: ring, which is a fundamental effect related with quantum
505: fluctuations. Third, we assume that the electron bunch duration
506: $\sigma_T$ is large enough so that $\omega \sigma_T \gg 1$: under
507: this assumption, phases $\omega t_k$ can be regarded as uniformly
508: distributed on the interval $(0, 2\pi)$. The assumption $\omega
509: \sigma_T \gg 1$ exploits the first large parameter of our theory,
510: and is justified by the fact that $\omega$ is the undulator
511: resonant frequency, which is high enough, in practical cases of
512: interest, to guarantee that $\omega \sigma_T \gg 1$ for any
513: realistic choice of $\sigma_T$. Based on the three previously
514: discussed properties, and with the help of the central limit
515: theorem, it can be demonstrated\footnote{The proof follows from a
516: slight generalization of Section 2.9 in \cite{GOOD}. Namely, it
517: can be shown by direct calculation that real and imaginary part of
518: the total phasor ${\bar{E}}_b$ are uncorrelated, with zero mean
519: and equal variance. Then, using the central limit theorem, we
520: conclude that ${\bar{E}}_b$ is a circular complex Gaussian random
521: variable (at fixed $z$, $\vec{r}_{{}}$ and $\omega$).} that the
522: real and the imaginary part of $\bar{E}_b$ are distributed in
523: accordance to a Gaussian law.
524: 
525: As a result, SR is a non-stationary Gaussian random process.
526: 
527: Because of this, higher-order correlation functions can be
528: expressed in terms of second-order correlation functions with the
529: help of the moment theorem \cite{GOOD}. As a result, the knowledge
530: of the second-order field correlation function in frequency
531: domain, $\Gamma_\omega$, is all we need to completely characterize
532: the signal from a statistical viewpoint. The following definition
533: holds:
534: 
535: \begin{equation}
536: \Gamma_{\omega}\left(z,\vec{r}_{{} 1},\vec{r}_{{}
537: 2},\omega_1,\omega_2\right) = \left<
538: {\bar{E}}_b\left(z,\vec{r}_{{}
539: 1},\omega_1\right){\bar{E}}^*_b\left(z,\vec{r}_{{}
540: 2},\omega_2\right) \right>~, \label{gamma}
541: \end{equation}
542: %
543: where brackets $<...>$ indicate ensemble average over electron
544: bunches. For any given function $w\left(\vec{\eta}_k, \vec{l}_k,
545: t_k\right)$, the ensemble average is defined as
546: 
547: \begin{eqnarray}
548: \left<w\left(\vec{\eta}_k, \vec{l}_k, t_k\right)\right> = \int
549: d\vec{\eta}_k \int d\vec{l}_k \int_{-\infty}^{\infty} d t_k
550: w\left(\vec{\eta}_k, \vec{l}_k, t_k\right) P\left(\vec{\eta}_k,
551: \vec{l}_k, t_k\right) \label{ensembledef}~,
552: \end{eqnarray}
553: %
554: where integrals in $d\vec{l}_k$ and $d\vec{\eta}_k$ span over all
555: offsets and deflections, and $P=P(\vec{\eta}_k, \vec{l}_k, t_k)$
556: indicates the probability density distribution in the joint random
557: variables $\vec{\eta}_k$, $\vec{l}_k$, and $t_k$. Note that, since
558: all electrons have the same probability of arrival around a given
559: offset, deflection, and time, $P$ is independent of $k$. Moreover,
560: already discussed independence of $t_k$ from $\vec{l}_k$ and
561: $\vec{\eta}_k$ allows to write $P$ as
562: 
563: \begin{eqnarray}
564: P\left(\vec{\eta}_k, \vec{l}_k, t_k\right) =
565: f_\bot\left(\vec{l}_k,\vec{\eta}_k\right) f(t_k)~.
566: \label{independence}
567: \end{eqnarray}
568: %
569: Here $f$ is the longitudinal bunch profile of the electron beam,
570: while $f_\bot$ is the transverse phase-space distribution.
571: 
572: Substituting Eq. (\ref{total}) in Eq. (\ref{gamma}) one has
573: 
574: \begin{eqnarray}
575: \Gamma_{\omega} = \left<\sum_{m,n=1}^{N_e} \bar{E}
576: \left(\vec{\eta}_m ,\vec{l}_m , z ,\vec{r}_{{} 1}, \omega_1\right)
577: \bar{E}^* \left(\vec{\eta}_n,\vec{l}_n, z,\vec{r}_{{} 2},
578: \omega_2\right) \exp{[i(\omega_1 t_m- \omega_2 t_n)]} \right> .\cr
579: && \label{gamma2}
580: \end{eqnarray}
581: %
582: Expansion of Eq. (\ref{gamma2}) gives
583: 
584: \begin{eqnarray}
585: &&\Gamma_{\omega} = \sum_{m=1}^{N_e} \left\langle \bar{E}
586: \left(\vec{\eta}_m,\vec{l}_m,z,\vec{r}_{{} 1}, \omega_1\right)
587: \bar{E}^*\left(\vec{\eta}_m,\vec{l}_m,z,\vec{r}_{{} 2},
588: \omega_2\right) \exp{[i( \omega_1- \omega_2) t_m]} \right\rangle
589: \cr && + \sum_{m\ne n} \left\langle
590: \bar{E}\left(\vec{\eta}_m,\vec{l}_m,z,\vec{r}_{{} 1},
591: \omega_1\right) \exp{(i \omega_1
592: t_m)}\right\rangle\left\langle\bar{E}^*
593: \left(\vec{\eta}_n,\vec{l}_n,z,\vec{r}_{{} 2}, \omega_2\right)
594: \exp{(- i \omega_2 t_n)} \right\rangle.\cr && \label{gamma3}
595: \end{eqnarray}
596: %
597: With the help of Eq. (\ref{ensembledef}) and Eq.
598: (\ref{independence}), the ensemble average $\langle \exp{(i \omega
599: t_k)} \rangle$ can be written as the Fourier transform of the
600: longitudinal bunch profile function $f$, that is
601: 
602: \begin{equation}
603: \left\langle \exp{(i \omega t_k)} \right\rangle =
604: \int_{-\infty}^{\infty} d t_k f(t_k) \exp(i\omega t_k) \equiv
605: \bar{f}(\omega)~. \label{FTlong}
606: \end{equation}
607: %
608: Using Eq. (\ref{FTlong}), Eq. (\ref{gamma3}) can be written as
609: 
610: \begin{eqnarray}
611: \Gamma_{\omega} &&= \sum_{m=1}^{N_e} \bar{f}( \omega_1- \omega_2)
612: \left \langle \bar{E} \left(\vec{\eta}_m,\vec{l}_m, z, \vec{r}_{{}
613: 1}, \omega_1\right) \bar{E}^*
614: \left(\vec{\eta}_m,\vec{l}_m,z,\vec{r}_{{} 2}, \omega_2\right)
615: \right\rangle \cr && + \sum_{m\ne n} \bar{f}( \omega_1)\bar{f}(-
616: \omega_2)  \left\langle
617: \bar{E}\left(\vec{\eta}_m,\vec{l}_m,z,\vec{r}_{{} 1},
618: \omega_1\right) \right\rangle
619: \left\langle\bar{E}^*\left(\vec{\eta}_n, \vec{l}_n,z,\vec{r}_{{}
620: 2}, \omega_2\right) \right\rangle ~, \label{gamma4}
621: \end{eqnarray}
622: %
623: where $\bar{f}^*(\omega_2)=\bar{f}(- \omega_2)$ because ${f}$ is a
624: real function. When the radiation wavelengths of interest are much
625: shorter than the bunch length we can safely neglect the second
626: term on the right hand side of Eq. (\ref{gamma4}) since the form
627: factor product $\bar{f}( \omega_1) \bar{f}(- \omega_2)$ goes
628: rapidly to zero for frequencies larger than the characteristic
629: frequency associated with the bunch length: think for instance, at
630: a centimeter long bunch compared with radiation in the Angstrom
631: wavelength range\footnote{When the radiation wavelength of
632: interested is comparable or longer than the bunch length, the
633: second term in Eq. (\ref{gamma4}) is dominant with respect to the
634: first, because it scales with the number of particles
635: \textit{squared}: in this case, analysis of the second term leads
636: to a treatment of coherent synchrotron radiation phenomena (CSR).
637: In this paper we will not be concerned with CSR and we will
638: neglect the second term in Eq. (\ref{gamma4}), assuming that the
639: radiation wavelength of interest is shorter than the bunch length.
640: Also note that $\bar{f}( \omega_1- \omega_2)$ depends on the
641: \textit{difference} between $ \omega_1$ and $ \omega_2$, and the
642: first term cannot be neglected.}. Therefore we write
643: 
644: \begin{eqnarray}
645: \Gamma_{\omega} &&= \sum_{m=1}^{N_e} \bar{f}( \omega_1- \omega_2)
646: \left \langle \bar{E} \left(\vec{\eta}_m,\vec{l}_m, z, \vec{r}_{{}
647: 1}, \omega_1\right) \bar{E}^*
648: \left(\vec{\eta}_m,\vec{l}_m,z,\vec{r}_{{} 2}, \omega_2\right)
649: \right\rangle \cr && = N_e \bar{f}( \omega_1- \omega_2) \left
650: \langle \bar{E}\left(\vec{\eta},\vec{l}, z, \vec{r}_{{} 1},
651: \omega_1\right) \bar{E}^* \left(\vec{\eta},\vec{l},z,\vec{r}_{{}
652: 2}, \omega_2\right) \right\rangle  ~. \label{gamma5}
653: \end{eqnarray}
654: %
655: As one can see from Eq. (\ref{gamma5}) each electron is correlated
656: just with itself: cross-correlation terms between different
657: electrons was included in the second term on the right hand side
658: of Eq. (\ref{gamma4}), that has been dropped.
659: 
660: If the dependence of $\bar{E}$ on $ \omega$ is slow enough,
661: $\bar{E}$ does not vary appreciably on the characteristic scale of
662: $\bar{f}$. Thus, we can substitute
663: $\bar{E}^*(\vec{\eta},\vec{l},z,\vec{r}_{{} 2}, \omega_2)$ with
664: $\bar{E}^*(\vec{\eta},\vec{l},z,\vec{r}_{{} 2}, \omega_1)$ in Eq.
665: (\ref{gamma5}).
666: %
667: \begin{figure}
668: \begin{center}
669: \includegraphics*[width=110mm]{figure3.eps}% Here is how to import EPS art
670: \caption{\label{rela} Schematic illustration of the relative
671: frequency dependence of the spectral correlation function
672: $\bar{f}(\omega-\omega')$ and of the cross-spectral density
673: function (the cross-power spectrum) $G(z,\vec{r}_{{}
674: 1},\vec{r}_{{} 2}, \omega)$ of SR at points $\vec{r}_{{} 1}$ and
675: $\vec{r}_{{} 2}$ at frequency $\omega$.}
676: \end{center}
677: \end{figure}
678: %
679: The situation is depicted in Fig. \ref{rela}. On the one hand, the
680: characteristic scale of $\bar{f}$ is given by $1/\sigma_T$, where
681: $\sigma_T$ is the characteristic bunch duration. On the other
682: hand, the minimal possible bandwidth of undulator radiation is
683: achieved on axis and in the case of a filament beam. It is peaked
684: around the resonant frequency ${\omega_r} = {2 \gamma_z^2
685: c}/{\lambdabar_w}$ ($\lambda_w$ being the undulator period and
686: $\gamma_z$ the average longitudinal Lorentz factor) and amounts to
687: $\omega_r/N_w$, $N_w\gg 1$ being the number of undulator periods
688: (of order $10^2 - 10^3$). Since $\omega_r/N_w$ is a minimum for
689: the radiation bandwidth, it should be compared with $1/\sigma_T$.
690: For instance, at wavelengths of order $1 \AA$, $N_w \sim 10^3$ and
691: $\sigma_T \sim 30$ ps (see \cite{PETR}), one has $\omega_r/N_w
692: \sim 2\cdot 10^{16}$ Hz, which is much larger than $1/\sigma_T
693: \sim 3\cdot 10^{10}$ Hz. From this discussion follows that, in
694: practical situations of interest, we can simplify Eq.
695: (\ref{gamma5}) to
696: 
697: \begin{eqnarray}
698: \Gamma_{\omega}(z,\vec{r}_{{} 1},\vec{r}_{{} 2}, \omega_1,
699: \omega_2) = N_e \bar{f}( \omega_1- \omega_2) G\left(z,\vec{r}_{{}
700: 1},\vec{r}_{{} 2}, \omega_1\right)~, \label{gamma6prima}
701: \end{eqnarray}
702: %
703: where
704: 
705: \begin{equation}
706: G(z,\vec{r}_{{} 1},\vec{r}_{{} 2}, \omega) \equiv \left\langle
707: \bar{E} \left(\vec{\eta},\vec{l},z,\vec{r}_{{} 1}, \omega\right)
708: \bar{E}^*\left(\vec{\eta},\vec{l},z,\vec{r}_{{} 2}, \omega\right)
709: \right\rangle~.\label{coore}
710: \end{equation}
711: %
712: Eq. (\ref{gamma6prima}) fully characterizes the system under study
713: from a statistical viewpoint. Correlation in frequency and space
714: are expressed by two separate factors. In particular, spatial
715: correlation is expressed by the cross-spectral density
716: function\footnote{Note, however, that $G$ depends on $\omega$.}
717: $G$. In other words, we are able to deal separately with spatial
718: and spectral part of the correlation function in space-frequency
719: domain under the non-restrictive assumption $\omega_r/N_w \gg
720: 1/\sigma_T$.
721: 
722: Eq. (\ref{gamma6prima}) is the result of our theoretical analysis
723: of the second-order correlation function in the space-frequency
724: domain. We can readily extend this analysis to the case when the
725: observation plane is behind a monochromator with transfer function
726: $T(\omega)$. In this case, Eq. (\ref{gamma6prima}) modifies to
727: 
728: \begin{eqnarray}
729: \Gamma_{\omega}\left(z,\vec{r}_{{} 1},\vec{r}_{{} 2}, \omega_1,
730: \omega_2\right) = N_e \bar{f}( \omega_1- \omega_2)
731: G\left(z,\vec{r}_{{} 1},\vec{r}_{{} 2}, \omega_1\right)
732: T(\omega_1) T^*(\omega_2)~. \label{gamma6}
733: \end{eqnarray}
734: %
735: It is worth to note that, similarly to Eq. (\ref{gamma6prima}),
736: also Eq. (\ref{gamma6}) exhibits separability of correlation
737: functions in frequency and space.
738: 
739: From now on we will be concerned with the calculation of the
740: cross-spectral density $G(z,\vec{r}_{{} 1},\vec{r}_{{} 2},
741: \omega)$, independently of the shape of the remaining factors on
742: the right hand side of Eq. (\ref{gamma6}).
743: 
744: \begin{figure}
745: \begin{center}
746: \includegraphics*[width=140mm]{figure4.eps}% Here is how to import EPS art
747: \caption{\label{ed12c} Measurement of the cross-spectral density
748: of a undulator source. (a) Young's double-pinhole interferometer
749: demonstrating the coherence properties of undulator radiation. The
750: radiation beyond the pinholes must be spectrally filtered by a
751: monochromator or detector (not shown in figure). (b) The fringe
752: visibility of the resultant interference pattern.}
753: \end{center}
754: \end{figure}
755: %
756: Before proceeding we introduce, for future reference, the notion
757: of spectral degree of coherence $g$, that can be presented as a
758: function of $\vec{r}_{{} 1}$ and $\vec{r}_{{} 2}$:
759: 
760: \begin{equation}
761: g\left(\vec{r}_{{} 1},\vec{r}_{{} 2}\right) =
762: \frac{{G}\left(\vec{r}_{{} 1},\vec{r}_{{} 2}\right)}{\left\langle
763: \left |\bar{E}\left(\vec{r}_{{} 1}\right)\right|^2\right \rangle
764: ^{1/2} \left \langle \left|\bar{E}\left(\vec{r}_{{}
765: 2}\right)\right|^2\right\rangle^{1/2}} ~. \label{normfine}
766: \end{equation}
767: %
768: Consider Fig. \ref{ed12c}, depicting a Young's double-pinhole
769: interferometric measure. Results of Young's experiment vary with
770: $\vec{r}_{{} 1}$ and $\vec{r}_{{} 2}$. The modulus of the spectral
771: degree of coherence, $|g|$ is related with the fringe visibility
772: of the interference pattern. In particular, the relation between
773: fringes visibility $V$ and $g$ is given by
774: 
775: \begin{eqnarray}
776: V = 2 \frac{ \left<\left|\bar{E}\left(\vec{r}_{{}
777: 1}\right)\right|^2\right>^{1/2}
778: \left<\left|\hat{E}\left(\vec{r}_{{}
779: 2}\right)\right|^2\right>^{1/2}}{
780: \left<\left|\hat{E}\left(\vec{r}_{{}1}\right)\right|^2\right>+
781: \left<\left|\hat{E}\left(\vec{r}_{{}2}\right)\right|^2\right>}
782: \left|g\left(\vec{r}_{{}1},\vec{r}_{{}2}\right)\right|~.
783: \label{Vfrin}
784: \end{eqnarray}
785: %
786: Phase of $g$ is related to the position of the fringes instead.
787: Thus, spectral degree of coherence and cross-spectral density are
788: related with both amplitude \textit{and} position of the fringes
789: and are physically measurable quantities that can be recovered
790: from a Young's interference experiment.
791: 
792: \subsection{\label{sub:temp} Relation between space-frequency and
793: space-time domain}
794: 
795: This paper deals with transverse coherence properties of SR
796: sources in the space-frequency domain. However, it is interesting
797: to briefly discuss relations with the space-time domain, and
798: concepts like quasi-stationarity, cross-spectral purity and
799: quasi-monochromaticity that are often considered in literature
800: \cite{GOOD,MAND}.
801: 
802: First, the knowledge of $\Gamma_{\omega}$ in frequency domain is
803: completely equivalent to the knowledge of the mutual coherence
804: function \cite{WOLF}:
805: 
806: \begin{equation}
807: \Gamma_{t}(z,\vec{r}_{{} 1},\vec{r}_{{} 2},t_1,t_2) = \left<
808: {{E}}(z,\vec{r}_{{} 1},t_1){{E}}^*(z,\vec{r}_{{} 2},t_2) \right>~.
809: \label{gammatime}
810: \end{equation}
811: %
812: Next to $\Gamma_t$, the complex degree of coherence is defined as
813: 
814: \begin{eqnarray}
815: \gamma_t(z,\vec{r}_{{} 1},\vec{r}_{{} 2},t_1,t_2)  =
816: \frac{\Gamma_{t}(z,\vec{r}_{{} 1},\vec{r}_{{}
817: 2},t_1,t_2)}{\left[\Gamma_{t}(z,\vec{r}_{{} 1},\vec{r}_{{}
818: 1},t_1,t_1) \Gamma_{t}(z,\vec{r}_{{} 2},\vec{r}_{{}
819: 2},t_2,t_2)\right]^{1/2} }~.\label{comdeg}
820: \end{eqnarray}
821: %
822: The presence of a monochromator (see Eq. (\ref{gamma6})) is
823: related with a bandwidth of interest $\Delta \omega_\mathrm{m}$,
824: centered around a given frequency $\omega_o$ (typically, the
825: undulator resonant frequency). Then, $T(\omega)$ is peaked around
826: $\omega_o$ and rapidly goes to zero as we move out of the range $(
827: \omega_o-\Delta \omega_\mathrm{m}/2, \omega_o+\Delta
828: \omega_\mathrm{m}/2)$. Using Eq. (\ref{gamma6}) we write the
829: mutual coherence function as
830: 
831: \begin{eqnarray}
832: \Gamma_t(z,\vec{r}_{{} 1},\vec{r}_{{} 2},t_1,t_2) &=&
833: \frac{N_e}{(2\pi)^2} \int_{-\infty}^{\infty} d\omega_1
834: \int_{-\infty}^{\infty} d\omega_2 \bar{f}(\omega_1 - \omega_2)
835: T(\omega_1) T^*(\omega_2) \cr && \times G(z,\vec{r}_{{}
836: 1},\vec{r}_{{} 2},\omega_2) \exp{(-i\omega_1 t_1)} \exp{(i\omega_2
837: t_2)} ~.\label{trasfgammabreak}
838: \end{eqnarray}
839: %
840: If the characteristic bandwidth  of the monochromator, $\Delta
841: \omega_\mathrm{m}$, is large enough so that $T$ does not vary
842: appreciably on the characteristic scale of $\bar{f}$, i.e. $\Delta
843: \omega_\mathrm{m} \gg 1/\sigma_T$, then
844: $T(\omega_1)T^*(\omega_2)\bar{f}(\omega_1- \omega_2)$ is peaked at
845: $\omega_1-\omega_2=0$. In this case the process is
846: quasi-stationary. With the help of new variables $\Delta \omega =
847: \omega_1 - \omega_2$ and $\bar{\omega}=(\omega_1+\omega_2)/2$, we
848: can simplify Eq. (\ref{trasfgammabreak}) accounting for the fact
849: that $\bar{f}(\omega_1- \omega_2)$ is strongly peaked around
850: $\Delta \omega=0$. In fact we can consider $T(\omega_1)
851: T^*(\omega_2) G(z,\vec{r}_{{} 1},\vec{r}_{{} 2},\omega_1) \simeq
852: |T(\bar{\omega})|^2 G(z,\vec{r}_{{} 1},\vec{r}_{{} 2},
853: \bar{\omega})$, so that we can integrate separately in $\Delta
854: \omega$ and $\bar{\omega}$ to obtain
855: 
856: \begin{eqnarray}
857: \Gamma_t(z,\vec{r}_{{} 1},\vec{r}_{{} 2},t_1,t_2) &=&
858: \frac{N_e}{(2\pi)^2} \int_{-\infty}^{\infty} d\Delta \omega
859: ~\bar{f}(\Delta \omega) \exp{\left(-i \Delta{\omega}
860: \bar{t}\right)} \cr && \times \int_{-\infty}^{\infty}
861: d\bar{\omega} ~|T(\bar{\omega})|^2 G(z,\vec{r}_{{} 1},\vec{r}_{{}
862: 2},\bar{\omega}) \exp{\left[-i \bar{\omega} \Delta t
863: \right]}~.\label{trasfgammabreak2}
864: \end{eqnarray}
865: %
866: where $\bar{t} = (t_1+t_2)/2$ and $\Delta t = t_1 - t_2$. This
867: means that the mutual coherence function can be factorized as
868: $\Gamma_t(\bar{t}, \Delta t) = f(\bar{t}) \widetilde{G}_t(\Delta
869: t)$. In the case no monochromator is present, $\widetilde{G}_t$
870: coincides with the Fourier transform of the cross-spectral
871: density, and the correspondent correlation function
872: $\Gamma_\omega(\bar{\omega},\Delta \omega)$ has been seen to obey
873: Eq. (\ref{gamma5}). Therefore:
874: 
875: $(a_1)$ The temporal correlation function $\widetilde{G}_t(\Delta
876: t)$ and the spectral density distribution of the source
877: $|T(\bar{\omega})|^2 G(\bar{\omega})$ form a Fourier pair.
878: 
879: $(b_1)$ The intensity distribution of the radiation pulse
880: $f(\bar{t})$ and the spectral correlation function $\bar{f}(\Delta
881: \omega)$ form a Fourier pair.
882: 
883: \begin{figure}
884: \begin{center}
885: \includegraphics*[width=140mm]{figure5.eps}% Here is how to import EPS art
886: \caption{\label{avecarr} Reciprocal width relations of Fourier
887: transform pairs.}
888: \end{center}
889: \end{figure}
890: %
891: Statement $(a_1)$ can be regarded as an analogue, for
892: quasi-stationary sources, of the well-known Wiener-Khinchin
893: theorem, which applies to stationary sources and states that the
894: temporal correlation function and the spectral density are a
895: Fourier pair.  Since there is symmetry between time and frequency
896: domains, an inverse Wiener-Khinchin theorem must also hold, and
897: can be obtained by the usual Wiener-Khinchin theorem by exchanging
898: frequencies and times. This is simply constituted by statement
899: $(b_1)$. Intuitively, statements $(a_1)$ and $(b_1)$ have their
900: justification in the reciprocal width relations of Fourier
901: transform pairs (see Fig. \ref{avecarr}).
902: 
903: It should be stressed that statistical optics is almost always
904: applied in the stationary case \cite{GOOD,MAND}. Definitions of
905: $\Gamma_t$ and $\gamma_t$ are also usually restricted to such
906: case. The case of a stationary process corresponds to the
907: asymptote of a Dirac $\delta$-function for the spectral
908: correlation function $\bar{f}$. The inverse Wiener-Khinchin
909: theorem applied to this asymptotic case would imply an infinitely
910: long radiation pulse, i.e. an infinitely long electron beam. In
911: contrast to this, the width of $\bar{f}$ is finite, and
912: corresponds to a finite width of $f$ of about $30$ ps. Thus, one
913: cannot talk about stationarity. However, when $\Delta
914: \omega_\mathrm{m} \gg 1/\sigma_T$, the spectral width of the
915: process (i.e. the width of $|T|^2G$ in $\omega$) is much larger
916: than the width of $\bar{f}$. In this case, the process is
917: quasi-stationary. The situation changes completely if a
918: monochromator with a bandwidth $\Delta \omega_\mathrm{m} \lesssim
919: 1/\sigma_T$ is present. In this case, Eq. (\ref{trasfgammabreak2})
920: cannot be used anymore, and one is not allowed to treat the
921: process as quasi-stationary. In the large majority of the cases
922: monochromator characteristics are not good enough to allow
923: resolution of $\bar{f}$. There are, however, some special cases
924: when $\Delta \omega_\mathrm{m} \lesssim 1/\sigma_T$. For instance,
925: in \cite{NOST} a particular monochromator is described with a
926: relative resolution of $10^{-8}$ at wavelengths of about $1 \AA$,
927: or $\omega_o \sim 2 \cdot 10^{19}$ Hz. Let us consider, as in
928: \cite{NOST}, the case of radiation pulses of $32$ ps duration.
929: Under the already accepted assumption $1/\sigma_T \ll
930: \omega_o/N_w$, we can identify the radiation pulse duration with
931: $\sigma_T$. Then we have $\Delta \omega_\mathrm{m} \sim 2 \cdot
932: 10^{11}$ Hz  which is of order of $2 \pi/\sigma_T \sim 2 \cdot
933: 10^{11}$ Hz: this means that the monochromator has the capability
934: of resolving $\bar{f}$.
935: 
936: Cases discussed up to now deal with radiation that is not
937: cross-spectrally pure \cite{MAN2} (or \cite{MAND}, paragraph
938: 4.5.1). In fact, the absolute value of the spectral degree of
939: coherence $g$ is a function of $\omega$. Moreover, as remarked in
940: \cite{COI1,COI2,COI3}, the spectrum of undulator radiation depends
941: on the observation point. This fact can also be seen in the time
942: domain from Eq. (\ref{trasfgammabreak}) or Eq.
943: (\ref{trasfgammabreak2}), because the complex degree of coherence
944: $\gamma_t$ cannot be split into a product of temporal and spatial
945: factors. However, if we assume $\Delta \omega_\mathrm{m} N_w /
946: \omega_o \ll 1$ (that is usually true), $G(z,\vec{r}_{{}
947: 1},\vec{r}_{{} 2},\omega)$ $ = G(z,\vec{r}_{{} 1},\vec{r}_{{}
948: 2},\omega_o)$ is a constant function of frequency within the
949: monochromator line, i.e. it is independent on the frequency
950: $\omega$. As a result $\Gamma_t$ in Eq. (\ref{trasfgammabreak})
951: can be split in the product of a temporal and a spatial factor and
952: therefore, in this case, light is cross-spectrally pure:
953: 
954: 
955: \begin{equation}
956: \Gamma_t(z_o,\vec{r}_{{} o1},\vec{r}_{{} o2},t_1,t_2) = N_e
957: \widetilde{g}_t(t_1,t_2) G\left(z,\vec{r}_{{} 1},\vec{r}_{{}
958: 2},\omega_o\right)~, \label{CScasegen}
959: \end{equation}
960: %
961: where $\widetilde{g}_t(t_1,t_2)$ is defined by
962: 
963: \begin{eqnarray}
964: \widetilde{g}_t = \frac{1}{(2\pi)^2} \int_{-\infty}^{\infty}
965: d\omega_1 \int_{-\infty}^{\infty} d\omega_2 \bar{f}(\omega_1 -
966: \omega_2) T(\omega_1) T^*(\omega_2) \exp{[i(\omega_2 t_2-\omega_1
967: t_1)]} ~.\cr &&\label{gtilda}
968: \end{eqnarray}
969: %
970: Note that, for instance, in the example considered before $\Delta
971: \omega_\mathrm{m} /\omega_o \simeq 10^{-8}$ and $N_w \simeq 10^2
972: \div 10^3$, i.e. $\Delta \omega_\mathrm{m} N_w / \omega_o \ll 1$.
973: 
974: 
975: It is important to remark that, since we are dealing with the
976: process in the space-frequency domain, whether the light is
977: cross-spectrally pure or not is irrelevant concerning the
978: applicability of our treatment, because we can study the
979: cross-spectral density $G$ for any frequency component.
980: 
981: Finally, for the sake of completeness, it is interesting to
982: discuss the relation between $G$ and the mutual intensity function
983: as usually defined in textbooks \cite{GOOD, MAND} in
984: \textit{quasimonochromatic} conditions. The assumption $\Delta
985: \omega_\mathrm{m} \gg 1/\sigma_T$ describes a quasi-stationary
986: process. In the limit $\sigma_T \longrightarrow \infty$ we have a
987: stationary process. Now letting $\Delta \omega_m \longrightarrow
988: 0$ slowly enough so that $\Delta \omega_\mathrm{m} \gg
989: 1/\sigma_T$,  Eq. (\ref{trasfgammabreak2}) remains valid while
990: both $\bar{f}(\Delta \omega)$ and $|T(\bar{\omega})|^2$ become
991: approximated better and better by Dirac $\delta$-functions,
992: $\delta(\Delta\omega)$ and $\delta(\bar{\omega}-\omega_o)$,
993: respectively. Then, $\Gamma_t \sim G \exp [-i \omega_o
994: (t_1-t_2)]$. Aside for an unessential factor, depending on the
995: normalization of $\bar{f}$ and $|T({\omega})|^2$, this relation
996: between $\Gamma_t$ and $G$ allows identification of the mutual
997: intensity function with $G$ as in \cite{GOOD,MAND}. In this case,
998: light is obviously cross-spectrally pure.
999: 
1000: 
1001: \subsection{\label{sub:fiel} Undulator field by a single particle with offset and deflection}
1002: 
1003: In order to give an explicit expression for the cross-spectral
1004: density of undulator radiation, we first need an explicit
1005: expression for $\bar{E} \left(\vec{\eta},\vec{l},z,\vec{r}_{{}},
1006: \omega\right)$, the field contribution from a single electron with
1007: given offset and deflection. This can be obtained by solving
1008: paraxial Maxwell's equations in the space-frequency domain with a
1009: Green's function technique. We refer the reader to \cite{OURF},
1010: where it was also shown that paraxial approximation always applies
1011: for ultrarelativistic systems with $1/\gamma^2\ll 1$, $\gamma$
1012: being the relativistic Lorentz factor. Since paraxial
1013: approximation applies, the envelope of the field
1014: $\widetilde{E}(\omega) = \bar{E}(\omega) \exp{(-{i\omega z}/{c})}$
1015: is a slowly varying function of $z$ on the scale of the wavelength
1016: $\lambda$, and for the sake of simplicity will also be named "the
1017: field". In Eq. (17) of reference \cite{OURF} an expression for
1018: $\vec{\widetilde{E}}(z, \vec{r},\omega)$ generated by an electron
1019: moving along a generic trajectory $\vec{r}(z)$ was found. Working
1020: out that equation under the resonance approximation for the case
1021: of a planar undulator where $\vec{r}_{}(z) = {K}/({\gamma k_w})
1022: \cos{(k_w z)} \vec{e}_x + \vec{l} +z \vec{\eta}$, $\vec{e}_x$
1023: being the unit vector in the $x$-direction, yields the
1024: horizontally polarized field
1025: 
1026: 
1027: \begin{eqnarray}
1028: \tilde{E}_{C}&=& \frac{(-e) K \omega   A_{JJ}}{2 c^2  \gamma}
1029: \int_{-L_w/2}^{L_w/2}  \frac{d z'}{{z}-{z}'} \exp \left\{i
1030: \left[\left(C+\frac{\omega \left|\vec{\eta}\right|^2}{2
1031: c}\right){z}' + \frac{\omega\left(\vec{r}-\vec{l}-\vec{\eta}z'
1032: \right)^2 }{2 c(z-z')}\right] \right\}~ .\cr &&
1033: \label{undunormfin00}
1034: \end{eqnarray}
1035: %
1036: Here we defined the detuning parameter $C = \omega
1037: /(2{\gamma}_z^2c)-k_w = ({\Delta\omega}/{\omega_r}) k_w$, where
1038: $\omega = \omega_r + \Delta \omega$. Thus, $C$ specifies "how
1039: much" $\omega$ differs from the fundamental resonance frequency
1040: ${\omega_r} = {2 \gamma_z^2 c}/{\lambdabar_w}$. The subscript "C"
1041: in $\widetilde{E}_C$ indicates that this expression is valid for
1042: arbitrary detuning parameter. Moreover, $\lambdabar_w \equiv 1/k_w
1043: = \lambda_w/(2\pi)$, $\lambda_w$ being the undulator period;
1044: $\gamma_z \equiv \gamma/(1+K^2/2)$; $K=(\lambda_w e H_w) / (2 \pi
1045: m_\mathrm{e} c^2)$ is the undulator parameter, $m_\mathrm{e}$
1046: being the electron mass and $H_w$ being the maximum of the
1047: magnetic field produced by the undulator on the $z$ axis. Finally,
1048: $L_w$  is the undulator length and $A_{JJ} \equiv
1049: J_0[{K^2}/({4+2K^2})] - J_1[ {K^2}/({4+2K^2})]$, $J_n$ indicating
1050: the Bessel function of the first kind of order $n$. It should be
1051: stressed that Eq. (\ref{undunormfio}) was derived under the
1052: resonance approximation meaning that the large parameter $N_w \gg
1053: 1$ was exploited, together with conditions $\Delta \omega/\omega_r
1054: \ll 1$ and $C+\omega r^2/(2 c z^2) \ll k_w$, meaning that we are
1055: looking at frequencies near the fundamental and angles within the
1056: main lobe of the directivity diagram. Moreover, the reader should
1057: keep in mind that no focusing elements are accounted for in the
1058: undulator. This fact is intrinsically related to the choice of
1059: $\vec{r}(z)$ done above.
1060: 
1061: Further algebraic manipulations  (see Appendix B of \cite{OURU})
1062: show that Eq. (\ref{undunormfin00}) can be rewritten as:
1063: 
1064: 
1065: \begin{eqnarray}
1066: \widetilde{E}_C &=& \frac{(-e) K \omega  A_{JJ}}{2 c^2 \gamma}
1067: \int_{-L_w/2}^{L_w/2} \frac{d z'}{z-z'} \cr && \times \exp
1068: \left\{i \left[C z' +
1069: \frac{\omega\left(\vec{r}-\vec{l}~~\right)^2}{2 c z} +\frac{\omega
1070: z z'}{2c (z-z')}
1071: \left(\frac{\vec{r}}{z}-\frac{\vec{l}}{z}-\vec{\eta}\right)^2
1072: \right] \right\} ~.\cr && \label{undunormfio}
1073: \end{eqnarray}
1074: %
1075: In this paper we will make a consistent use of dimensional
1076: analysis, which allows one to classify the grouping of dimensional
1077: variables in a way that is most suitable for subsequent study.
1078: Normalized units will be defined as
1079: 
1080: \begin{eqnarray}
1081: &&\hat{E} = -\frac{2 c^2  \gamma}{K \omega e  A_{JJ}}
1082: \widetilde{E}~, \cr && \vec{\hat{\eta}} =\vec{{\eta}}
1083: \sqrt{\frac{\omega L_w}{c}}  ~,\cr &&\hat{C} = L_w C = 2 \pi N_w
1084: \frac {\omega-\omega_r}{\omega_r}~,\cr&&\vec{\hat{r}}_{{} }
1085: ={\vec{r}}_{{}} \sqrt{{\omega \over{L_w c}}}~,\cr&&\vec{\hat{l}}
1086: ={\vec{l}}\sqrt{{\omega \over{L_w c}}}~,\cr&&
1087: \hat{z}={z\over{L_w}}~.\label{Cnorm}
1088: \end{eqnarray}
1089: %
1090: Moreover,  for any distance $\hat{z}$, we introduce
1091: $\vec{\hat{\theta}} = \vec{\hat{r}}/\hat{z}$. The algorithm for
1092: calculating the cross-spectral density will be formulated in terms
1093: of dimensionless fields. Therefore we re-write Eq.
1094: (\ref{undunormfio}) as
1095: 
1096: 
1097: 
1098: \begin{eqnarray}
1099: \hat{E}_C=  \exp\left[\frac{i
1100: \hat{z}}{2}\left|\vec{\hat{\theta}}-\frac{\vec{\hat{l}}}{\hat{z}}\right|^2\right]
1101: \Psi_{C} \left(\hat{z},\hat{C},
1102: \left|\vec{\hat{\theta}}-\frac{\vec{\hat{l}}}{\hat{z}}
1103: -\vec{\hat{\eta}}\right|\right)~, \label{undunormfin}
1104: \end{eqnarray}
1105: %
1106: where
1107: 
1108: \begin{eqnarray}
1109: \Psi_{C}(\hat{z},\hat{C},\alpha) \equiv \int_{-1/2}^{1/2}
1110: \frac{d\hat{z}'}{\hat{z}-\hat{z}'} \exp \left\{i
1111: \left[\hat{C}\hat{z}' +\frac{\hat{z}\hat{z}' \alpha^2
1112: }{2(\hat{z}-\hat{z}')}\right] \right\} ~.\label{psig}
1113: \end{eqnarray}
1114: %
1115: A physical picture of the evolution of the field along the $z$
1116: direction was given in reference \cite{OURF}, where Fourier optics
1117: ideas were used to develop a formalism ideally suited for the
1118: analysis of any SR problem. In that reference, the use of Fourier
1119: optics led to establish basic foundations for the treatment of SR
1120: fields, and in particular of undulator radiation, in terms of
1121: laser beam optics. Radiation from an ultra-relativistic electron
1122: can be interpreted as radiation from a virtual source, which
1123: produces a laser-like beam. In principle, such virtual source can
1124: be positioned everywhere down the beam, but there is a particular
1125: position where it is similar, in many aspects, to the waist of a
1126: laser beam. In the case of an undulator this location is the
1127: center of the insertion device. A virtual source located at that
1128: position ("the" virtual source) exhibits a plane wavefront.
1129: Therefore, it is completely specified by a real-valued amplitude
1130: distribution of the field (see Eq. (34) of \cite{OURF}). This
1131: amplitude can be derived from the far zone field distribution.
1132: Free-space propagation from the virtual source through the near
1133: zone and up to the far-zone, can be performed with the help of the
1134: Fresnel formula:
1135: 
1136: 
1137: 
1138: \begin{equation}
1139: { \hat{E}}( {z},\vec{r}_{}) = \frac{i }{2 \pi ( \hat{z}-
1140: \hat{z}_s)} \int d \vec{\hat{r}'}_{{}}~
1141: \hat{E}(\hat{z}_s,\vec{\hat{r}'}) \exp{\left[\frac{i
1142: \left|{\vec{\hat{r}}}-\vec{ \hat{r}'}\right|^2}{2  (\hat{z}-
1143: \hat{z}_s)}\right]}~, \label{fieldpropback}
1144: \end{equation}
1145: %
1146: where the integral is performed over the transverse plane, and
1147: $z_s$ is the virtual source position down the beamline.
1148: 
1149: These considerations were applied in \cite{OURF} to the case of
1150: undulator radiation under the applicability region of the resonant
1151: approximation. With reference to Fig. \ref{geo}, we let $z=0$ be
1152: the center of the undulator. Thus, the position of the virtual
1153: source is fixed in the center of the undulator too, $z_s=0$. For
1154: simplicity, the resonance condition with the fundamental harmonic
1155: was assumed satisfied, i.e. $\omega = \omega_r$. For this case, an
1156: analytical description of undulator radiation was provided.
1157: %
1158: \begin{figure}
1159: \begin{center}
1160: \includegraphics*[width=110mm]{figure6.eps}% Here is how to import EPS art
1161: \caption{\label{geo} Illustration of the undulator geometry and of
1162: the observation plane. }
1163: \end{center}
1164: \end{figure}
1165: %
1166: The horizontally-polarized field produced by a single electron
1167: with offset $\vec{\hat{l}}$ and deflection $\vec{\hat{\eta}}$ in
1168: the far-zone (i.e. at  $\hat{z} \gg 1$) can be represented by the
1169: scalar quantity\footnote{Note that for a particle moving on axis,
1170: at $\vec{\hat{l}}=0$ and $\vec{\hat{\eta}}=0$ the quadratic phase
1171: $\exp[i \hat{z} \hat{\theta}^2/2]$ in Eq. (\ref{undurad4}) is
1172: indicative of a spherical wavefront in paraxial approximation on
1173: the observation plane. When $\vec{\hat{l}}$ is different from
1174: zero, the laser-like beam is shifted, and this justifies the
1175: present of extra-factors including $\vec{\hat{l}}$. When the
1176: particle also has a deflection $\vec{\hat{\eta}}$, the laser-like
1177: beam is tilted, but the wavefront remains spherical. Since the
1178: observation plane remains orthogonal to the $z$ axis, the phase
1179: factor before $\Psi_f$ does not include $\eta$ and thus it does
1180: not depend on the combination
1181: $\vec{\hat{\theta}}-{\hat{l}}/{\hat{z}}-\vec{\hat{\eta}}$.}:
1182: 
1183: 
1184: \begin{eqnarray}
1185: \hat{E}_{f}\left(\hat{z},\vec{\hat{\eta}}, \vec{\hat{l}},
1186: \vec{\hat{\theta}}\right)&=& \frac{1}{\hat{z}} \exp\left[i\left(
1187: \frac{\hat{z} \hat{\theta}^2}{2}-
1188: \vec{\hat{\theta}}\cdot\vec{\hat{l}}
1189: +\frac{\hat{l}^2}{2\hat{z}^2}~\right)\right]
1190: \Psi_f\left(\left|\vec{\hat{\theta}}-\frac{\hat{l}}{\hat{z}}-\vec{\hat{\eta}}\right|\right)~
1191: , \label{undurad4}
1192: \end{eqnarray}
1193: %
1194: where
1195: 
1196: \begin{eqnarray}
1197: \Psi_f(\alpha) \equiv
1198: \mathrm{sinc}\left[\frac{\alpha^2}{4}\right]~ , \label{bisgg}
1199: \end{eqnarray}
1200: %
1201: subscript $f$ indicating the "far-zone". The field distribution of
1202: the virtual source positioned at $z=0$, corresponding to the waist
1203: of our laser-like beam was found to be:
1204: 
1205: \begin{eqnarray}
1206: \hat{{E}}_{0}\left(0,\vec{\hat{\eta}}, \vec{\hat{l}},
1207: \vec{\hat{r}}_{{}}\right) &=& - i \pi \exp\left[i \vec{\hat{\eta}}
1208: \cdot \left(\vec{\hat{r}}_{}-\vec{\hat{l}}~\right) \right]
1209: \Psi_0\left({\left|~\vec{\hat{r}}_{
1210: {}}-\vec{\hat{l}}~~\right|}\right)~, \label{undurad5gg}
1211: \end{eqnarray}
1212: %
1213: where
1214: 
1215: 
1216: \begin{eqnarray}
1217: \Psi_0(\alpha) \equiv \frac{1}{\pi}\left[\pi - 2\mathrm{Si}
1218: \left(\alpha^2\right)\right]~ . \label{undurad4bisgg0}
1219: \end{eqnarray}
1220: %
1221: where $\mathrm{Si}(z)=\int_0^z dt \sin(t)/t$ indicates the sin
1222: integral function and subscript "0" is indicative of the source
1223: position. Plots of $\Psi_f$ and $\Psi_0$ are given in Fig.
1224: \ref{psif} and Fig. \ref{psio}. It should be noted here that the
1225: independent variable in both plots is the dummy variable $\alpha$.
1226: The characteristic transverse range of the field in the far zone
1227: is in units of the radiation diffraction angle
1228: $\sqrt{\lambdabar/L_w}$, while the characteristic transverse range
1229: of the field at the source is in units of the radiation
1230: diffraction size $\sqrt{\lambdabar L_w}$.
1231: 
1232: \begin{figure}
1233: \begin{center}
1234: \includegraphics*[width=110mm]{figure7.eps}% Here is how to import EPS art
1235: \caption{\label{psif} Universal function $\Psi_f(\alpha)$, used to
1236: calculate the far-zone radiation field of a single electron at the
1237: fundamental harmonic at perfect resonance.}
1238: \end{center}
1239: \end{figure}
1240: %
1241: \begin{figure}
1242: \begin{center}
1243: \includegraphics*[width=110mm]{figure8.eps}% Here is how to import EPS art
1244: \caption{\label{psio} Universal function $\Psi_0(\alpha)$, used to
1245: calculate the radiation field of a single electron at the
1246: fundamental harmonic at perfect resonance on the virtual source
1247: plane.}
1248: \end{center}
1249: \end{figure}
1250: %
1251: Finally, with the help of the Fresnel propagation formula Eq.
1252: (\ref{fieldpropback}), we found the following expression for the
1253: field distribution at any distance $\hat{z} > 1/2$ from the
1254: virtual source:
1255: 
1256: \begin{eqnarray}
1257: \hat{E}\left(\hat{z},\vec{\hat{\eta}},
1258: \vec{\hat{l}},\vec{\hat{\theta}}_{}\right) &=& \exp\left[\frac{i
1259: \hat{z} }{2}\left|\vec{\hat{\theta}}_{}
1260: -\frac{\vec{\hat{l}}}{\hat{z}}~\right|^2\right]
1261: \Psi\left(\hat{z},\left|\vec{\hat{\theta}}_{}-
1262: \frac{\vec{\hat{l}}}{\hat{z}}-\vec{\hat{\eta}}\right|\right)~,
1263: \label{Esumg}
1264: \end{eqnarray}
1265: %
1266: where we defined
1267: 
1268: \begin{eqnarray}
1269: \Psi(\hat{z},\alpha) \equiv \exp\left[-i \frac{\hat{z}\alpha^2}{2}
1270: \right] \left\{\mathrm{Ei} \left[\frac{i \hat{z}^2\alpha^2
1271: }{2{\hat{z}} - 1}\right]- \mathrm{Ei} \left[\frac{i
1272: \hat{z}^2\alpha^2 }{2{\hat{z}}  + 1}\right]\right\}~,
1273: \label{undurad4bisggz}
1274: \end{eqnarray}
1275: %
1276: and $\mathrm{Ei}(z)=-\int_{-z}^{\infty} dt \exp(-t)/t$ indicates
1277: the exponential integral function. Eq. (\ref{Esumg}) is a
1278: particular case of Eq. (\ref{undunormfin}) at perfect resonance.
1279: Note that free space basically acts as a spatial Fourier
1280: transformation. This means that the field in the far zone is,
1281: aside for a phase factor, the Fourier transform of the field at
1282: any position $z$ down the beamline. It is also, aside for a phase
1283: factor, the spatial Fourier transform of the virtual source:
1284: 
1285: \begin{eqnarray}
1286: {\hat{E}}_0\left(0,\vec{\hat{\eta}},
1287: \vec{\hat{l}},\vec{\hat{r}}_{}\right)&=& -\frac{i  \hat{z}}{2 \pi
1288: } \int d\vec{\hat{\theta}}\exp{\left[-\frac{i  |\vec{
1289: \hat{\theta}}|^2}{2} \hat{z}\right]}{
1290: \hat{E}}_f\left(\hat{z},\vec{\hat{\eta}}, \vec{\hat{l}},
1291: \vec{\hat{\theta}}\right)\exp\left[{i } \vec{\hat{r}}_{{}}\cdot
1292: \vec{\hat{\theta}}\right] ~ . \label{virfiemody}
1293: \end{eqnarray}
1294: %
1295: It follows that
1296: 
1297: \begin{eqnarray}
1298: \int d \vec{{\theta}}~ \Psi_f(\theta) \exp{\left[{i  }
1299: {\vec{\hat{r}}}_{}\cdot\vec{{\theta}} \right]} &=& 2 \pi
1300: \int_0^{\infty} d \theta~ \theta J_o\left( {|\vec{\hat{r}}}_{}|
1301: \theta \right) \Psi_f(\theta)   =  2\pi^2
1302: \Psi_0\left(|\vec{\hat{r}}_{}|\right) ~.\label{psi0psif}
1303: \end{eqnarray}
1304: %
1305: We conclude verifying that Eq. (\ref{Esumg}) is in agreement with
1306: Eq. (\ref{undurad4}) and Eq. (\ref{undurad5gg}) $\hat{z} \gg 1$
1307: and for $\hat{z}=0$, respectively. Consider first $\hat{z}=0$. For
1308: positive numbers $\alpha^2 \equiv
1309: \left|\vec{\hat{r}}-\vec{\hat{l}}-\hat{z}\vec{\hat{\eta}}\right|^2
1310: > 0$, we have
1311: 
1312: \begin{eqnarray}
1313: -i [\pi +2 \mathrm{Si}(\alpha^2)] = \mathrm{Ei}(- i \alpha^2)
1314: -\mathrm{Ei}( i \alpha^2)~, \label{siei}
1315: \end{eqnarray}
1316: %
1317: and Eq. (\ref{Esumg})  yields back Eq. (\ref{undurad5gg}).
1318: 
1319: Consider now the limit for $\hat{z} \gg 1$. For positive numbers
1320: $\xi^2 \equiv \left|\vec{\hat{\theta}} -\vec{\hat{\eta}}\right|^2
1321: > 0$ one has
1322: 
1323: \begin{eqnarray}
1324: &&\exp\left[-\frac{i\hat{z}\xi^2}{2}\right]\left[
1325: \mathrm{Ei}\left(\frac{i \xi^2 \hat{z}^2}{2 \hat{z} -1}\right) -
1326: \mathrm{Ei}\left(\frac{i \xi^2 \hat{z}^2}{2 \hat{z}
1327: +1}\right)\right] \longrightarrow
1328: \frac{1}{\hat{z}}\mathrm{sinc}\left(\frac{\xi^2}{4}\right)~,\label{sincei}
1329: \end{eqnarray}
1330: %
1331: as it can be directly seen comparing Eq. (\ref{Esumg}) with Eq.
1332: (\ref{undunormfin})), where integration in Eq. (\ref{psig}) is
1333: performed directly at $\hat{C} = 0$ and in the limit for $\hat{z}
1334: \gg 1$. Thus, Eq. (\ref{Esumg}) yields back Eq. (\ref{undurad4}),
1335: as it must be.
1336: 
1337: Finally, it should be noted that expressions in the present
1338: Section \ref{sub:fiel} have been derived for $\omega > 0$.
1339: Expressions for the field at negative values of $\omega$ can be
1340: obtained based on the property $\vec{\bar{E}}(-\omega) =
1341: \vec{\bar{E}}^*(\omega)$ starting from explicit expressions at
1342: $\omega>0$.
1343: 
1344: 
1345: \subsection{\label{sub:evol} Cross-spectral density of an undulator source
1346: and its free-space propagation.}
1347: 
1348: 
1349: From now on we consider a normalized expression of the
1350: cross-spectral density, $\hat{G}$, that is linked with $G$ in Eq.
1351: (\ref{coore}) by a proportionality factor
1352: 
1353: \begin{equation}
1354: \hat{G} =  \left(\frac{2 c^2  \gamma}{K \omega e A_{JJ}}\right)^2
1355: G ~.\label{inetransf}
1356: \end{equation}
1357: %
1358: Moreover, we introduce variables
1359: 
1360: \begin{equation}
1361: \Delta \vec{{r}} = {\vec{\hat{r}}_{1}-\vec{\hat{r}}_{2}}~,~~
1362: \vec{\bar{r}} = \frac{\vec{\hat{r}}_{1}+\vec{\hat{r}}_{2}}{2}~.
1363: \label{deltabarx}
1364: \end{equation}
1365: %
1366: and
1367: 
1368: \begin{equation}
1369: \Delta \vec{\theta} =
1370: {\vec{\hat{\theta}}_{1}-\vec{\hat{\theta}}_{2}}~,~~
1371: \vec{\bar{\theta}} =
1372: \frac{\vec{\hat{\theta}}_{1}+\vec{\hat{\theta}}_{2}}{2}~,
1373: \label{deltabarth}
1374: \end{equation}
1375: %
1376: where, as before, $\vec{\hat{\theta}} = \vec{\hat{r}}/z$. Thus,
1377: Eq. (\ref{coore}) can now be written as
1378: 
1379: \begin{equation}
1380: \hat{G}(\hat{z},\vec{\bar{r}},\Delta \vec{r}, \hat{C}) \equiv
1381: \left\langle \hat{E} \left(\vec{\hat{\eta}},\vec{\hat{l}},\hat{z},
1382: \vec{\bar{r}}+ \frac{\Delta\vec{r}}{2}, \hat{C}\right)
1383: \hat{E}^*\left(\vec{\hat{\eta}},\vec{\hat{l}},\hat{z},
1384: \vec{\bar{r}}- \frac{\Delta\vec{r}}{2}, \hat{C}\right)
1385: \right\rangle~.\label{coore2}
1386: \end{equation}
1387: %
1388: On the one hand, the cross-spectral density as is defined in Eq.
1389: (\ref{coore2}) includes the product of fields which obey the free
1390: space propagation relation Eq. (\ref{fieldpropback}). On the other
1391: hand, the averaging over random variables commutes with all
1392: operations involved in the calculation of the field propagation.
1393: More explicitly, introducing the notation
1394: $\hat{E}(z)=\mathcal{O}[\hat{E}(z_s)]$ as a shortcut for Eq.
1395: (\ref{fieldpropback}) one can write
1396: 
1397: \begin{eqnarray}
1398: \hat{G}(z)  = \left\langle
1399: \mathcal{O}\left[\hat{E}(\hat{z}_s)\right]\mathcal{O}^*\left[\hat{E}^*(\hat{z}_s)\right]\right\rangle
1400: =
1401: \mathcal{O}\cdot\mathcal{O}^*\left[\left\langle\hat{E}(\hat{z}_s)\hat{E}^*(\hat{z}_s)\right\rangle\right]
1402: = \mathcal{O}\cdot\mathcal{O}^*\left[G(\hat{z}_s)\right] ~,
1403: \label{comm}
1404: \end{eqnarray}
1405: %
1406: where $\mathcal{O}$ may also represent, more in general, any
1407: linear operator. Once the cross-spectral density at the source is
1408: known, Eq. (\ref{comm}) provides an algorithm to calculate the
1409: cross-spectral density at any position $\hat{z}$ down the beamline
1410: (in the free-space case). Similarly, propagation through a complex
1411: optical system can be performed starting from the knowledge of
1412: $\hat{G}(\hat{z}_s)$. As a result, the main problem to solve in
1413: order to characterize the cross-spectral density at the specimen
1414: position is to calculate the cross-spectral density at the virtual
1415: source. For the undulator case, we fix the position of the source
1416: in the center of the undulator $\hat{z}_s=0$. This is the main
1417: issue this paper is devoted to. However, free-space propagation is
1418: also treated, and may be considered an illustration of how our
1419: main result can be used in a specific case.
1420: 
1421: Based on Eq. (\ref{coore2}) and on results in Section
1422: \ref{sub:fiel} we are now ready to present an expression for the
1423: cross-spectral density at any position down the beamline, always
1424: keeping in mind that the main result we are looking for is the
1425: cross-spectral density at the virtual source position.
1426: 
1427: We begin giving a closed expression for $\hat{G}$ valid at any
1428: value of the detuning parameter $\hat{C}$ by substituting Eq.
1429: (\ref{undunormfin}) in Eq. (\ref{coore2}), and replacing the
1430: ensemble average with integration over the transverse beam
1431: distribution function. We thus obtain
1432: 
1433: \begin{eqnarray}
1434: &&\hat{G}\left(\hat{z},\hat{C},\vec{\bar{\theta}},\Delta
1435: \vec{\theta}\right) = \exp\left[{i \hat{z} \vec{\bar{\theta}}\cdot
1436: \Delta\vec{\theta}} \right]\int d\vec{\hat{l}}
1437: \exp\left[-{i}\vec{\hat{l}}\cdot \Delta \vec{\theta}\right]\cr
1438: &&\times  \int d\vec{\hat{\eta}}
1439: f_{\bot}\left(\vec{\hat{l}},\vec{\hat{\eta}}\right)
1440: {\Psi}_{C}\left(\hat{z},\hat{C},\left|\vec{\bar{\theta}}+\frac{\Delta
1441: \vec{\theta}}{2}-\frac{\vec{\hat{l}}}{\hat{z}}-\vec{\hat{\eta}}\right|\right)
1442: {\Psi}^*_{C}\left(\hat{z},\hat{C},\left|\vec{\bar{\theta}}-\frac{\Delta
1443: \vec{\theta}}{2}-\frac{\vec{\hat{l}}}{\hat{z}}-\vec{\hat{\eta}}\right|\right)~.\cr
1444: && \label{Cany}
1445: \end{eqnarray}
1446: %
1447: It is often useful to substitute the integration variables
1448: $\vec{\hat{l}}$ with $\vec{\phi} \equiv
1449: -\vec{\bar{\theta}}+\vec{\hat{l}}/\hat{z}+\vec{\hat{\eta}}$ In
1450: fact, in this way, Eq. (\ref{Cany}) becomes
1451: 
1452: 
1453: \begin{eqnarray}
1454: &&\hat{G}\left(\hat{z},\hat{C},\vec{\bar{\theta}},\Delta
1455: \vec{\theta}\right) = \hat{z}^2 \exp\left[{i
1456: \hat{z}\vec{\bar{\theta}}\cdot \Delta\vec{\theta}} \right]\int
1457: d\vec{\phi} \int d\vec{\hat{\eta}} \exp\left[-{i \hat{z}}
1458: \left(\vec{\phi}+\vec{\bar{\theta}}-\vec{\hat{\eta}}\right)\cdot
1459: \Delta \vec{\theta}\right]\cr &&\times
1460: f_\bot\left(\vec{\phi}+\vec{\bar{\theta}}-\vec{\hat{\eta}},\vec{\hat{\eta}}\right)
1461: {\Psi}_{C}\left(\hat{z},\hat{C},\left|\vec{\phi}-\frac{\Delta
1462: \vec{\theta}}{2}\right|\right)
1463: {\Psi}^*_{C}\left(\hat{z},\hat{C},\left|\vec{\phi}+\frac{\Delta
1464: \vec{\theta}}{2}\right|\right)~, \label{Cany2}
1465: \end{eqnarray}
1466: %
1467: For choices of $f$ of particular interest (e.g. product of
1468: Gaussian functions for both transverse and angle distributions),
1469: integrals in $d\vec{\eta}$ can be performed analytically, leaving
1470: an expression involving two integrations only, and still quite
1471: generic.
1472: 
1473: Eq. (\ref{Cany2}) is as far as we can get with this level of
1474: generality, and can be exploited with the help of numerical
1475: integration techniques. However, it still depends on six
1476: parameters at least: four parameters\footnote{At least. This
1477: depends on the number of parameters needed to specify $f_\bot$.
1478: For a Gaussian distribution in phase space, four parameters are
1479: needed, specifying rms transverse size and angular divergence in
1480: the horizontal and vertical direction.} are needed to specify
1481: $f_\bot$, plus the detuning parameter $\hat{C}$ and the distance
1482: $\hat{z}$.
1483: 
1484: In the following we will assume $\hat{C} \ll 1$, that allows us to
1485: take advantage of analytical presentations for the single-particle
1486: field obtained in \cite{OURF} and reported before. This means that
1487: monochromatization is good enough to neglect finite bandwidth of
1488: the radiation around the fundamental frequency. By this, we
1489: automatically assume that monochromatization is performed around
1490: the fundamental frequency. It should be noted, however, that our
1491: theory can be applied to the case monochromatization is performed
1492: at other frequencies too. Analytical presentation of the
1493: single-particle field cannot be used in full generality, but for
1494: any fixed value $\hat{C}$ of interest one may tabulate the special
1495: function $\Psi_C$ once and for all, and use it in place of $\Psi$
1496: throughout the paper\footnote{Of course, selection of a particular
1497: value $\hat{C}$ still implies a narrow monochromator bandwidth
1498: around that value.}. From this viewpoint, although the case of
1499: prefect resonance studied here is of practical importance in many
1500: situations, it should be considered as a particular illustration
1501: of our theory only.
1502: 
1503: Also note that in Eq. (\ref{Cany}) the electron beam energy spread
1504: is assumed to be negligible. Contrarily to the monochromator
1505: bandwidth, the energy spread is fixed for a given facility: its
1506: presence constitutes a fundamental effect. In order to
1507: quantitatively account for it, one should sum the dimensionless
1508: energy-spread parameter $\hat{\Delta}_E = 4 \pi N_w
1509: \delta\gamma/\gamma$ to $\hat{C}$ in Eq. (\ref{Cany}) and,
1510: subsequently, integration should be extended over the
1511: energy-spread distribution. Typical energy spread $\delta
1512: \gamma/\gamma$ for third generation light sources is of order $0.1
1513: \%$. For ERL sources this figure is about an order of magnitude
1514: smaller, $\delta \gamma/\gamma \sim 0.01 \%$.
1515: 
1516: \begin{figure}
1517: \begin{center}
1518: \includegraphics*[width=140mm]{Espread.eps}% Here is how to import EPS art
1519: \caption{Study of $\hat{I}$ given in Eq. (\ref{enspread}) for
1520: different valued of $\hat{C}$ and $\hat{\Delta}_E$. Here $N_w
1521: =70$. Plot (a) : comparison, at $\hat{C}=0$, of the case for
1522: negligible energy spread with the case $\hat{\Delta}_E = 0.88$.
1523: Plot (b) : comparison, at negligible energy spread, of the case
1524: $\hat{C}=0$ with the case $\hat{C} = \pm 0.5$. Plot (c) :
1525: comparison of the case $\hat{C}=0$ at negligible energy spread
1526: with the case $\hat{C} = \pm 0.5$ at $\Delta \omega/\omega \simeq
1527: 0.1 \%$. \label{espread} }
1528: \end{center}
1529: \end{figure}
1530: %
1531: 
1532: In order to study the impact of a finite energy spread parameter
1533: $\hat{\Delta}_{E}$, of a finite radiation bandwidth and of a
1534: relatively small detuning from the fundamental we consider an
1535: expression for the intensity of a diffraction-limited beam
1536: $\hat{z}^2  \hat{I} = \hat{z}^2 \langle |\hat{E}|^2 \rangle$
1537: including both $\hat{\Delta}_{E}$ and $\hat{C}$:
1538: 
1539: \begin{eqnarray}
1540: \hat{z}^2 \hat{I}(\hat{\theta},\hat{C},\hat{\Delta}_E) &&=
1541: \frac{1}{\sqrt{2\pi}\hat{\Delta}_E} \int_{-\infty}^{\infty} d
1542: \hat{\xi}_E \exp\left[-\frac{\hat{\xi}_E^2}{2 \hat{\Delta}_E^2}
1543: \right] \mathrm{sinc}^2
1544: \left(\frac{\hat{C}+\hat{\xi}_E}{2}+\frac{\hat{\theta}^2}{4}\right)~.\cr
1545: && \label{enspread}
1546: \end{eqnarray}
1547: %
1548: We plotted $\hat{z}^2 \hat{I}$ for different values of $\hat{C}$
1549: and $\hat{\Delta}_E$ in Fig. \ref{espread}. First, in Fig.
1550: \ref{espread} (a), we compared, at $\hat{C}=0$, the case for
1551: negligible energy spread with the case $\hat{\Delta}_E = 0.88$,
1552: corresponding to $\delta \gamma/\gamma \simeq 0.1 \%$ at $N_w
1553: =70$, typical of third generation sources. As one can see, maximal
1554: intensity differences are within $10 \%$. Second, in Fig.
1555: \ref{espread} (b) we compared, at negligible energy spread, the
1556: case $\hat{C}=0$ with $\hat{C} = \pm 0.5$ at $N_w = 70$,
1557: corresponding to a shift $\Delta \omega/\omega \simeq 0.1 \%$. One
1558: can see that also in this case maximal intensities differ of about
1559: $10 \%$. Analysis of Fig. \ref{espread} (c), where the case
1560: $\hat{C}=0$ at negligible energy spread is compared with cases
1561: $\hat{C} = \pm 0.5$ at $\Delta \omega/\omega \simeq 0.1 \%$ and
1562: $N_w = 70$ leads to a similar result. This reasoning allows to
1563: conclude that our simplest analytical illustrations can be applied
1564: to practical cases of interest involving third generation sources
1565: and undulators with up to $70$ periods with good accuracy. It
1566: should be remarked that such illustration holds for the first
1567: harmonic only. In fact, while the shape of $\hat{z}^2 \hat{I}$ is
1568: still given by Eq. (\ref{enspread}) in the case of odd harmonic of
1569: order $h$, parameters $\hat{C}$ and $\hat{\Delta}_E$ are modified
1570: according to $\hat{C}_h = h \hat{C}$ and $\hat{\Delta}_{E~h}=h
1571: \hat{\Delta}_{E}$, decreasing the applicability of our analytical
1572: results.
1573: 
1574: %When $\delta\gamma/\gamma \sim 0.01 \%$ the asymptotic case for a
1575: %monoenergetic beam, Eq. (\ref{Cany}), can be used with high
1576: %accuracy. For $\delta\gamma/\gamma \sim 0.1 \%$ Eq. (\ref{Cany})
1577: %can be used as a first estimation for the cross-spectral density.
1578: 
1579: With this in mind, we can present an expression for the
1580: cross-spectral density at $\hat{C} \ll 1$ based on Eq.
1581: (\ref{coore2}) and Eq. (\ref{Esumg}). Substituting the latter in
1582: the former we obtain an equation for
1583: $\hat{G}(\hat{z},\vec{\bar{\theta}},\Delta \vec{\theta})$ that can
1584: be formally derived from Eq. (\ref{Cany}) by substitution of
1585: $\Psi_C$ with $\Psi$. Similarly as before, one may give
1586: alternative presentation of
1587: $\hat{G}(\hat{z},\vec{\bar{\theta}},\Delta \vec{\theta})$
1588: replacing the integration variables $\vec{\hat{l}}$ with
1589: $\vec{\phi} \equiv
1590: -\vec{\bar{\theta}}+\vec{\hat{l}}/\hat{z}+\vec{\hat{\eta}}$. This
1591: results in another expression for
1592: $\hat{G}(\hat{z},\vec{\bar{\theta}},\Delta \vec{\theta})$ that can
1593: be formally derived from Eq. (\ref{Cany2}) by substituting
1594: $\Psi_C$ with $\Psi$. This last expression presents the
1595: cross-spectral density in terms of a convolution of the transverse
1596: electron beam phase space distribution with an analytical
1597: function, followed by Fourier transformation\footnote{Aside for an
1598: inessential multiplicative constant. This remark also applies in
1599: what follows.}.
1600: 
1601: One may obtain an expression for $\hat{G}$ at $\hat{z}=0$ as a
1602: limiting case of Eq. (\ref{Cany}) or Eq. (\ref{Cany2}) at
1603: $\hat{C}=0$. It is however simpler to do so by substituting Eq.
1604: (\ref{undurad5gg}) in Eq. (\ref{coore2}) that gives
1605: 
1606: \begin{eqnarray}
1607: \hat{G}\left(0,\vec{\bar{r}},\Delta \vec{r}\right) &=& \int
1608: d\vec{\eta} \exp\left[i \vec{{\eta}} \cdot \Delta \vec{r}\right]
1609: \cr && \times \int d\vec{l} f_\bot\left(\vec{l},\vec{\eta}\right)
1610: \pi^2 \Psi_0\left(\left|\vec{\bar{r}}-\frac{\Delta \vec{r}}{2} -
1611: \vec{l}\right|\right) \Psi_0\left(\left|\vec{\bar{r}}+\frac{\Delta
1612: \vec{r}}{2} - \vec{l}\right|\right)~,\label{Gnor}
1613: \end{eqnarray}
1614: %
1615: where the function $\Psi_0$ has already been defined in Eq.
1616: (\ref{undurad4bisgg0}). The product
1617: $\Psi_0\left(\left|\vec{\bar{r}}-{\Delta \vec{r}}/{2}
1618: \right|\right)
1619:  \Psi_0\left(\left|\vec{\bar{r}}+{\Delta \vec{r}}/{2}\right| \right)$ is a
1620: four-dimensional , analytical function in $\vec{\bar{r}}$ and
1621: $\Delta \vec{r}$. Eq. (\ref{Gnor}) tells that the cross-spectral
1622: density at the virtual source position can be obtained convolving
1623: $\Psi_0\left(\left|\vec{\bar{r}}-{\Delta \vec{r}}/{2}
1624: \right|\right) \Psi_0\left(\left|\vec{\bar{r}}+{\Delta
1625: \vec{r}}/{2} \right|\right)$ with the transverse beam distribution
1626: at $z=0$, $f_\bot\left(\vec{\hat{l}},\vec{\hat{\eta}}\right)$,
1627: considered as a function of $\vec{\hat{l}}$, and taking Fourier
1628: transform with respect to $\vec{\hat{\eta}}$. When the betatron
1629: functions have minima in the center of the undulator we have
1630: 
1631: \begin{eqnarray}
1632: f_\bot\left(\vec{\hat{l}},\vec{\hat{\eta}}\right) =
1633: f_l\left(\vec{\hat{l}}\right)
1634: f_\eta\left(\vec{\hat{\eta}}\right)~. \label{ff}
1635: \end{eqnarray}
1636: %
1637: Then, Eq. (\ref{Gnor}) becomes
1638: 
1639: 
1640: \begin{eqnarray}
1641: \hat{G}\left(0,\vec{\bar{r}},\Delta \vec{r}\right) &=& \int
1642: d\vec{\hat{\eta}} f_\eta\left(\vec{\hat{\eta}}\right) \exp\left[i
1643: \vec{\hat{\eta}} \cdot \Delta \vec{r}\right] \cr && \times \int
1644: d\vec{\hat{l}} f_l\left(\vec{\hat{l}}\right) \pi^2
1645: \Psi_0\left(\left|\vec{\bar{r}}-\frac{\Delta \vec{r}}{2} -
1646: \vec{\hat{l}}\right|\right)
1647: \Psi_0\left(\left|\vec{\bar{r}}+\frac{\Delta \vec{r}}{2} -
1648: \vec{\hat{l}}\right|\right)~,\label{Gnor2}
1649: \end{eqnarray}
1650: %
1651: that will be useful later on. In this case, the cross-spectral
1652: density is the product of two separate factors. First, the Fourier
1653: transform of the distribution of angular divergence of electrons.
1654: Second, the convolution of the transverse electron beam
1655: distribution with the four-dimensional function
1656: $\Psi_0\left(\left|\vec{\bar{r}}-{\Delta \vec{r}}/{2}
1657: \right|\right) \Psi_0\left(\left|\vec{\bar{r}}+{\Delta
1658: \vec{r}}/{2} \right|\right)$.
1659: 
1660: Eq. (\ref{Cany}) (or Eq. (\ref{Cany2})) constitutes the most
1661: general result in in the calculation of the cross-spectral density
1662: for undulator sources. Its applicability is not restricted to
1663: third generation light sources. In particular, it can be used for
1664: arbitrary undulator sources like ERLs \cite{EDGA} or XFEL
1665: spontaneous undulators \cite{XFEL}. Eq. (\ref{Cany2}) has been
1666: derived, in fact, under the only constraints $\gamma^2 \gg 1$,
1667: $N_w \gg 1$ and $\sigma_T \omega_r \gg N_w$. Note that $\sigma_T
1668: \sim 30 $ ps for a typical SR source, whereas $\sigma_T \sim 100$
1669: fs for an XFEL spontaneous undulator source or an ERL. Yet, for
1670: all practical cases of interest, $\sigma_T \omega \gg N_w$. As we
1671: have seen before, Eq. (\ref{Cany2}) further simplifies  in the
1672: particular but practical case of perfect resonance, i.e. in the
1673: limit for $\hat{C} \ll 1$. A particularly important asymptote of
1674: Eq. (\ref{Cany}) at perfect resonance is at the virtual source
1675: position, described by Eq. (\ref{Gnor}), which express the
1676: cross-spectral density in the undulator center. While Eq.
1677: (\ref{Cany}) (or Eq. (\ref{Cany2})) solves all problems concerning
1678: characterization of transverse coherence properties of light in
1679: free-space, the knowledge of Eq. (\ref{Gnor}) constitutes, in the
1680: presence of optical elements, the first (and main) step towards
1681: the characterization of SR light properties at the specimen
1682: position. In fact, the tracking of the cross-spectral density can
1683: be performed with the help of standard statistical optics
1684: formalism developed for the solution of problems dealing with
1685: partially coherent sources. Finally, it should be noted that in
1686: the case of XFELs and ERLs, there is no further simplification
1687: that we may apply to previously found equations. In particular,
1688: the transverse electron beam phase space should be considered as
1689: the result of a start-to-end simulation or, better, of
1690: experimental diagnostics measurements in a operating machine. On
1691: the contrary, as we will see, extra-simplifications can be
1692: exploited in the case of third-generation light sources, allowing
1693: for the development of a more specialized theory.
1694: 
1695: Inspection of  Eq. (\ref{Cany}) or Eq. (\ref{Gnor}) results in the
1696: conclusion that a Gaussian-Schell model cannot be applied to
1697: describe partially coherent SR light. In fact, functions $\Psi_C$,
1698: $\Psi$, $\Psi_0$ and $\Psi_f$ are of non-Gaussian nature, as the
1699: laser-like beam they can be ascribed to is non-Gaussian. This
1700: explains our words in the Introduction, where we stated that
1701: \cite{COI1,COI2,COI3} are of general theoretical interest, but
1702: they do not provide a satisfactory approximation to third
1703: generation SR sources.
1704: 
1705: As a final remark to this Section, we should discuss the relation
1706: of our approach with that given, in terms of Wigner distribution,
1707: in \cite{KIM2,KIM3}. As said in Section \ref{sec:intro}, treatment
1708: based on Wigner distribution is equivalent to treatment based on
1709: cross-spectral density. We chose to use cross-spectral density
1710: because such quantity is straightforwardly physically measurable,
1711: being related to the outcome of a Young's experiment. Essentially,
1712: one can obtain a Wigner distribution $\hat{W}$ from $\hat{G}$ by
1713: means of an inverse Fourier transformation:
1714: 
1715: \begin{eqnarray}
1716: \hat{W} \left(\hat{z},\vec{\bar{r}}, \vec{\bar{u}}\right) =
1717: \frac{1}{4\pi^2} \int d(\Delta\vec{r}~) \exp\left[-i\Delta
1718: \vec{r}\cdot \vec{\bar{u}}
1719: \right]\hat{G}\left(\hat{z},\vec{\bar{r}},\Delta \vec{r}\right) ~.
1720: \label{W1}
1721: \end{eqnarray}
1722: %
1723: Thus, Eq. (\ref{Gnor2}) gives
1724: 
1725: \begin{eqnarray}
1726: \hat{W}\left(0,\vec{\bar{r}},\vec{\bar{u}}\right) &=& \int
1727: d\vec{\hat{l}} \int d\vec{\hat{\eta}}
1728: f\left(\vec{\hat{l}},\vec{\hat{\eta}}\right)
1729: \hat{W}_o\left(0,\vec{\bar{r}}-\vec{\hat{l}},\vec{\bar{u}}-\vec{\hat{\eta}}\right)~.
1730: \label{W2}
1731: \end{eqnarray}
1732: %
1733: where
1734: 
1735: \begin{eqnarray}
1736: \hat{W}_o\left(0,\vec{\alpha},\vec{\delta}\right) = \frac{1}{4}
1737: \int d(\Delta \vec{r}~) \exp\left[-i\Delta \vec{r}\cdot
1738: \vec{\delta} \right] \Psi_0\left(\left|\vec{\alpha}-\frac{\Delta
1739: \vec{r}}{2}\right|\right)
1740: \Psi_0\left(\left|\vec{\alpha}+\frac{\Delta
1741: \vec{r}}{2}\right|\right)~,\label{WWW}
1742: \end{eqnarray}
1743: %
1744: The Wigner distribution at $\hat{z}=0$ is presented as a
1745: convolution product between the electron phase-space and a
1746: universal function $\hat{W}_o$. This result may be directly
1747: compared (aside for different notation) with \cite{KIM2,KIM3},
1748: where the Wigner distribution is presented as a convolution
1749: between the electrons phase space and a universal function as
1750: well. The study in \cite{KIM2,KIM3} ends at this point, presenting
1751: expressions for arbitrary detuning parameter. On the contrary, in
1752: the following Sections we will take advantage of expressions at
1753: perfect resonance, of small and large parameters related to third
1754: generation light sources and of specific characteristics of the
1755: electron beam distribution. This will allow us to develop a
1756: comprehensive theory of third generation SR sources.
1757: 
1758: 
1759: \section{\label{sec:main} Theory of transverse coherence for
1760: third-generation light sources}
1761: 
1762: \subsection{\label{sub:cross} Cross-Spectral Density}
1763: 
1764: 
1765: We now specialize our discussion to third-generation light
1766: sources.  We assume that the motion of particles in the horizontal
1767: and vertical directions are completely uncoupled. Additionally, we
1768: assume a Gaussian distribution of the electron beam in the phase
1769: space. These two assumptions are practically realized, with good
1770: accuracy, in storage rings. For simplicity, we also assume that
1771: the minimal values of the beta-functions in horizontal and
1772: vertical directions are located at the virtual source position
1773: $\hat{z} = 0$, that is often (but not
1774: always\footnote{Generalization to the case when this assumption
1775: fails is straightforward.}) the case in practice. $f_\bot =
1776: f_{\eta_x}(\hat{\eta}_x) f_{\eta_x}(\hat{\eta}_y)
1777: f_{l_x}(\hat{l}_x) f_{l_x}(\hat{l}_y)$ with
1778: 
1779: \begin{eqnarray}
1780: && f_{\eta_x}(\hat{\eta}_x) = \frac{1}{\sqrt{2\pi D_x}}
1781: \exp{\left(-\frac{\hat{\eta}_x^2}{2 D_x}\right)}~,~~~~
1782: f_{\eta_y}(\hat{\eta}_y)  = \frac{1}{\sqrt{2\pi D_y}}
1783: \exp{\left(-\frac{\hat{\eta}_y^2}{2 D_y}\right)}~,\cr &&
1784: f_{l_x}(~\hat{l}_x) =\frac{1}{\sqrt{2\pi N_x} }
1785: \exp{\left(-\frac{\hat{l}_x^2}{2 N_x}\right)}~,~~~~~
1786: f_{l_y}(~\hat{l}_y)=\frac{1}{\sqrt{2\pi N_y} }
1787: \exp{\left(-\frac{\hat{l}_y^2}{2 N_y}\right)}~.\label{distr}
1788: \end{eqnarray}
1789: %
1790: Here
1791: 
1792: \begin{equation}
1793: D_{x,y} = \frac{\sigma_{x',y'}^2} {{\lambdabar}/L_w}~,~~~~~
1794: N_{x,y} = \frac{\sigma^2_{x,y}} {\lambdabar L_w}~,\label{enne}
1795: \end{equation}
1796: %
1797: $\sigma_{x,y}$ and $\sigma_{x',y'}$ being rms transverse bunch
1798: dimensions and angular spreads. Parameters $N_{x,y}$ will be
1799: indicated as the beam diffraction parameters, are analogous to
1800: Fresnel numbers and correspond to the normalized square of the
1801: electron beam sizes, whereas $D_{x,y}$ represent the normalized
1802: square of the electron beam divergences.  Consider the reduced
1803: emittances $\hat{\epsilon}_{x,y} = \epsilon_{x,y}/\lambdabar$,
1804: where $\epsilon_{x,y}$ indicate the geometrical emittance of the
1805: electron beam in the horizontal and vertical directions. Since we
1806: restricted our model to third generation light sources, we can
1807: consider $\hat{\epsilon}_x\gg 1$. Moreover, since betatron
1808: functions are of order of the undulator length, we can also
1809: separately accept
1810: 
1811: \begin{eqnarray}
1812: N_x \gg 1~,~~~~~D_x \gg 1, \label{major}
1813: \end{eqnarray}
1814: %
1815: still retaining full generality concerning values of $N_y$ and
1816: $D_y$, due to the small coupling coefficient between horizontal
1817: and vertical emittance.
1818: 
1819: Exploitation of the extra-parameter $\hat{\epsilon}_x\gg 1$ (or
1820: equivalently $N_x\gg 1$ and $D_x \gg 1$) specializes our theory to
1821: the case of third-generation sources.
1822: 
1823: With this in mind we start to specialize our theory beginning with
1824: the expression for the cross-spectral density at the virtual
1825: source, i.e. Eq. (\ref{Gnor2}). After the change of variables
1826: $\vec{\phi} \longrightarrow -\vec{\bar{r}}+\vec{\hat{l}}$, and
1827: making use of Eq. (\ref{distr}), Eq. (\ref{Gnor2}) becomes
1828: 
1829: \begin{eqnarray}
1830: \hat{G}\left(0,\vec{\bar{r}},\Delta \vec{r}\right) &=&
1831: \frac{\pi}{2 \sqrt{N_x N_y}} \exp \left[-\frac{(\Delta x)^2
1832: D_x}{2}\right] \exp \left[-\frac{(\Delta y)^2 D_y}{2}\right]\cr &&
1833: \times \int_{-\infty}^{\infty} d \phi_x \int_{-\infty}^{\infty} d
1834: \phi_y \exp\left[-\frac{\left(\phi_x+\bar{x}\right)^2}{2
1835: N_x}\right] \exp\left[-\frac{\left(\phi_y+\bar{y}\right)^2}{2
1836: N_y}\right] \cr && \times
1837: \Psi_0\left\{\left[\left({\phi_{x}}+\frac{\Delta
1838: x}{2}\right)^2+\left({\phi_{y}}+\frac{\Delta y}{2} \right)^2
1839: \right]^{1/2}\right\} \cr && \times
1840: \Psi_0\left\{\left[\left({\phi_{x}}-\frac{\Delta x}{2}
1841: \right)^2+\left({\phi_{y}}-\frac{\Delta y}{2} \right)^2
1842: \right]^{1/2}\right\}~,\label{Gnor3}
1843: \end{eqnarray}
1844: %
1845: where $\Delta x$ and $\Delta y$ indicate components of $\Delta
1846: \vec{r}$. In Eq. (\ref{Gnor3}) the range of variable $\phi_x$  is
1847: effectively limited up to values $|\phi_x| \sim 1$. In fact,
1848: $\phi_x$ enters the expression for $\Psi_0$. It follows that at
1849: values larger than unity the integrand in Eq. (\ref{Gnor3}) is
1850: suppressed. Then, since $N_x \gg 1$, we can neglect $\phi_x$ in
1851: the exponential function. Moreover $D_x \gg 1$ and from the
1852: exponential function in $D_x$ follows that $\Delta x \ll 1$ can be
1853: neglected in $\Psi_0$. As a result, Eq. (\ref{Gnor3}) is
1854: factorized in the product of a horizontal cross-spectral density
1855: $G_x(0,\bar{x}, \Delta x)$ and a vertical cross-spectral density
1856: $G_y(0,\bar{x}, \Delta x)$:
1857: 
1858: \begin{eqnarray}
1859: \hat{G}\left(0,\vec{\bar{r}},\Delta \vec{r}\right) =
1860: \hat{G}_x\left(0,{\bar{x}},\Delta {x}\right)
1861: \hat{G}_y\left(0,{\bar{y}},\Delta {y}\right)~, \label{factorizeee}
1862: \end{eqnarray}
1863: %
1864: where
1865: 
1866: \begin{eqnarray}
1867: \hat{G}_x\left(0,{\bar{x}},\Delta {x}\right) &=&
1868: \sqrt{\frac{\pi}{{N_x}}} \exp \left[-\frac{(\Delta x)^2D_x
1869: }{2}\right] \exp\left[-\frac{\bar{x}^2}{2 N_x}\right]
1870: ~,\label{Gnor3x}
1871: \end{eqnarray}
1872: %
1873: \begin{eqnarray}
1874: \hat{G}_y\left(0,{\bar{y}},\Delta {y}\right) &=&
1875: \frac{1}{2}\sqrt{\frac{\pi}{N_y}}  \exp \left[-\frac{(\Delta
1876: y)^2D_y}{2}\right] \int_{-\infty}^{\infty} d \phi_y
1877: \int_{-\infty}^{\infty} d \phi_x
1878: \exp\left[-\frac{\left(\phi_y+\bar{y}\right)^2}{2 N_y}\right] \cr
1879: && \times
1880: \Psi_0\left\{\left[{\phi_{x}}^2+\left({\phi_{y}}+\frac{\Delta
1881: y}{2} \right)^2 \right]^{1/2}\right\}
1882: \Psi_0\left\{\left[{\phi_{x}}^2+\left({\phi_{y}}-\frac{\Delta
1883: y}{2} \right)^2 \right]^{1/2}\right\}~.\cr &&\label{Gnor3y}
1884: \end{eqnarray}
1885: %
1886: Note that, in virtue of Eq. (\ref{fieldpropback}), factorization
1887: holds in general, at any position $\hat{z}$. This allows us to
1888: separately study $\hat{G}_x$ and $\hat{G}_y$. $\hat{G}_x$
1889: describes a quasi-homogeneous Gaussian source, which will be
1890: treated in Section \ref{sub:gaus}. Here we will focus our
1891: attention on $\hat{G}_y$ only. It should be remarked that the
1892: quasi-homogenous Gaussian source asymptote is obtained from Eq.
1893: (\ref{Gnor3y}) in the limit $N_y \gg 1$ and $D_y \gg 1$. In other
1894: words, normalization constants in Eq. (\ref{Gnor3x}) and Eq.
1895: (\ref{Gnor3y}) are chosen in such a way that Eq. (\ref{Gnor3y})
1896: reduces to Eq. (\ref{Gnor3x}) in the limit $N_y \gg 1$ and $D_y
1897: \gg 1$ (with the obvious substitution $x \longrightarrow y$). It
1898: should be clear that this normalization is most natural, but not
1899: unique. The only physical constraint that normalization of Eq.
1900: (\ref{Gnor3x}) and Eq. (\ref{Gnor3y}) should obey is that the
1901: product $\hat{G}_x \hat{G}_y$ should not change.
1902: 
1903: Let us define the two-dimensional universal function
1904: $\mathcal{S}(\alpha,\delta)$ as\footnote{$\mathcal{S}$ stands for
1905: "Source".}
1906: 
1907: \begin{eqnarray}
1908: \mathcal{S}(\alpha,\delta) = \mathcal{K}_S \int_{-\infty}^{\infty}
1909: d \phi_x
1910: \Psi_0\left\{\left[{\phi_{x}}^2+\left(\alpha+\frac{\delta}{2}\right)^2
1911: \right]^{1/2}\right\}
1912: \Psi_0\left\{\left[{\phi_{x}}^2+\left(\alpha-
1913: \frac{\delta}{2}\right)^2 \right]^{1/2}\right\}~. \cr &&
1914: \label{mtcm}
1915: \end{eqnarray}
1916: %
1917: The normalization constant $\mathcal{K}_S \simeq 0.714$ is chosen
1918: in such a way that $\mathcal{S}(0,0)=1$. We can present
1919: $\hat{G}_y$ at the virtual source with the help of $\mathcal{S}$
1920: as
1921: 
1922: \begin{figure}
1923: \begin{center}
1924: \includegraphics*[width=140mm]{figure9.eps}% Here is how to import EPS art
1925: \caption{\label{FTM} Plot of the universal function
1926: ${\mathcal{S}}$, used to calculate the cross-spectral density at
1927: the virtual source when $N_x \gg1$, $D_x \gg 1$, $N_y$ and $D_y$
1928: are arbitrary. }
1929: \end{center}
1930: \end{figure}
1931: %
1932: 
1933: \begin{eqnarray}
1934: \hat{G}_y\left(0,{\bar{y}},\Delta {y}\right) &=&
1935: \frac{1}{2}\sqrt{\frac{\pi}{N_y}}  \exp \left[-\frac{(\Delta y)^2
1936: D_y }{2}\right] \int_{-\infty}^{\infty} d \phi_y
1937: \exp\left[-\frac{\left(\phi_y+\bar{y}\right)^2}{2 N_y}\right] \cr
1938: && \times~\frac{1}{\mathcal{K}_S} \mathcal{S}(\phi_y,\Delta y) ~.
1939: \label{Gnor4y}
1940: \end{eqnarray}
1941: %
1942: Therefore, $\hat{G}_y$ at the virtual source is found by
1943: convolving an universal function, $\mathcal{S}$, with a Gaussian
1944: function and multiplying the result by another Gaussian function.
1945: 
1946: Note that in the limit $N_x \gg 1$ and $D_x \gg 1$ there is no
1947: influence of the electron beam distribution along the vertical
1948: direction on $\hat{G}_x$. In spite of this, in the same limit, Eq.
1949: (\ref{Gnor4y}) shows that there is an influence of the horizontal
1950: electron beam distribution on $\hat{G}_y$, due to the
1951: non-separability of the function $\Psi_0$ in $\mathcal{S}$. In
1952: fact, contrarily to the case of a Gaussian laser beam,
1953: $\Psi_0[{\phi_{x}}^2+(\alpha- {\delta}/{2})^2 ] \ne
1954: \Psi_0[{\phi_{x}}^2 ] \Psi_0[(\alpha- {\delta}/{2})^2 ]$, and the
1955: integral in $d \phi_x$, that is a remainder of the integration
1956: along the x-direction, is still present in the definition of
1957: $\mathcal{S}$. However, such influence of the horizontal electron
1958: beam distribution is independent of $N_x$ and $D_x$. As a
1959: consequence, $\mathcal{S}$ is a universal function.
1960: 
1961: A plot of ${\mathcal{S}}$ is presented in Fig. \ref{FTM}.
1962: ${\mathcal{S}}$ is a real function. Then, $\hat{G}$ at the virtual
1963: plane is also real. Moreover, ${\mathcal{S}}$ is invariant for
1964: exchange of $\alpha$ with $\delta/2$.
1965: 
1966: 
1967: \begin{figure}
1968: \begin{center}
1969: \includegraphics*[width=140mm]{figure10.eps}% Here is how to import EPS art
1970: \caption{\label{M} Plot of the universal function $\mathcal{F}$,
1971: used to calculate the cross-spectral density in the far zone when
1972: $N_x \gg1$, $D_x \gg 1$, $N_y$ and $D_y$ are arbitrary. }
1973: \end{center}
1974: \end{figure}
1975: %
1976: 
1977: 
1978: Let us now deal with the evolution of the cross-spectral density
1979: $\hat{G}_y$ in free-space. In principle, one may use Eq.
1980: (\ref{Gnor4y}) and apply Eq. (\ref{comm}), remembering Eq.
1981: (\ref{fieldpropback}). It is however straightforward to use
1982: directly Eq. (\ref{Cany2}) at $\hat{C} \ll 1$. Under the
1983: assumption $N_x \gg 1$ and $D_x \gg 1$, as has been already
1984: remarked, factorization of $\hat{G}$ as a product of $\hat{G}_x$
1985: and $\hat{G}_y$ holds for any value of $\hat{z}$.  Isolating these
1986: factors in Eq. (\ref{Cany2}) at $\hat{C} \ll 1$ and using Eq.
1987: (\ref{distr}) one obtains
1988: 
1989: 
1990: \begin{eqnarray}
1991: &&\hat{G}_y\left(\hat{z},{\bar{\theta}_y},\Delta {\theta}_y\right)
1992: = \frac{\hat{z}^2}{2 \pi^2 \sqrt{2 N_y D_y}} \exp\left[{i
1993: \hat{z}\bar{\theta}_y \Delta{\theta}_y } \right]\int d{\phi}_y
1994: \exp\left[-{i \hat{z}}\left({\phi}_y+{\bar{\theta}_y}\right)
1995: \Delta {\theta}_y\right] \cr && \times \int d\hat{\eta}_y
1996: \exp\left[i \hat{z}{\hat{\eta}_y}\Delta {\theta}_y\right]
1997: \exp\left[-\frac{\hat{\eta}_y^2}{2 D_y}\right]
1998: \exp\left[-\frac{\left({\phi}_y
1999: +\bar{\theta}_y-\hat{\eta}_y\right)^2}{2 N_y/\hat{z}^2} \right]\cr
2000: && \times \int d \phi_x
2001: {\Psi}\left\{\hat{z},\left[{\phi}_x^2+\left(\phi_y-\frac{\Delta
2002: \theta_y }{2}\right)^2\right]^{1/2}\right\}
2003: {\Psi}^*\left\{\hat{z},\left[{\phi}_x^2+\left(\phi_y+\frac{\Delta
2004: \theta_y}{2}\right)^2\right]^{1/2}\right\}~. \label{Gnoryany}
2005: \end{eqnarray}
2006: %
2007: The integral in $d \hat{\eta}_y$ (second line) can be calculated
2008: analytically yielding
2009: 
2010: 
2011: \begin{eqnarray}
2012: &&\hat{G}_y\left(\hat{z},{\bar{\theta}}_y,\Delta \theta_y\right) =
2013: \frac{\hat{z}}{2\pi\sqrt{ \pi} \sqrt{N_y/\hat{z}^2 + D_y}}
2014: \exp\left[{\hat{z}}{i \bar{\theta}_y \Delta\theta_{y} } \right]
2015: \exp\left[-\frac{D_y N_y \Delta \theta_y^2 }{2\left(N_y/\hat{z}^2
2016: + D_y \right)}\right] \cr && \times \int d{\phi}_y
2017: \exp\left[-\frac{i \left(\phi_y+\bar{\theta}_y\right) \Delta
2018: \theta_y N_y/\hat{z}}{N_y/\hat{z}^2 + D_y }\right]
2019: \exp\left[-\frac{\left({\phi}_y+{\bar{\theta}_y}\right)^2}{2\left(N_y/\hat{z}^2
2020: + D_y \right)}\right] \cr && \times \int d \phi_x
2021: {\Psi}\left\{\hat{z},\left[{\phi}_x^2+\left(\phi_y-\frac{\Delta
2022: \theta_y }{2}\right)^2\right]^{1/2}\right\}
2023: {\Psi}^*\left\{\hat{z},\left[{\phi}_x^2+\left(\phi_y+\frac{\Delta
2024: \theta_y}{2}\right)^2\right]^{1/2}\right\}~. \label{Gnoryany2}
2025: \end{eqnarray}
2026: %
2027: We are now interested in discussing the far-zone limit of Eq.
2028: (\ref{Gnoryany2}). Up to now we dealt with the far-zone region of
2029: the field from a single particle, Eq. (\ref{undurad4}). In this
2030: case, the field exhibits a spherical wavefront. Such wavefront
2031: corresponds to the quadratic phase factor in Eq. (\ref{undurad4}).
2032: Note that when the electron is moving on-axis, Eq.
2033: (\ref{undurad4}) consists of the product of $\hat{z}^{-1} \exp[i
2034: \hat{z} \hat{\theta}^2/2]$ by a real function independent of
2035: $\hat{z}$. Such field structure can be taken as a definition of
2036: far-zone. A similar definition can be used for the far-zone
2037: pertaining the cross-spectral density. We regard the quadratic
2038: phase factor $\exp[{\hat{z}}{i \bar{\theta}_y \Delta\theta_{y} }
2039: ]$ in Eq. (\ref{Gnoryany2}) as the equivalent, in terms of
2040: cross-spectral density, of the quadratic phase factor for the
2041: single-particle field. We therefore take as definition of far-zone
2042: the region of parameters where $\hat{G}_y(\hat{z},
2043: \bar{\theta}_y,\Delta \theta_y) = \hat{z}^{-1}
2044: h(\bar{\theta}_y,\Delta \theta_y) \exp[{\hat{z}}{i \bar{\theta}_y
2045: \Delta\theta_{y} }]$, $h$ being a real function, and remains like
2046: that for larger values of $\hat{z}$.
2047: 
2048: Let us discuss the definition of the far-zone region in terms of
2049: problem parameters. In the single-particle situation, the only
2050: parameter of the problem was $\hat{z}$ and, as is intuitively
2051: sound, the far-zone region was shown to coincide with the limit
2052: $\hat{z} \gg 1$. In the case of Eq. (\ref{Gnoryany2}), we deal
2053: with three parameters $\hat{z}$, ${N}_y$ and ${D}_y$. Therefore we
2054: should find that the far-zone is defined in terms of conditions
2055: involving all three parameters.
2056: 
2057: When $N_y\lesssim 1$ and $D_y \lesssim 1$, analysis of Eq.
2058: (\ref{Gnoryany2}) shows that the far-zone region is for $\hat{z}
2059: \gg 1$. In this case the definition of far-zone for $\hat{G}_y$
2060: coincides with that of far-zone for the field of a single
2061: particle.
2062: 
2063: However, when either or both $N_y \gg 1$ or $D_y \gg 1$ the
2064: situation is different, and one finds that the far-zone condition
2065: is a combination of $\hat{z}$, $N_y$, and $D_y$. In all these
2066: cases, detailed mathematical analysis of Eq. (\ref{Gnoryany2})
2067: shows that far-zone is reached when
2068: 
2069: \begin{eqnarray}
2070: \max[N_y,1] \ll \hat{z}^2 \max[D_y,1] ~.\label{farzo}
2071: \end{eqnarray}
2072: %
2073: As it will be clearer after reading Section \ref{sub:geop}, but
2074: can also be seen considering the definition of our dimensionless
2075: units, the physical meaning of comparisons of $N_y$ and $D_y$ with
2076: unity in condition (\ref{farzo}) is that of a comparison between
2077: diffraction-related parameters (diffraction angle and diffraction
2078: size) and beam-related parameters (divergence and size of the
2079: electron beam).
2080: 
2081: When ${N}_y \gg 1$, but $D_y \lesssim 1$ condition (\ref{farzo})
2082: reads $\hat{z}^2 \gg N_y \gg 1$. This result is in agreement with
2083: intuition. The far-zone condition for $\hat{G}_y$ does not
2084: coincide with that for the field of a single electron, but it is
2085: anyway reached far away from the source, at $\hat{z} \gg 1$. As we
2086: will see in Section \ref{sub:nong}, the case ${N}_y \gg 1$ with
2087: $D_y \lesssim 1$ corresponds to a quasi-homogeneous non-Gaussian
2088: source. In Section \ref{sub:nong} we will see that in this case
2089: the VCZ theorem is applicable, and its region of applicability is
2090: in agreement with our definition of far-zone $\hat{z}^2 \gg N_y
2091: \gg 1$.
2092: 
2093: When $D_y \gg 1$ (with arbitrary $N_y$) and condition
2094: (\ref{farzo}) holds, analysis shows that the phase factor under
2095: the integral sign in Eq. (\ref{Gnoryany2}) can be neglected.
2096: Moreover, in this case, the Gaussian function in $\phi_y +
2097: \bar{\theta}_y$ has a width in $\phi_y$ much larger than unity,
2098: while the integral in $d \phi_x$ in second line of Eq.
2099: (\ref{Gnoryany2}) has a width in $\phi_y$ of order unity, because
2100: $\Psi$ does not depend on parameters. Therefore, the dependence in
2101: $\phi_y$ in the Gaussian function can
2102: always be omitted, and the Gaussian function factors out of the integral sign in $d \phi_y$. %It follows that Eq. (\ref{Gnoryany2})
2103: %simplifies to
2104: %
2105: %
2106: %\begin{eqnarray}
2107: %&&\hat{G}_y\left(\hat{z},{\bar{y}},\Delta {y}\right) = \frac{
2108: %\hat{z}\exp\left[{i \hat{z}\bar{\theta}_y \Delta{\theta}_y }
2109: %\right]}{2 \pi\sqrt{ \pi D_y}} \exp\left[-\frac{N_y \Delta
2110: %\theta_y^2 }{2}\right] \exp\left[-\frac{{\bar{\theta}_y}^2}{2 D_y
2111: %}\right]  \int d{\phi}_y \cr && \times\int d \phi_x
2112: %{\Psi}\left\{\hat{z},\left[{\phi}_x^2+\left(\phi_y-\frac{\Delta
2113: %\theta_y }{2}\right)^2\right]^{1/2}\right\}
2114: %{\Psi}^*\left\{\hat{z},\left[{\phi}_x^2+\left(\phi_y+\frac{\Delta
2115: %\theta_y }{2}\right)^2\right]^{1/2}\right\}~. \cr &&
2116: %\label{Gnoryany4}
2117: %\end{eqnarray}
2118: %%
2119: One is left with the product of exponential functions and the
2120: double integral
2121: 
2122: \begin{eqnarray}
2123: \tilde{f}(\Delta \theta_y) &\equiv&  \int d{\phi}_y\int d \phi_x
2124: \cr && \times {\Psi}\left\{\hat{z},\left[{\phi}_x^2+
2125: \left(\phi_y-\frac{\Delta
2126: \theta_y}{2}\right)^2\right]^{1/2}\right\}
2127: {\Psi}^*\left\{\hat{z},\left[{\phi}_x^2+\left(\phi_y+ \frac{\Delta
2128: \theta_y}{2}\right)^2\right]^{1/2}\right\}~, \cr &&
2129: \label{doubleint}
2130: \end{eqnarray}
2131: %
2132: which has a very peculiar property. In fact, it does not depend on
2133: $\hat{z}$. The proof is based on the autocorrelation theorem in
2134: two dimensions, and is given in detail in Appendix C of reference
2135: \cite{OURU}. This quite remarkable property of $\tilde{f}$ carries
2136: the consequence that substitution of $\Psi$ with $\Psi_f$ can be
2137: performed in Eq. (\ref{doubleint}) without altering the final
2138: result.
2139: 
2140: When both $D_y \gg 1$ and $N_y \gg 1$, one may neglect the
2141: dependence in $\Delta \theta_y$ in functions $\Psi$ and $\Psi^*$
2142: in Eq. (\ref{doubleint}), because the exponential function in
2143: $\Delta \theta_y$ before the integral sign limits the range of
2144: $\Delta \theta_y$ to values of order $1/\sqrt{N_y} \ll 1$. As a
2145: result, the double integration in $d \phi_x $ and $d \phi_y$
2146: yields a constant, and the description of the photon beam is
2147: independent of $\Psi$, i.e. does not include diffraction effects.
2148: This result is intuitive: when the electron beam size and
2149: divergence is large compared to the diffraction size and
2150: divergence, the photon beam can be described in terms of the
2151: phase-space distribution of the electron beam. This approach will
2152: be treated in more detail in Section \ref{sub:geop}.
2153: 
2154: Finally, when $D_y \gg 1$ and $N_y \lesssim 1$, diffraction
2155: effects cannot be neglected, nor can be the dependence in $\Delta
2156: \theta_y$ in $\Psi$ and $\Psi^*$. In this case, from condition
2157: (\ref{farzo}) we obtain that the far-zone coincides with $D_y
2158: \hat{z}^2 \gg 1$. This result is completely counterintuitive. In
2159: fact, since $D_y \gg 1$, the far-zone is reached for values
2160: $\hat{z} \sim 1$, i.e. at the very end of the undulator. Yet,
2161: diffraction effects cannot be neglected, and the field from a
2162: single electron is far from exhibiting a spherical wavefront at
2163: $\hat{z} \sim 1$. This paradox is solved by the special property
2164: of the double integral in Eq. (\ref{doubleint}), that allows one
2165: to substitute $\Psi$ with $\Psi_f$ independently of the value of
2166: $\hat{z}$.
2167: 
2168: As it will be discussed in Section \ref{sub:gaus} and Section
2169: \ref{sub:nong}, the case $D_y \gg 1$ corresponds to a
2170: quasi-homogeneous Gaussian source when $N_y \gg 1$ and to a
2171: quasi-homogeneous non-Gaussian source when $N_y \lesssim 1$. It
2172: will be shown that the VCZ theorem is applicable to these
2173: situations. In particular, the applicability region of the VCZ
2174: theorem will be seen to be in agreement with our definition of
2175: far-zone.
2176: 
2177: Our discussion can be summarized in a single statement. The far
2178: zone is defined, in terms of problem parameters, by condition
2179: (\ref{farzo}). Remembering this condition one can derive the
2180: far-zone expression for $\hat{G}_y$ simplifying Eq.
2181: (\ref{Gnoryany2}):
2182: 
2183: \begin{eqnarray}
2184: \hat{G}_y(\hat{z},\bar{\theta}_y,\Delta {\theta}_y) &=&
2185: \frac{1}{\hat{z}} \frac{1}{2\pi\sqrt{\pi D_y}} {\exp{\left[i
2186: \hat{z} \bar{\theta}_y \Delta {\theta}_y \right]}} \exp{\left[-
2187: \frac{N_y \Delta {\theta}_y^2}{2} \right]} \cr &&\times
2188: \int_{-\infty}^{\infty} d {\phi}_y
2189: \exp{\left[-\frac{(\bar{\theta}_y+{\phi}_y)^2}{2 D_y}\right]}
2190: \frac{1}{\mathcal{K}_F}\mathcal{F}\left({\phi}_y,\Delta
2191: {\theta}_y\right) ~,\label{resu1}
2192: \end{eqnarray}
2193: %
2194: where the universal two-dimensional
2195: function\footnote{$\mathcal{F}$ stands for "Far".}
2196: $\mathcal{F}(\alpha,\delta)$ is normalized in such a way that
2197: $\mathcal{F}(0,0)=1$ and reads :
2198: 
2199: \begin{eqnarray}
2200: \mathcal{F}(\alpha,\delta)&=& \mathcal{K}_F
2201: {\int_{-\infty}^{\infty} d {\phi}_x \Psi_f\left\{\left[{{\phi}_x^2
2202: +\left(\alpha-\frac{\delta}{2}\right)^2}\right]^{1/2}\right\}}
2203: \Psi_f\left\{\left[{{\phi}_x^2
2204: +\left(\alpha+\frac{\delta}{2}\right)^2}\right]^{1/2}\right\}~,\cr
2205: && \label{Motherofall}
2206: \end{eqnarray}
2207: %
2208: with $\mathcal{K}_F={3}/({8 \sqrt{\pi}})$.
2209: 
2210: 
2211: Note that, similarly to the source case, the right hand side of
2212: Eq. (\ref{resu1}) is found by convolving an universal function,
2213: $\mathcal{F}$, with a Gaussian function and multiplying the result
2214: by another Gaussian function.
2215: 
2216: A plot of $\mathcal{F}$ function defined is given in Fig. \ref{M}.
2217: $\mathcal{F}$ is a real function.  Thus,  only the geometrical
2218: phase factor in Eq. (\ref{resu1}) prevents $\hat{G}_y$ in the
2219: far-zone from being real. Another remarkable property of
2220: $\mathcal{F}$ is its invariance for exchange of $\alpha$ with
2221: $\delta/2$. Also, $\mathcal{F}$ is invariant for exchange of
2222: $\alpha$ with $-\alpha$ (or $\delta$ with $-\delta$).
2223: 
2224: Finally, from Eq. (\ref{psi0psif}) and Eq. (\ref{mtcm}) we have
2225: 
2226: \begin{eqnarray}
2227: {\mathcal{S}}(x,y) = \frac{\mathcal{K}_S}{4\pi^3 \mathcal{K}_F}
2228: \int_{-\infty}^{\infty} d\alpha \int_{-\infty}^{\infty} d\delta
2229: \exp\left[ 2i x \alpha +\frac{i y \delta}{2}\right]
2230: \mathcal{F}(\alpha,\delta)~. \label{calM}
2231: \end{eqnarray}
2232: %
2233: 
2234: \subsection{\label{sub:cross2} Intensity distribution}
2235: 
2236: With expressions for the cross-spectral density at hand, it is now
2237: possible to investigate the intensity distribution\footnote{Words
2238: "intensity distribution" include some abuse of language here and
2239: in the following. What we really calculate is the ensemble average
2240: of the square modulus of the normalized field, $\hat{I}_x
2241: \hat{I}_y = \langle |\hat{E}|^2\rangle$. Conversion to dimensional
2242: units, followed by multiplication by $c/(4\pi^2)$ yields the
2243: spectral density normalized to the electron number $N_e$.} along
2244: the beamline, letting $\Delta \vec{r} = 0$ in the expression for
2245: $\hat{G}$. Since ${N}_x \gg 1$ and $D_x \gg 1$, factorization of
2246: the cross-spectral density still holds. Therefore we can
2247: investigate the intensity profile along the vertical direction
2248: without loss of generality.
2249: 
2250: Posing $\Delta y = 0$ in Eq. (\ref{Gnor4y}) we obtain the
2251: intensity profile at the virtual source,
2252: $\hat{I}_{y}\left(0,{\bar{y}}\right)$, as a function of $\bar{y}$:
2253: 
2254: \begin{eqnarray}
2255: \hat{I}_y\left(0,{\bar{y}}\right) &=& \frac{1}{2 \mathcal{K}_S
2256: }\sqrt{\frac{\pi}{N_y}} \int_{-\infty}^{\infty} d \phi_y
2257: \exp\left[-\frac{\left(\phi_y+\bar{y}\right)^2}{2 N_y}\right]~
2258: \mathcal{I}_S(\phi_y) ~, \label{Inty0}
2259: \end{eqnarray}
2260: %
2261: where we introduced the universal function
2262: 
2263: \begin{eqnarray}
2264: \mathcal{I}_S(\alpha)\equiv\mathcal{S}(\alpha,0) = \mathcal{K}_S
2265: \int_{-\infty}^{\infty} d \phi_x
2266: \Psi_0^2\left(\sqrt{\phi_x^2+\alpha^2}\right)~. \label{Bdefine}
2267: \end{eqnarray}
2268: %
2269: A change of the integration variable: $\phi_x \longrightarrow x
2270: \equiv (\phi_x^2+\alpha^2)^{1/2}$ allows the alternative
2271: representation:
2272: 
2273: \begin{eqnarray}
2274: \mathcal{I}_S(\alpha)= 2\mathcal{K}_S \int_{0}^{\infty} d x
2275: \frac{\mathrm{rect}\left[\alpha/(2 x)
2276: \right]}{\sqrt{1-\alpha^2/x^2}}\Psi_0^2\left(x\right)~, \label{B2}
2277: \end{eqnarray}
2278: %
2279: where the function $\mathrm{rect}(\xi)$ is defined following
2280: \cite{GOOD}: it is equal to unity for $|\xi|\leqslant 1/2$ and
2281: zero otherwise.
2282: 
2283: \begin{figure}
2284: \begin{center}
2285: \includegraphics*[width=140mm]{figure11.eps}% Here is how to import EPS art
2286: \caption{\label{ftbetaplot} The universal function
2287: ${\mathcal{I}_S}$, used to calculate intensity at the
2288: virtual-source position.}
2289: \end{center}
2290: \end{figure}
2291: %
2292: 
2293: The intensity at the virtual-source position is given in terms of
2294: a convolution of a Gaussian function with the universal function
2295: $\mathcal{I}_S$. A plot of $\mathcal{I}_S$ is given in Fig.
2296: \ref{ftbetaplot} .
2297: 
2298: 
2299: Similar derivations can be performed in the far zone. Posing
2300: $\Delta \theta_y = 0$ in Eq. (\ref{resu1}) we obtain the
2301: directivity diagram of the radiation as a function of
2302: $\bar{\theta}_y$:
2303: 
2304: 
2305: \begin{eqnarray}
2306: \hat{I}_y(\hat{z},\bar{\theta}_y) &=& \frac{1}{\hat{z}}
2307: \frac{1}{\sqrt{2 \pi D_y} \mathcal{K}_F } \int_{-\infty}^{\infty}
2308: d {\phi}_y \exp{\left[-\frac{(\bar{\theta}_y+{\phi}_y)^2}{2
2309: D_y}\right]} \mathcal{I}_F\left({\phi}_y\right) ~,\label{Intyf}
2310: \end{eqnarray}
2311: %
2312: where we defined
2313: 
2314: \begin{eqnarray}
2315: \mathcal{I}_F(\alpha) \equiv  \mathcal{F}\left(\alpha,0\right)
2316: &=&\mathcal{K}_F {\int_{-\infty}^{\infty} d {\phi}_x
2317: \Psi_f^2\left[\left({{\phi}_x^2 +\alpha^2}\right)^{1/2}\right]}
2318: \cr &=&\mathcal{K}_F {\int_{-\infty}^{\infty} d {\phi}_x
2319: \mathrm{sinc}^2\left[\frac{{{\phi}_x^2 +\alpha^2}}{4}\right]}~.
2320: \label{Isintro}
2321: \end{eqnarray}
2322: %
2323: 
2324: \begin{figure}
2325: \begin{center}
2326: \includegraphics*[width=140mm]{figure12.eps}% Here is how to import EPS art
2327: \caption{\label{ISnorm} The universal function $\mathcal{I}_F$,
2328: used to calculate the intensity in the far zone.}
2329: \end{center}
2330: \end{figure}
2331: %
2332: 
2333: The intensity in the far-zone is given in terms of a convolution
2334: of a Gaussian function with the universal function
2335: $\mathcal{I}_F$. A plot of $\mathcal{I}_F$ is given in Fig.
2336: \ref{ISnorm}.
2337: 
2338: \subsection{\label{sub:degree} Spectral degree of coherence}
2339: 
2340: We can now present expressions for the spectral degree of
2341: coherence at the virtual source (that is a real quantity) and in
2342: the far-zone (that is not real). Eq. (\ref{normfine}) can be
2343: written for the vertical direction and in normalized units as:
2344: 
2345: 
2346: \begin{equation}
2347: \hat{g}_y\left(\hat{z},{\bar{y}},\Delta y\right) =
2348: \frac{\hat{G}_y\left(\hat{z},{\bar{y}},\Delta {y}\right)}{
2349: \left[\hat{I}_y(\hat{z},\bar{y}+\Delta y/2)\right]^{1/2}\left[
2350: \hat{I}_y(\hat{z},\bar{y}-\Delta y/2)\right]^{1/2}} ~.
2351: \label{normfine2}
2352: \end{equation}
2353: %
2354: Substitution of Eq. (\ref{Gnor4y}) and Eq. (\ref{Inty0}) in Eq.
2355: (\ref{normfine2}) gives the spectral degree of coherence at the
2356: virtual source:
2357: 
2358: \begin{eqnarray}
2359: \hat{g}_y\left(0,{\bar{y}},\Delta {y}\right) &=& \exp
2360: \left[-\frac{(\Delta y)^2 D_y }{2}\right] \int_{-\infty}^{\infty}
2361: d \phi_y \exp\left[-\frac{\left(\phi_y+\bar{y}\right)^2}{2
2362: N_y}\right] ~ \mathcal{S}(\phi_y,\Delta y)\cr && \times \left\{
2363: \int_{-\infty}^{\infty} d \phi_y
2364: \exp\left[-\frac{\left(\phi_y+\bar{y}+\Delta y/2\right)^2}{2
2365: N_y}\right]~ \mathcal{I}_S(\phi_y)\right\}^{-1/2} \cr &&\times
2366: \left\{ \int_{-\infty}^{\infty} d \phi_y
2367: \exp\left[-\frac{\left(\phi_y+\bar{y}-\Delta y/2\right)^2}{2
2368: N_y}\right]~ \mathcal{I}_S(\phi_y)\right\}^{-1/2} ~. \label{g0}
2369: \end{eqnarray}
2370: %
2371: Similarly, substitution of Eq. (\ref{resu1}) and Eq. (\ref{Intyf})
2372: in Eq. (\ref{normfine2}) gives the spectral degree of coherence in
2373: the far zone:
2374: 
2375: 
2376: \begin{eqnarray}
2377: \hat{g}_y(\hat{z},\bar{\theta}_y,\Delta {\theta}_y) &=&
2378: {\exp{\left[i  \hat{z} \bar{\theta}_y \Delta {\theta}_y \right]}}
2379: \exp{\left[-\frac{  (\Delta {\theta}_y)^2 N_y}{2} \right]}\cr &&
2380: \times \int_{-\infty}^{\infty} d {\phi}_y
2381: \exp{\left[-\frac{(\bar{\theta}_y+{\phi}_y)^2}{2 D_y}\right]}
2382: \mathcal{F}\left({\phi}_y,\Delta {\theta}_y\right)\cr && \times
2383: \left\{\int_{-\infty}^{\infty} d {\phi}_y
2384: \exp{\left[-\frac{(\bar{\theta}_y+\Delta \theta_y/2+
2385: {\phi}_y)^2}{2 D_y}\right]}
2386: \mathcal{I}_F\left({\phi}_y\right)\right\}^{-1/2}\cr && \times
2387: \left\{\int_{-\infty}^{\infty} d {\phi}_y
2388: \exp{\left[-\frac{(\bar{\theta}_y-\Delta \theta_y/2+
2389: {\phi}_y)^2}{2 D_y}\right]}
2390: \mathcal{I}_F\left({\phi}_y\right)\right\}^{-1/2}~.\label{gfar}
2391: \end{eqnarray}
2392: %
2393: 
2394: 
2395: \subsection{\label{sub:limit} Influence of horizontal emittance on
2396: vertical coherence for $D_y \ll 1$ and $N_y \ll 1$}
2397: 
2398: The theory developed up to now is valid for arbitrary values of
2399: $N_y$ and $D_y$. In the present Section \ref{sub:limit} we discuss
2400: an application in the limiting case for ${D}_y \ll 1$ and ${N}_y
2401: \ll 1$ corresponding to third generation light sources operating
2402: in  the soft X-ray range.
2403: 
2404: At the virtual source position, the following simplified
2405: expression for the spectral degree of coherence in the vertical
2406: direction is derived from Eq. (\ref{g0}):
2407: 
2408: \begin{eqnarray}
2409: \hat{g}_y\left(0,{\bar{y}},\Delta {y}\right) =
2410: \mathcal{X}_S({\bar{y}},\Delta {y}) ~, \label{g0soft}
2411: \end{eqnarray}
2412: %
2413: where
2414: 
2415: \begin{equation}
2416: \mathcal{X}_S(\alpha,\delta) = \frac{{\mathcal{S}}(\alpha,\delta)}
2417: {\left[{\mathcal{I}_S}(\alpha-\delta/2)\right]^{1/2}
2418: \left[{\mathcal{I}_S}(\alpha+\delta/2)\right]^{1/2}}~.
2419: \label{mathcalchi}
2420: \end{equation}
2421: %
2422: Thus, $\hat{g}_y$ is given by the universal function
2423: $\mathcal{X}_S$. A plot of $\mathcal{X}_S$ is given in Fig.
2424: \ref{ftchi}.
2425: 
2426: 
2427: \begin{figure}
2428: \begin{center}
2429: \includegraphics*[width=140mm]{figure13.eps}% Here is how to import EPS art
2430: \caption{\label{ftchi} Plot of the universal function
2431: $\mathcal{X}_S$, used to calculate the modulus of the spectral
2432: degree of coherence of the source when $\hat{N}_x \gg1$,
2433: $\hat{D}_x \gg 1$, $\hat{N}_y \ll 1$ and $\hat{D}_y \ll 1$.}
2434: \end{center}
2435: \end{figure}
2436: %
2437: 
2438: \begin{figure}
2439: \begin{center}
2440: \includegraphics*[width=140mm]{figure14.eps}% Here is how to import EPS art
2441: \caption{\label{chi} Plot of the universal function
2442: $\mathcal{X}_F$, used to calculate the modulus of the spectral
2443: degree of coherence in the far zone when $\hat{N}_x \gg1$,
2444: $\hat{D}_x \gg 1$, $\hat{N}_y \ll 1$ and $\hat{D}_y \ll 1$.}
2445: \end{center}
2446: \end{figure}
2447: %
2448: 
2449: Similarly, in  the far zone, one obtains from Eq. (\ref{gfar}):
2450: 
2451: \begin{eqnarray}
2452: \hat{g}_y(\hat{z},\bar{\theta}_y,\Delta {\theta}_y) &=&
2453: {\exp{\left[i \hat{z} \bar{\theta}_y \Delta {\theta}_y \right]}}
2454: \mathcal{X}_F({\bar{\theta}_y},\Delta \theta_{y}) ~.
2455: \label{gfarsoft}
2456: \end{eqnarray}
2457: %
2458: Here the universal function $\mathcal{X}_F$ is given by
2459: 
2460: \begin{equation}
2461: \mathcal{X}_F{\left(\alpha,\delta\right)} =
2462: \frac{\mathcal{F}(\alpha,\delta)}{\left[\mathcal{I}_F(\alpha-\delta/2)\right]^{1/2}
2463: \left[\mathcal{I}_F(\alpha+\delta/2)\right]^{1/2}}~.
2464: \label{chidef}
2465: \end{equation}
2466: %
2467: Thus, $|\hat{g}_y|$ is given by the universal function
2468: $\mathcal{X}_F$. A plot of $\mathcal{X}_F$ is presented in Fig.
2469: \ref{chi} as a function of dummy variable $\alpha$ and $\delta$.
2470: It is straightforward to see that
2471: $\mathcal{X}_F({\bar{\theta}_y},\Delta \theta_{y})$ is symmetric
2472: with respect to $\Delta {\theta}_y$ and with respect to the
2473: exchange of $\Delta {\theta}_y/2$  with $\bar{\theta}_y$. When
2474: $\bar{\theta}_y =0$, i.e. $\hat{\theta}_{y1}= -\hat{\theta}_{y2}$,
2475: we obviously obtain $\mathcal{X}_F(0,\Delta {\theta}_y)=1$ that
2476: corresponds to complete coherence at this particular value of
2477: $\bar{\theta}_y$. However, since $|\hat{g}_y| = \mathcal{X}_F$
2478: oscillates from positive to negative values, in general one never
2479: has full coherence in the vertical direction, even in the case of
2480: zero vertical emittance. Note that this effect does not depend, in
2481: the limit for ${N}_x \gg 1$ and ${D}_x \gg 1$, on the actual
2482: values of ${N}_x$ and ${D}_x$.
2483: 
2484: 
2485: \begin{figure}
2486: \begin{center}
2487: \includegraphics*[width=140mm]{plot2Dcompnew.eps}% Here is how to import EPS art
2488: \caption{\label{plot2Dcomp} Comparison between some zeros of
2489: $\mathcal{X}_F$, at coordinates  $(\bar{\theta}_y,\Delta
2490: {\theta}_{y}$ (black circles), and the directivity diagram of
2491: undulator radiation in the vertical direction at very large
2492: horizontal electron beam divergence $\hat{D}_x \gg 1$ and
2493: negligible vertical divergence $\hat{D}_y \ll 1$ (solid line).}
2494: \end{center}
2495: \end{figure}
2496: %
2497: In other words, as it is evident from Fig. \ref{chi},
2498: $\mathcal{X}_F({\bar{\theta}_y},\Delta \theta_{y})$ exhibits many
2499: different zeros in $\bar{\theta}_y$ for any fixed value of $\Delta
2500: {\theta}_y$. In Fig. \ref{plot2Dcomp} some of these zeros are
2501: illustrated with black circles on the plane
2502: $(\bar{\theta}_y,\Delta {\theta}_{y})$. Consider a two-pinhole
2503: experiment as in Fig. \ref{ed12c}. Once a certain distance
2504: $\hat{z} \Delta {\theta}_y$ between the two pinholes is fixed,
2505: Fig. \ref{plot2Dcomp} illustrates at what position of the pinhole
2506: system, $\bar{\theta}_{y}$, the spectral degree of coherence in
2507: the vertical direction drops from unity to zero for the first
2508: time.
2509: 
2510: To estimate the importance of this effect, it is crucial to
2511: consider the position of $\bar{\theta}_{y}$ in the directivity
2512: diagram of the radiant intensity, that coincides in this
2513: case\footnote{In other words, it can be shown that $\mathcal{I}_F$
2514: is the directivity diagram corresponding to the case $N_x\gg1$,
2515: $D_x \gg1$, $N_y \ll 1$, $D_y \ll 1$.} with
2516: $\mathcal{I}_F(\bar{\theta}_y)$  (solid line in Fig.
2517: \ref{plot2Dcomp}).
2518: %
2519: From Fig. \ref{plot2Dcomp} one can see that
2520: $\mathcal{X}_F({\bar{\theta}_y},\Delta \theta_{y})$ drops to zero
2521: for the first time at $\Delta {\theta}_y \sim 2$ $\bar{\theta}_y
2522: \sim 2$, where the X-ray flux is still intense. This behavior of
2523: the degree of coherence may influence particular kind of
2524: experiments. To give an example we go back to the two-pinhole
2525: setup in Fig. \ref{ed12c}. After spatial filtering in the
2526: horizontal direction, one will find that for some vertical
2527: position $\bar{\theta}_y$ of the pinholes (at fixed $\Delta
2528: {\theta}_y$), well within the radiation pattern diagram, there
2529: will be no fringes, while for some other vertical position there
2530: will be perfect visibility. Without the knowledge of the function
2531: $\mathcal{X}_F$ it would not be possible to fully predict the
2532: outcomes of a two-pinhole experiment. Results described here
2533: should be considered as an illustration of our general theory that
2534: may or may not, depending on the case under study, have practical
2535: influence. It may, however, be the subject of experimental
2536: verification.
2537: 
2538: \subsubsection{Discussion}
2539: 
2540: Although the $\mathcal{X}_F$ is independent of $N_x$ and $D_x$,
2541: its actual shape is determined by the presence of a large
2542: horizontal emittance. Let us show this fact. If both $N_{x,y} \ll
2543: 1$ and $D_{x,y}\ll 1$ (filament beam limit), one would have had
2544: $|\hat{G}| = \Psi_f\left(\left|\vec{\theta}_1\right|\right)
2545: \Psi_f\left(\left|\vec{\theta_2}\right|\right)$, so that
2546: $|\hat{g}|=1$, strictly. Note that in this case $\hat{G}$ could
2547: not be factorized. When instead $N_x \gg 1$ and $D_x \gg 1$ the
2548: cross-spectral density can be factorized according to Eq.
2549: (\ref{factorizeee}), but the integration in $d\phi_x$, that
2550: follows from an integration over the horizontal electron beam
2551: distribution, is still included in the vertical cross-spectral
2552: density $\hat{G}_x$ (and, therefore, in $\mathcal{X}_F$). This
2553: results in the outcome described above and can be traced back to
2554: the non-Gaussian nature of $\Psi_f$. Note that if one adopted a
2555: Gaussian-Schell model, the cross-spectral density could have been
2556: split in the product of Gaussian intensity and Gaussian spectral
2557: degree of coherence and $\Psi_f$, being a product of Gaussians,
2558: would have been separable. Then, $|\hat{g}_y|$ would have been
2559: constant for $N_x\gg 1$ and $D_x \gg 1$. As a result, the effect
2560: described here would not have been recognized. This fact
2561: constitutes a particular realization of our general remarks about
2562: Gaussian-Schell models at the end of Section \ref{sub:evol}.
2563: 
2564: As a final note to the entire Section, we stress the fact that our
2565: theory of partial coherence in third generation light sources is
2566: valid under several non-restrictive assumptions. Alongside with
2567: previously discussed conditions $\gamma^2 \gg 1$, $N_w \gg 1$,
2568: $\sigma_T \omega \gg N_w$ and the assumption of perfect resonance
2569: (i.e. the limit $\hat{C} \ll 1$), we assumed separability and
2570: particular shape of the electron beam phase space (see Eq.
2571: (\ref{distr})). Moreover, for third generation light sources we
2572: assumed $\epsilon_x \gg \lambdabar$ (up to the VUV range).
2573: Together with $\beta_{x,y} \simeq L_w$ this implies $N_x \gg 1$
2574: and $D_x \gg 1$, allowing for separability of the cross-spectral
2575: density in horizontal and vertical factors. Moreover, due to $N_x
2576: \gg 1$ and $D_x \gg 1$, we are dealing with quasi-homogenous
2577: Gaussian sources in the horizontal direction. This particular kind
2578: of sources will be treated in detail as an asymptote of our
2579: general theory in the next Section. In the vertical direction
2580: instead, we still have fully generic sources. We showed how the
2581: vertical cross-spectral density can be expressed in terms of
2582: convolutions between two-dimensional universal functions and
2583: Gaussian functions. A particularly interesting case is that of
2584: quasi-homogeneous non-Gaussian sources that will also be treated
2585: as an asymptotic case in the following Section dedicated to
2586: quasi-homogeneous sources.
2587: 
2588: \section{\label{sec:quas} Quasi-homogeneous asymptotes for
2589: undulator sources}
2590: 
2591: In Section \ref{sec:main} we developed a general theory of
2592: transverse coherence properties of third generation light sources.
2593: In this Section we consider the class of quasi-homogeneous sources
2594: for undulator devices as an asymptotic limit for that theory.
2595: 
2596: Quasi-homogeneous sources are defined by the fact that the
2597: cross-spectral density of the virtual source (positioned at $z=0$)
2598: can be written as:
2599: 
2600: 
2601: \begin{eqnarray}
2602: \hat{G}\left(0,\vec{\bar{r}},\Delta \vec{r}\right) =
2603: \hat{I}\left(0,\vec{\bar{r}}\right) \hat{g}\left(\Delta
2604: \vec{r}\right) ~.\label{introhdopo}
2605: \end{eqnarray}
2606: %
2607: The definition of quasi-homogeneity amounts to a factorization of
2608: the cross-spectral density as the product of the field intensity
2609: distribution and the spectral degree of coherence. A set of
2610: necessary and sufficient conditions for such factorization to be
2611: possible follows: (i) the radiation intensity at the virtual
2612: source varies very slowly with the position across the source on
2613: the scale of the field correlation length and (ii) the spectral
2614: degree of coherence depends on the positions across the source
2615: only through the difference $\Delta \vec{r}$.
2616: 
2617: Factorization of Eq. (\ref{Gnor3}) as in Eq. (\ref{introhdopo}),
2618: for third generation light sources, is equivalent to a particular
2619: choice of the region of parameters for the electron beam: ${N}_x
2620: \gg 1$, ${D}_x \gg 1$ and either (or both) ${N}_y \gg 1$ and
2621: ${D}_y \gg 1$\footnote{These conditions describe the totality of
2622: third generation quasi-homogeneous sources. In fact, while a
2623: purely mathematical analysis indicates that factorization of Eq.
2624: (\ref{Gnor3}) is equivalent to more generic conditions (${N}_x
2625: \gg1$ and ${N}_y \gg 1$, or ${D}_x \gg1$ and ${D}_y \gg 1$),
2626: comparison with third generation source parameters ($N_x \gg 1$
2627: and $D_x \gg 1$) reduces such conditions to the already mentioned
2628: ones.}. In this case, the reader may verify that conditions (i)
2629: and (ii) are satisfied.
2630: 
2631: In the horizontal direction, we have both $N_x \gg 1$ and $D_x \gg
2632: 1$ for wavelengths up to the VUV range, so that factorization of
2633: the cross-spectral density in horizontal and vertical
2634: contributions $\hat{G}_x$ and $\hat{G}_y$ always holds.
2635: 
2636: Let us first consider $\hat{G}_y$. Depending on the values of
2637: $N_y$ and $D_y$ we may have Gaussian quasi-homogeneous sources
2638: characterized by a Gaussian transverse distribution of intensity
2639: ($N_y \gg 1$ and $D_y \gg 1$), as well as non-Gaussian
2640: quasi-homogeneous sources ($N_y \gg 1$ and $D_y \lesssim 1$ or
2641: $N_y \lesssim 1$ and $D_y \gg 1$). Gaussian quasi-homogenous
2642: sources are to be expected in the vertical direction in the hard
2643: X-ray limit, where diffraction effects play no role. On the
2644: contrary, diffraction effects must be accounted for when dealing
2645: with non-Gaussian quasi-homogeneous sources.
2646: 
2647: Let us now consider $\hat{G}_x$. In the horizontal direction both
2648: $N_x \gg 1$ and $D_x \gg 1$. It follows that Gaussian
2649: quasi-homogeneous sources find a very natural application in the
2650: description of the cross-spectral density in the horizontal
2651: direction, from the hard X-ray to the VUV range.
2652: 
2653: We will see that the VCZ theorem applies to all quasi-homogeneous
2654: cases. Actually, the concept of far-zone for quasi-homogeneous
2655: sources can be introduced as the region in the parameter space
2656: $\hat{z}, N_{x,y}, D_{x,y}$ such that the VCZ theorem holds. We
2657: will see that these condition coincides with condition
2658: (\ref{farzo}) given before.
2659: 
2660: \subsection{\label{sub:gaus} Gaussian undulator sources}
2661: 
2662: When $N_x \gg 1$ and $D_x \gg 1$ Eq. (\ref{Gnor3x}) applies. When
2663: $N_y \gg 1$ and $D_y \gg 1$, Eq. (\ref{Gnor3y}) reduces to
2664: 
2665: \begin{eqnarray}
2666: \hat{G}_y\left(0,{\bar{y}},\Delta {y}\right) &=&
2667: \sqrt{\frac{\pi}{{N_y}}} \exp \left[-\frac{(\Delta y)^2D_y
2668: }{2}\right] \exp\left[-\frac{\bar{y}^2}{2 N_y}\right]
2669: ~,\label{Gnorgauss0}
2670: \end{eqnarray}
2671: %
2672: that is equivalent to Eq. (\ref{Gnor3x}): thus, identical
2673: treatments hold separately in the horizontal and vertical
2674: directions. For this reasons, and for simplicity of notation, we
2675: drop all subscripts "x" or "y" in the present Section
2676: \ref{sub:gaus}, and we substitute letters "x" and "y" in variables
2677: with the more generic "r". However, as stated before, the Gaussian
2678: quasi-homogeneous model primarily describes third generation light
2679: sources in the horizontal direction.
2680: 
2681: Eq. (\ref{introhdopo}) is satisfied. Moreover,
2682: 
2683: \begin{eqnarray}
2684: \hat{I}\left(0,{\bar{r}}\right)=\sqrt{\frac{\pi}{{N}}}
2685: \exp\left[-\frac{\bar{r}^2}{2 N}\right] \label{integauss}
2686: \end{eqnarray}
2687: %
2688: and
2689: 
2690: \begin{eqnarray}
2691: \hat{g}(\Delta r) = \exp \left[-\frac{(\Delta r)^2D }{2}\right]
2692: ~.\label{degregauss0}
2693: \end{eqnarray}
2694: %
2695: Propagation of Eq. (\ref{Gnorgauss0}) can be found taking the
2696: limit of Eq. (\ref{Gnoryany2}) for $N\gg 1$ and $D\gg 1$, which
2697: yields an analytical expression for the evolution of the
2698: cross-spectral density based on Eq. (\ref{Cany2}) at $\hat{C} \ll
2699: 1$ :
2700: 
2701: 
2702: 
2703: \begin{eqnarray}
2704: \hat{G}(\hat{z},\bar{r},\Delta r) &=& \frac{\sqrt{\pi}}{{\hat{z}
2705: \sqrt{A+D}}} \exp\left[-\frac{\bar{r}^{2}}{2({A}+{D})
2706: \hat{z}^2}\right] \exp\left[ i \frac{\bar{r}\Delta
2707: {r}}{\hat{z}}\right]\cr &&\times \exp\left[-  i
2708: \frac{{A}\bar{r}\Delta {r}}{\hat{z} ({A}+{D})}\right]
2709: \exp\left[-\frac{{A} {D} (\Delta r)^{2}}{2({A}+{D}) }\right]
2710: ~,\label{Ggaussz}
2711: \end{eqnarray}
2712: %
2713: where
2714: 
2715: \begin{equation}
2716: {A} = \frac{{N}}{\hat{z}^2} ~.\label{a2set}
2717: \end{equation}
2718: %
2719: We have
2720: 
2721: \begin{eqnarray}
2722: \hat{I}(\hat{z},\bar{r}) &=& \frac{\sqrt{\pi}}{{\hat{z}
2723: \sqrt{A+D}}} \exp\left[-\frac{\bar{r}^{2}}{2({A}+{D})
2724: \hat{z}^2}\right] ~,\label{Igaussz}
2725: \end{eqnarray}
2726: %
2727: and
2728: 
2729: \begin{eqnarray}
2730: \hat{g}(\hat{z},\Delta r) &=& \exp\left[i \frac{\bar{r}\Delta
2731: {r}}{\hat{z}}\right] \exp\left[-  i \frac{{A}\bar{r}\Delta
2732: {r}}{\hat{z} ({A}+{D})}\right] \exp\left[-\frac{{A} {D} (\Delta
2733: r)^{2}}{2({A}+{D}) }\right] ~.\label{degregaussz}
2734: \end{eqnarray}
2735: %
2736: Note that, due to the phase factors in Eq. (\ref{Ggaussz}), only
2737: the virtual source at $\hat{z}=0$ constitutes a quasi-homogeneous
2738: source.
2739: 
2740: Geometrical interpretation of ${A}$ is the dimensionless square of
2741: the apparent angular size of the electron bunch at the observer
2742: point position, calculated as if the beam was positioned at
2743: $\hat{z}=0$.
2744: 
2745: The far-zone for quasi-homogeneous Gaussian sources is given by
2746: condition $A \ll D$, which can be retrieved by condition
2747: (\ref{farzo}) or directly by Eq. (\ref{Ggaussz}). In this case,
2748: simplification of Eq. (\ref{Ggaussz}) or use of Eq. (\ref{resu1})
2749: in the limit for $N\gg 1$ and $D \gg 1$ yields the far-zone
2750: cross-spectral density:
2751: 
2752: 
2753: \begin{eqnarray}
2754: \hat{G}\left(\hat{z},{\bar{\theta}},\Delta {\theta}\right) &=&
2755: \frac{1}{\hat{z}} \sqrt{\frac{\pi}{D}} \exp\left[i \hat{z}
2756: {\bar{\theta}} \Delta {\theta}\right] \exp\left[-\frac{N (\Delta
2757: \theta)^2}{2}\right] \exp\left[-\frac{\bar{\theta}^2}{2 D}\right]
2758: ~,\cr &&\label{Gnorgaussf}
2759: \end{eqnarray}
2760: %
2761: so that
2762: 
2763: 
2764: \begin{eqnarray}
2765: \hat{I}\left(\hat{z},{\bar{\theta}}\right) &=& \frac{1}{\hat{z}}
2766: \sqrt{\frac{\pi}{D}} \exp\left[-\frac{\bar{\theta}^2}{2 D}\right]
2767: ~,\label{Integaussf}
2768: \end{eqnarray}
2769: %
2770: and
2771: 
2772: \begin{eqnarray}
2773: \hat{g}\left(\hat{z},\Delta {\theta}\right) &=&  \exp\left[i
2774: \hat{z} {\bar{\theta}}  \Delta {\theta}\right] \exp\left[-\frac{N
2775: (\Delta \theta)^2}{2}\right] ~.\label{degregaussf}
2776: \end{eqnarray}
2777: %
2778: Similarly as before we suppressed subscripts "x" or "y" in the
2779: symbol $\theta$.
2780: 
2781: Analysis of Eq. (\ref{Gnorgauss0}) and Eq. (\ref{Gnorgaussf})
2782: allows to conclude that
2783: 
2784: $(a_2)$ the spectral degree of coherence of the field at the
2785: source plane $g({\Delta {r}})$ and the angular distribution of the
2786: radiant intensity $I({\bar{\theta}})$ are a Fourier pair.
2787: 
2788: $(b_2)$ the spectral degree of coherence of the far field
2789: $g(\Delta {\theta})$ and the source-intensity distribution
2790: $I({\bar{r}})$ are, apart for a simple geometrical phase factor, a
2791: Fourier pair.
2792: 
2793: The statement $(b_2)$ is a version of the VCZ theorem valid for
2794: quasi-homogeneous sources.  Statement $(a_2)$ instead, regards the
2795: symmetry between space and angle domains, and can be seen as an
2796: inverse VCZ theorem.
2797: 
2798: This discussion underlines the link between the VCZ theorem and
2799: the Wiener-Khinchin theorem. Exactly as the space domain has a
2800: reciprocal description in terms of transverse (two-dimensional)
2801: wave vectors, the time domain has a reciprocal description in
2802: terms of frequency. The reader will recognize the analogy between
2803: statements $(a_2)$ and $(b_2)$, with statements $(a_1)$ and
2804: $(b_1)$ discussed in Section \ref{sub:temp}. In particular, the
2805: VCZ is analogous to the inverse Wiener-Khincin theorem. Similarly,
2806: separability of $G$ in Eq. (\ref{introhdopo}) in the product of
2807: spectral degree of coherence and intensity is analogous to
2808: separability of $\Gamma_\omega$, in Eq. (\ref{gamma6prima}) in the
2809: product of spectral correlation function and spectral density
2810: distribution of the source.
2811: 
2812: Let us calculate the transverse coherence length $\hat{\xi}_c$ as
2813: a function of the observation distance $\hat{z}$. We introduce the
2814: coherence length following the definition by Mandel \cite{WOLF}.
2815: The coherence length, naturally normalized to the diffraction
2816: length $\sqrt{L_w c/\omega}$ is defined as
2817: 
2818: \begin{equation}
2819: \hat{\xi}_c(\hat{z}) =  \int_{-\infty}^{\infty} |g(\hat{z},\Delta
2820: r )|^2 d(\Delta r)~, \label{cohlen}
2821: \end{equation}
2822: %
2823: Performing the integration in Eq. (\ref{cohlen}) with the help of
2824: Eq. (\ref{degregaussz}) yields:
2825: 
2826: \begin{equation}
2827: \hat{\xi}_c(\hat{z}) =  {\sqrt{\pi}}
2828: \left(\frac{1}{{A}}+\frac{1}{{D}}\right)^{1/2}~.\label{cohlen2}
2829: \end{equation}
2830: %
2831: \begin{figure}
2832: \begin{center}
2833: \includegraphics*[width=140mm]{figure15.eps}% Here is how to import EPS art
2834: \caption{\label{uno} Coherence length $\hat{\xi}_c$ as a function
2835: of $\hat{z}$ and asymptotic behaviors for $\hat{z} \longrightarrow
2836: 1/2$ and $\hat{z} \gg 1$. Here ${N} = 10^3$ and ${D} = 10$.}
2837: \end{center}
2838: \end{figure}
2839: %
2840: 
2841: The coherence length in Eq. (\ref{cohlen2}) exhibits linear
2842: dependence on $\hat{z}$, that is $\hat{\xi}_c \longrightarrow
2843: \sqrt{\pi/ {N}}~{\hat{z}}$ while for $\hat{z} \longrightarrow 1/2$
2844: that is at the end of the undulator, it converges to a constant
2845: $\hat{\xi}_c \longrightarrow [\pi/(4{N})+\pi/{D}]^{1/2}$. Eq.
2846: (\ref{cohlen2}) and its asymptotes are presented in Fig. \ref{uno}
2847: for the case ${N} = 10^3$, ${D} = 10$. At the exit of the
2848: undulator, $\hat{\xi}_c \sim 1/\sqrt{{D}}$, because ${N} \gg {D}$.
2849: On the other hand, horizontal dimension of the light spot is
2850: simply proportional to $\sqrt{{N}}$. This means that the
2851: horizontal dimension of the light spot is determined by the
2852: electron beam size, as is intuitive, while the beam angular
2853: distribution is printed in the fine structures of the intensity
2854: function that are of the dimension of the coherence length.  In
2855: the limit for $A \ll D$ the situation is reversed. The radiation
2856: field at the source can be presented as a superposition of plane
2857: waves, all at the same frequency $\omega$, but with different
2858: propagation angles with respect to the $z$-direction. Since the
2859: radiation at the exit of the undulator is partially coherent, a
2860: spiky angular distribution of intensity is to be expected. The
2861: nature of the spikes is easily described in terms of Fourier
2862: transform theory. From Fourier transform theorem  we can expect an
2863: angular spectrum with Gaussian envelope and rms width
2864: $\sqrt{{D}}$. Also, the angular distribution of intensity should
2865: contain spikes with characteristic width $1/\sqrt{N}$, as a
2866: consequence of the reciprocal width relations of Fourier transform
2867: pairs (see Fig. \ref{spikesp}). This is realized in mathematics by
2868: the expression for the cross-spectral density, Eq. (\ref{Ggaussz})
2869: and by the equation for the coherence length, Eq. (\ref{cohlen2}).
2870: 
2871: It is also important to remark that the asymptotic behavior for
2872: ${A} \ll 1$ of $\hat{g}$, that is Eq. (\ref{degregaussf}) and
2873: $\hat{\xi}_c$, that is
2874: 
2875: \begin{equation}
2876: \hat{\xi}_c \longrightarrow \sqrt{\frac{\pi}{
2877: N}}{\hat{z}}\label{cohlen2vcz}
2878: \end{equation}
2879: %
2880: \begin{figure}
2881: \begin{center}
2882: \includegraphics*[width=140mm]{figure16.eps}% Here is how to import EPS art
2883: \caption{\label{spikesp} Physical interpretation of the
2884: generalized VCZ theorem. If the radiation beyond the source plane
2885: is partially coherent, a spiky angular distribution of intensity
2886: is expected. The nature of these spikes is easily described in
2887: Fourier transform notations. We can expect that typical width of
2888: the angular distribution of intensity should be of order $(\omega
2889: \Delta/c)^{-1}$, where $\Delta$ is the typical linear dimension of
2890: spatially random intensity fluctuations. If the source has
2891: transverse size $d$, the angular distribution of intensity should
2892: contain spikes with typical width of about $(\omega d/c)^{-1}$, a
2893: consequence of the reciprocal width relations of Fourier transform
2894: pairs.}
2895: \end{center}
2896: \end{figure}
2897: %
2898: are direct application of VCZ theorem.  In fact, the last
2899: exponential factor on the right hand side of Eq. (\ref{Ggaussz})
2900: is simply linked with the Fourier transform of
2901: $f_{{l}}\left(\hat{l}\right)$. We derived Eq. (\ref{Ggaussz}) for
2902: ${N} \gg 1$ and ${D} \gg 1$, with $ A \ll D$: in non-normalized
2903: units these conditions mean that the VCZ theorem is applicable
2904: when the electron beam divergence is much larger than the
2905: diffraction angle, i.e. $\sigma'^2 \gg \lambda /(2\pi L_w)$, the
2906: electron beam dimensions are much larger than the diffraction
2907: size\footnote{We do not agree with statement in \cite{TAKA}: "the
2908: electron-beam divergence must be much smaller than the photon
2909: divergence" for the VCZ theorem to apply. This would imply
2910: $\sigma' \ll \sqrt{\lambda /(2\pi L_w)}$ (reference \cite{TAKA},
2911: page 571, Eq. (57)).}, i.e. $\sigma^2 \gg \lambda L_w/2\pi $, and
2912: $({\sigma'} {z})^2 \gg {\sigma}^2$.
2913: 
2914: In \cite{GOOD} (paragraph 5.6.4) a rule of thumb is given for the
2915: applicability region of the generalization of the VCZ theorem to
2916: quasi-homogeneous sources. The rule of thumb requires $z > 2 d
2917: \Delta/\lambda$ where $d$ is "the maximum linear dimension of the
2918: source", that is the diameter of a source with uniform intensity
2919: and $\Delta$ "represents the maximum linear dimension of a
2920: coherence area of the source". In our case, since $\sigma$ is the
2921: rms source dimension, $d\simeq 2\sigma$. Moreover, from Eq.
2922: (\ref{cohlen2}) we have $\Delta = \xi_c \simeq \lambda/(2
2923: \sqrt{\pi} \sigma')$. The rule of thumb then requires $z >2
2924: \sigma/( \sqrt{\pi} \sigma')$: in dimensionless form this reads $
2925: \hat{z} \gtrsim \sqrt{{N}/{D}}$. This is parametrically in
2926: agreement with our limiting condition $A \ll D$, even though these
2927: two conditions are different when it come to actual estimations:
2928: our condition $A \ll D$ is, in fact, only an asymptotic one. To
2929: see how well it works in reality we might consider the plot in
2930: Fig. \ref{uno}. There ${N} = 10^3$ and ${D} = 10$. Following
2931: \cite{GOOD} we may conclude that a good condition for the
2932: applicability of the VCZ theorem should be $\hat{z} \gtrsim 10$.
2933: However as it is seen from the figure, the linear asymptotic
2934: behavior is not yet a good approximation at $\hat{z} \simeq 10$.
2935: This may be ascribed to the fact that the derivation in
2936: \cite{GOOD} is not generally valid, but has been carried out for
2937: sources which drop to zero very rapidly outside the maximum linear
2938: dimension $d$ and whose correlation function also drops rapidly to
2939: zero very rapidly outside maximum linear dimension $\Delta$.
2940: However, at least parametrically, the applicability of the VCZ
2941: theorem in the asymptotic limit $A \ll D$ can be also expected
2942: from the condition $z > 2 d \Delta/\lambda$ in \cite{GOOD}.
2943: 
2944: 
2945: We conclude that the far field limit ${A} \ll {D}$ corresponds
2946: with the applicability region of the VCZ theorem. When this is the
2947: case, the VCZ theorem applies and the modulus of the spectral
2948: degree of coherence in the far field, forms a Fourier pair with
2949: the intensity distribution of the virtual source.  In particular
2950: one concludes that the rms width of the virtual source is
2951: $\sqrt{N}$. In our study case for $N \gg 1$ and $D \gg 1$, such a
2952: relation between the rms width of the spectral degree of coherence
2953: in the far field and the rms dimension of the virtual source is
2954: also a relation between the rms width of the cross-spectral
2955: density function in the far field and the rms dimension of the
2956: electron beam at the plane of minimal beta function in the center
2957: of the undulator. In dimensional units one can write the value
2958: $\sigma_c$ of the rms width of the spectral degree of coherence
2959: $\hat{g}(\Delta \vec{r})$ in the far field as
2960: 
2961: \begin{equation}
2962: \sigma_c = \frac{\lambda z}{2\pi \sigma} ~.\label{eqdimboh}
2963: \end{equation}
2964: %
2965: Here $\sigma$ is, as usual, the rms dimension of the electron
2966: beam. These few last remarks help to clarify what is the size of
2967: the source in the VCZ theorem, that is far from being a trivial
2968: question. In several papers \cite{YAB1,PFEI} the  rms electron
2969: beam size is recovered from the measurement of the transverse
2970: coherence length and subsequent application of the VCZ theorem,
2971: under the assumptions that the VCZ theorem can indeed be applied.
2972: In this regard, in \cite{YAB1} Section V, one may find a statement
2973: according to which the rms electron beam size "is only the average
2974: value along the undulator" because "the beta function has a large
2975: variation along the undulator". Another example dealing with the
2976: same issue is given in reference \cite{PFEI}. This paper (as well
2977: as reference \cite{YAB1}) reports experimental results. However,
2978: authors of \cite{PFEI} observe a disagreement between the electron
2979: beam rms size reconstructed from the VCZ theorem and beam
2980: diagnostics result of about a factor $2$. They ascribe this
2981: variation to the variation of the electron beam size along the
2982: undulator. In footnote 25 of reference \cite{PFEI}, one may read:
2983: "The precise shape and width of the x-ray intensity distribution
2984: in the source plane are directly connected to the properties of
2985: the electron beam. It would not be surprising if the limited depth
2986: of focus of the parabolically shaped electron beta function in the
2987: undulator translates into a virtually enlarged x-ray source
2988: size.". At first glance it looks like if the SR source had a
2989: finite longitudinal dimension, and the virtual source size
2990: depended on variations of the beta function along the undulator.
2991: However, as we have seen before, the concept of virtual source
2992: involves a single transverse plane, and in the most general case
2993: any variation of the beta function does not affect the virtual
2994: source size. In our case of quasi-homogeneous Gaussian source, the
2995: virtual source is located where the beta function is minimal, and
2996: its size coincides with the transverse size of the electron beam
2997: at that location.
2998: 
2999: Finally, it should be remarked that the VCZ theorem can only be
3000: applied to quasi-homogeneous sources. Consider for example
3001: \cite{YAB1,YAB2}, where a characterization of the vertical
3002: emittance in Spring-8 is reported, based on the measurement of the
3003: X-Ray beam coherence length in the far zone. The experiment was
3004: performed at the beamline BL29XU. Based on the assumption of
3005: validity of the VCZ theorem, it was found that the rms electron
3006: beam size at the undulator center (corresponding to the minimal
3007: value of the beta function) was $s_y \simeq 4.5 ~\mu$m, and that
3008: the coupling factor between horizontal and vertical emittance was
3009: down to the value $\zeta \simeq 0.12 \%$, which corresponds to an
3010: extremely small vertical emittance $\epsilon_y = 3.6$
3011: pm$\cdot$rad. A resolution limit of this method was also
3012: discussed, based on numerical calculations of the radiation size
3013: from a single electron performed at the exit of the undulator,
3014: $s_p \simeq 1.6 ~\mu$m at $E_p = 14.41 $ keV for the $4.5$ m long
3015: undulator used in the experiment. The resolution limit of the
3016: measurement of $s_y$ was estimated to be about $1~ \mu$m.
3017: Considering propagation laws for Gaussian beams it seems
3018: reasonable that the radiation spot size at the virtual source,
3019: located in the center of the undulator, be smaller than $s_p$.
3020: However, undulator radiation from a single electron cannot be
3021: considered a Gaussian beam. In particular, under the resonance
3022: approximation the field at the exit of the undulator exhibits a
3023: singularity and is not suitable for evaluation of the radiation
3024: spot size.
3025: 
3026: Use of $s_p$ led to an underestimation of the virtual source size
3027: and, thus, an overestimation of the resolution. Let us show this
3028: fact. Based on Eq. (\ref{undurad5gg}), we can determine the
3029: correct virtual source size of undulator radiation from a single
3030: electron. Let us consider the case when the single electron is
3031: emitting photons at the fundamental harmonic with energy $E =
3032: 14.41$ keV. The angular frequency of light oscillations is given,
3033: in this case, by $\omega = 2.2 \cdot 10^{19}$ Hz. For an undulator
3034: length $L_w = 4.5$ m, the normalization factor for the transverse
3035: size, $(L_w c/\omega)^{1/2}$, is about $8~\mu$m. From Fig.
3036: \ref{psio} obtain the dimensionless Half Width Half Maximum (HWHM)
3037: radiation size from a single electron (i.e. the HWHM width of the
3038: intensity distribution at the virtual source, located at the
3039: center of the undulator). This HWHM dimensionless value is about
3040: $0.7$. It follows that the HWHM value of the radiation spot size
3041: from a single electron is about $0.7 \cdot (c L_w/\omega)^{1/2}
3042: \simeq 6~ \mu$m. Therefore, the resolution in \cite{YAB1,YAB2}, is
3043: estimated to be better than the correct value.
3044: 
3045: Note that the HWHM radiation spot size from a \textit{single
3046: electron} is larger than the \textit{rms electron beam size} $s_y
3047: \simeq 4.5 ~ \mu$m found by means of coherence measurements. One
3048: concludes that the uncertainty due to finite resolution is larger
3049: than the measured electron beam size. This suggests that the
3050: analysis of experimental results \cite{YAB1,YAB2} might include a
3051: logical flaw. Authors of that reference assume the validity of the
3052: VCZ theorem in the vertical direction. If one assumes their result
3053: of a vertical emittance ${\epsilon}_y \simeq 0.3 \lambda/(2 \pi)$,
3054: it follows \textit{a posteriori} that the VCZ theorem could not
3055: have been applied in first instance (in this experiment the value
3056: of the beta function was ${\beta} \simeq L_w$). We suggest that
3057: analysis of experimental results in \cite{YAB1,YAB2} should be
3058: based, instead, on the study of transverse coherence for
3059: non-homogeneous undulator sources in free space made in the
3060: previous Section \ref{sec:main}.
3061: 
3062: 
3063: 
3064: 
3065: \subsection{\label{sub:nong} Non-Gaussian undulator sources}
3066: 
3067: Let us now turn to the analysis of non-Gaussian quasi-homogeneous
3068: sources on the basis of Eq. (\ref{Gnor3y}) and Eq. (\ref{resu1}).
3069: We still assume that $N_x \gg 1$ and $D_x \gg 1$. Moreover, as
3070: before, we assume that the minimal beta function is located at the
3071: undulator center.
3072: 
3073: \subsubsection{Case of a large electron-beam size $N_y \gg 1$ and $D_y
3074: \lesssim 1$.}
3075: 
3076: Eq. (\ref{Gnor3y}) simplifies to
3077: 
3078: \begin{eqnarray}
3079: \hat{G}_y\left(0,{\bar{y}},\Delta {y}\right) &=&
3080: \sqrt{\frac{\pi}{N_y}}  \exp \left[-\frac{(\Delta
3081: y)^2D_y}{2}\right] \exp\left[-\frac{\bar{y}^2}{2 N_y}\right]
3082: \gamma_S(\Delta y)~,\cr &&\label{nong10}
3083: \end{eqnarray}
3084: %
3085: where we defined the universal function $\gamma_S(\alpha)$ as
3086: 
3087: \begin{eqnarray}
3088: \gamma_S(\alpha)&=&\frac{1}{2}\int_{-\infty}^{\infty} d \phi_y
3089: \int_{-\infty}^{\infty} d \phi_x
3090: \Psi_0\left\{\left[{\phi_{x}}^2+\left({\phi_{y}}+\frac{\alpha}{2}
3091: \right)^2 \right]^{1/2}\right\}
3092: \Psi_0\left\{\left[{\phi_{x}}^2+\left({\phi_{y}}-\frac{\alpha}{2}
3093: \right)^2 \right]^{1/2}\right\}\cr && = \frac{1}{2 \mathcal{K}_S}
3094: \int_{-\infty}^{\infty} d\phi_y \mathcal{S}(\phi_y,\alpha) ~.
3095: \label{gammadef}
3096: \end{eqnarray}
3097: %
3098: 
3099: \begin{figure}
3100: \begin{center}
3101: \includegraphics*[width=140mm]{figure17.eps}% Here is how to import EPS art
3102: \caption{\label{gammas} Plot of the universal function
3103: $\gamma_S(\alpha)$, used to calculate the cross-spectral density
3104: of a quasi-homogeneous source when ${N}_x \gg1$, ${D}_x \gg 1$,
3105: ${N}_y \gg 1$ and ${D}_y \lesssim 1$.}
3106: \end{center}
3107: \end{figure}
3108: %
3109: 
3110: A plot of $\gamma_S(\alpha)$ is given in Fig. \ref{gammas}. Main
3111: features of $\gamma_S$ are a strong non-Gaussian shape, and the
3112: fact that also negative values are assumed. Eq. (\ref{introhdopo})
3113: is satisfied. Moreover,
3114: 
3115: \begin{eqnarray}
3116: \hat{I}_y\left(0,{\bar{y}}\right) &=& \sqrt{\frac{\pi}{ N_y}}
3117: \exp\left[-\frac{\bar{y}^2}{2 N_y}\right] ~,\cr &&\label{nong10In}
3118: \end{eqnarray}
3119: %
3120: and
3121: 
3122: \begin{eqnarray}
3123: \hat{g}_y\left(0,\Delta {y}\right) &=& \exp \left[-\frac{(\Delta
3124: y)^2D_y}{2}\right] \gamma_S(\Delta y)~,\cr &&\label{nong10de}
3125: \end{eqnarray}
3126: %
3127: In the far-zone, Eq. (\ref{resu1}) simplifies to
3128: 
3129: \begin{eqnarray}
3130: \hat{G}_y(\hat{z},\bar{\theta}_y,\Delta {\theta}_y) &=&
3131: \frac{1}{2\pi \hat{z} \sqrt{\pi D_y} \mathcal{K}_F} {\exp{\left[i
3132: \hat{z} \bar{\theta}_y \Delta {\theta}_y \right]}} \exp{\left[-
3133: \frac{N_y \Delta {\theta}_y^2}{2} \right]}\cr &&\times
3134: \int_{-\infty}^{\infty} d \hat{\phi}_y
3135: \exp{\left[-\frac{(\bar{\theta}_y+\hat{\phi}_y)^2}{2 D_y}\right]}
3136: \mathcal{I}_F(\phi_y) ~,\label{nong1f}
3137: \end{eqnarray}
3138: %
3139: where $\mathcal{I}_F$ has already been defined in Eq.
3140: (\ref{ISnorm}).  Moreover,
3141: 
3142: 
3143: \begin{eqnarray}
3144: \hat{I}_y(\hat{z},\bar{\theta}_y) &=& \frac{1}{2\pi \hat{z}
3145: \sqrt{\pi D_y} \mathcal{K}_F} \int_{-\infty}^{\infty} d
3146: \hat{\phi}_y \exp{\left[-\frac{(\bar{\theta}_y+\hat{\phi}_y)^2}{2
3147: D_y}\right]} \mathcal{I}_F(\phi_y) ~,\label{nong1fIn}
3148: \end{eqnarray}
3149: %
3150: and
3151: 
3152: 
3153: \begin{eqnarray}
3154: \hat{g}_y(\hat{z},\Delta {\theta}_y) &=& {\exp{\left[i \hat{z}
3155: \bar{\theta}_y \Delta {\theta}_y \right]}} \exp{\left[- \frac{N_y
3156: \Delta {\theta}_y^2}{2} \right]} ~\label{nong1fde}
3157: \end{eqnarray}
3158: %
3159: Note that, due to geometrical phase factor in Eq. (\ref{nong1f}),
3160: only the virtual source at $\hat{z}=0$ constitutes a
3161: quasi-homogeneous source.
3162: 
3163: Also, $\gamma_S$ and $\mathcal{I}_F$ basically form a Fourier
3164: pair:
3165: 
3166: \begin{eqnarray}
3167: \gamma_S(\Delta {y}) = \frac{1}{{2}\pi^2 \mathcal{K}_F}
3168: \int_{-\infty}^{\infty} d {\phi}_y \exp{\left[i \Delta {y}
3169: {\phi}_y\right]} \mathcal{I}_F({\phi}_y)~,\label{G2D3lastlastno2}
3170: \end{eqnarray}
3171: %
3172: meaning that the inverse VCZ theorem is satisfied for $N_{x,y} \ll
3173: \hat{z}^2$.
3174: 
3175: Finally, it is possible to calculate $\gamma_S$ analytically. To
3176: this purpose, it is sufficient to note that the Fourier transform
3177: 
3178: \begin{eqnarray}
3179: {\widetilde{\gamma}}_S(\xi,\eta) = \int_{-\infty}^{\infty} d
3180: \hat{\phi}_x \int_{-\infty}^{\infty} d \hat{\phi}_y
3181:  \exp{\left[i(\xi \hat{\phi}_x+\eta \hat{\phi}_y)\right]}
3182: \mathrm{sinc}^2\left(\frac{\hat{\phi}_x^2+\hat{\phi}_y^2}{4}\right)
3183: ~\label{G2D3lastlastno3}
3184: \end{eqnarray}
3185: %
3186: can be evaluated with the help of the Bessel-Fourier formula as
3187: 
3188: \begin{eqnarray}
3189: {\widetilde{\gamma}}_S(\lambda) &=& 2\pi \int_{0}^{\infty} d
3190: {\phi} ~{\phi} J_0\left({\phi} \lambda\right)
3191: \mathrm{sinc}^2\left(\frac{{\phi}^2}{4}\right)  \cr &=& 2\pi
3192: \left[\pi+ \lambda^2 \mathrm{Ci}\left(\frac{\lambda^2}{2}\right)-
3193: 2 \sin\left(\frac{\lambda^2}{2}\right)- 2
3194: \mathrm{Si}\left(\frac{\lambda^2}{2}\right)
3195: \right]~,\label{G2D3lastlastno4}\end{eqnarray}
3196: %
3197: where $\lambda^2 = \xi^2+\eta^2$, $\phi^2 = \phi_x^2+\phi_y^2$,
3198: $\mathrm{Si}(\cdot)$ is the sine integral function and
3199: $\mathrm{Ci}(\cdot)$ is the cosine integral function. Thus,
3200: letting $\xi=0$ and $\eta = \Delta y$ one has
3201: 
3202: 
3203: \begin{eqnarray}
3204: \gamma_S(\Delta {y}) =\frac{1}{\pi} \left[\pi+ (\Delta y)^2
3205: \mathrm{Ci}\left(\frac{(\Delta y)^2}{2}\right)- 2
3206: \sin\left(\frac{(\Delta y)^2}{2}\right)- 2
3207: \mathrm{Si}\left(\frac{(\Delta y)^2}{2}\right)
3208: \right]~.\label{G2D3lastlastno5}
3209: \end{eqnarray}
3210: %
3211: 
3212: \subsubsection{Case of a large electron-beam divergence $D_y \gg 1$ and $N_y \lesssim 1$.}
3213: 
3214: A similar analysis can be given in the case of a large
3215: electron-beam divergence. In this case Eq. (\ref{Gnor3y})
3216: simplifies to give
3217: 
3218: 
3219: \begin{eqnarray}
3220: \hat{G}_y\left(0,{\bar{y}},\Delta {y}\right) &=&
3221: \frac{1}{2\mathcal{K}_S} \sqrt{\frac{\pi}{N_y}}  \exp
3222: \left[-\frac{(\Delta y)^2D_y}{2}\right] \int_{-\infty}^{\infty} d
3223: \phi_y \exp\left[-\frac{\left(\phi_y+\bar{y}\right)^2}{2
3224: N_y}\right] \mathcal{I}_S(\phi_y)~,\cr &&\label{nong20}
3225: \end{eqnarray}
3226: %
3227: where $\mathcal{I}_S$ has already been defined in Eq. (\ref{B2}).
3228: Eq. (\ref{introhdopo}) is satisfied. Moreover,
3229: 
3230: 
3231: \begin{eqnarray}
3232: \hat{I}_y\left(0,{\bar{y}}\right) &=& \frac{1}{2 \mathcal{K}_S
3233: }\sqrt{\frac{\pi}{N_y}} \int_{-\infty}^{\infty} d \phi_y
3234: \exp\left[-\frac{\left(\phi_y+\bar{y}\right)^2}{2 N_y}\right]
3235: \mathcal{I}_S(\phi_y)~,\cr &&\label{nong20In}
3236: \end{eqnarray}
3237: %
3238: 
3239: \begin{eqnarray}
3240: \hat{g}_y\left(0,\Delta {y}\right) &=& \exp \left[-\frac{(\Delta
3241: y)^2D_y}{2}\right]~.\cr &&\label{nong20de}
3242: \end{eqnarray}
3243: %
3244: In the far zone, Eq. (\ref{resu1}) simplifies to
3245: 
3246: \begin{eqnarray}
3247: \hat{G}(\hat{z},\bar{\theta}_y,\Delta {\theta}_y) &=&
3248: \frac{1}{\hat{z}} \sqrt{\frac{\pi}{D_y}} {\exp{\left[i \hat{z}
3249: \bar{\theta}_y \Delta {\theta}_y \right]}} \exp{\left[- \frac{N_y
3250: \Delta {\theta}_y^2}{2} \right]}
3251: \exp{\left[-\frac{\bar{\theta}_y^2}{2 D_y}\right]}\gamma_F(\Delta
3252: \theta_y) ~,\cr &&\label{nong2f}
3253: \end{eqnarray}
3254: %
3255: where
3256: 
3257: \begin{eqnarray}
3258: \gamma_F(\alpha) &=& \frac{1}{2\pi^2} \int_{-\infty}^{\infty} d
3259: {\phi}_y \int_{-\infty}^{\infty} d {\phi}_x \cr && \times
3260: \Psi_f\left\{\left[{{\phi}_x^2
3261: +\left(\phi_y-\frac{\alpha}{2}\right)^2}\right]^{1/2}\right\}
3262: \Psi_f\left\{\left[{{\phi}_x^2
3263: +\left(\phi_y+\frac{\alpha}{2}\right)^2}\right]^{1/2}\right\} \cr
3264: && = \frac{1}{2\pi^2 \mathcal{K}_F}\int_{-\infty}^{\infty} d
3265: \phi_y \mathcal{F}(\phi_y,\alpha) ~.\label{betadef}
3266: \end{eqnarray}
3267: %
3268: Moreover
3269: 
3270: 
3271: \begin{eqnarray}
3272: \hat{I}_y(\hat{z},\bar{\theta}_y) &=& \frac{1}{\hat{z}}
3273: \sqrt{\frac{\pi}{D_y}} \exp{\left[-\frac{\bar{\theta}_y^2}{2
3274: D_y}\right]} ~ ,\cr &&\label{nong2fIn}
3275: \end{eqnarray}
3276: %
3277: and
3278: 
3279: \begin{eqnarray}
3280: \hat{g}(\hat{z},\Delta {\theta}_y) &=& {\exp{\left[i \hat{z}
3281: \bar{\theta}_y \Delta {\theta}_y \right]}} \exp{\left[- \frac{N_y
3282: \Delta {\theta}_y^2}{2} \right]} \gamma_F(\Delta \theta_y) ,\cr
3283: &&\label{nong2fde}
3284: \end{eqnarray}
3285: %
3286: 
3287: 
3288: 
3289: \begin{figure}
3290: \begin{center}
3291: \includegraphics*[width=140mm]{figure18.eps}% Here is how to import EPS art
3292: \caption{\label{gammaf} Plot of the universal function $\gamma_F$,
3293: used to calculate the cross-spectral density in the far zone when
3294: $\hat{N}_x \gg1$, $\hat{D}_x \gg 1$, $\hat{N}_y \lesssim 1$ and
3295: $\hat{D}_y \gg 1$.}
3296: \end{center}
3297: \end{figure}
3298: %
3299: 
3300: A plot of $\gamma_F(\alpha)$ is given in Fig. \ref{gammaf}. As for
3301: $\gamma_S$, main features of $\gamma_F$ are a strong non-Gaussian
3302: shape, and the fact that also negative values are assumed. Due to
3303: the phase factors in Eq. (\ref{nong2f}), only the virtual source
3304: at $\hat{z}=0$ constitutes a quasi-homogeneous source.
3305: 
3306: Also, $\mathcal{I}_S$ and $\gamma_F$ form a Fourier pair:
3307: 
3308: \begin{eqnarray}
3309: {\mathcal{I}_S}(\bar{y}) = \frac{\mathcal{K}_S}{\pi}
3310: \int_{-\infty}^{\infty} d {\phi}_y \exp\left[-i \bar{y}
3311: {\phi}_y\right] \gamma_F({\phi}_y) ~,\label{BB}
3312: \end{eqnarray}
3313: %
3314: meaning that the VCZ theorem is satisfied for $\hat{z}^2 D_{x,y}
3315: \gg 1$. Since $D_{x,y} \gg 1$ this means that, in this case, the
3316: far zone begins at the very exit of the undulator, at $\hat{z}
3317: \sim 1$.
3318: 
3319: Finally, starting with the representation of $\mathcal{I}_S$ in
3320: Eq. (\ref{B2}), we can write (see \cite{GOOD} Appendix A.3.)
3321: 
3322: 
3323: \begin{eqnarray}
3324: {\mathcal{I}_S}(\bar{y}) = \mathcal{K}_S \int_{-\infty}^{\infty} d
3325: {\phi}_y \exp\left[-i \bar{y} {\phi}_y\right] \int_{0}^{\infty} d
3326: \alpha ~ \alpha J_0(\alpha \phi_y) \Psi_0^2(\alpha) ~.\label{BBQ1}
3327: \end{eqnarray}
3328: %
3329: Comparison with Eq. (\ref{BB}) yields the following alternative
3330: representation of $\gamma_F$ in terms of a one-dimensional
3331: integral involving special functions:
3332: 
3333: \begin{eqnarray}
3334: \gamma_F(\phi_y) = \pi \int_{0}^{\infty} d \alpha ~ \alpha
3335: J_0(\alpha \phi_y) \Psi_0^2(\alpha) ~.\label{BBQ2}
3336: \end{eqnarray}
3337: %
3338: 
3339: 
3340: 
3341: 
3342: \subsection{\label{sub:accu} Accuracy of the quasi-homogeneous approximation}
3343: 
3344: As we have discussed before, the quasi-homogeneous approximation
3345: can be applied when either or both $N_y \gg 1$ or $D_y \gg 1$. In
3346: this Section we will see that the accuracy of the
3347: quasi-homogeneous approximation scales as $(\sqrt{\max[N_y,1]
3348: \max[D_y,1]}~)^{-1}$. As we have previously seen, in the
3349: particular case when both $N_y \gg 1$ and $D_y \gg 1$, a Gaussian
3350: model may be used. We will see that the accuracy of such model is
3351: worse than that of the quasi-homogeneous model, and scales as
3352: $\max(1/\sqrt{D_y},1/\sqrt{N_y})$.
3353: 
3354: We begin demonstrating that, when the source is quasi-homogeneous,
3355: we may take the approximation $\hat{\mathcal{S}}(\Delta {y},
3356: \hat{\phi}) \simeq \gamma_S(\Delta {y}) \mathcal{I}_S(\hat{\phi})$
3357: in Eq. (\ref{Gnor4y}) with an accuracy scaling as
3358: $(\sqrt{\max[N_y,1] \max[D_y,1]}~)^{-1}$. First, let us introduce
3359: a normalized version of the one-dimensional inverse Fourier
3360: transform of the function $\mathcal{F}$, that is
3361: 
3362: \begin{eqnarray}
3363: \bar{\mathcal{F}}(u,v) &=& \frac{1}{2\pi^2
3364: \mathcal{K}_F}\int_{-\infty}^{\infty}
3365: \mathcal{F}\left(\alpha,v\right) \exp\left[2 i u \alpha \right] d
3366: \alpha = \frac{1}{2\mathcal{K}_S}\int_{-\infty}^{\infty}
3367: {\mathcal{S}}\left(u,\delta\right) \exp\left[-i v\frac{\delta}{2}
3368: \right] d \delta,\cr &&\label{WIG2}
3369: \end{eqnarray}
3370: %
3371: where the normalization factor is chosen in such as way that
3372: $\bar{\mathcal{F}}(0,0)=1$. The cross-spectral density  in Eq.
3373: (\ref{Gnor4y}) can therefore be written as
3374: 
3375: \begin{eqnarray}
3376: \hat{G}_y(0,\bar{y},\Delta {y}) &=&
3377: \frac{1}{4\sqrt{2}\pi^2}\exp\left[- \frac{{D}_y (\Delta {y}^2)}{2}
3378: \right] \cr && \times \int_{-\infty}^{\infty} d {u} \exp\left[-i
3379: \frac{{u}\bar{y}}{2}\right] \exp\left[-\frac{{N}_y
3380: {u}^2}{2}\right] \bar{\mathcal{F}} (\Delta {y}, {u})~,\cr
3381: &&\label{gen2sepa}
3382: \end{eqnarray}
3383: %
3384: having used the convolution theorem. Under the quasi-homogeneous
3385: assumption, we can approximate $\bar{\mathcal{F}} (\Delta {y},
3386: {u}) \simeq \bar{\mathcal{F}} (\Delta {y}, 0)\bar{\mathcal{F}} (0,
3387: {u})$. To show this, let us represent $\bar{\mathcal{F}}(u,v)$
3388: using a Taylor expansion around the point $(0,0)$. One obtains
3389: 
3390: \begin{eqnarray}
3391: \bar{\mathcal{F}}(u,v) &=& 1 + \sum_{k=1}^{\infty} \frac{1}{k!}
3392: \left[u^k \frac{\partial^k \bar{\mathcal{F}}(u,0)}{\partial
3393: u^k}\Bigg|_{u=0}+v^k \frac{\partial^k
3394: \bar{\mathcal{F}}(0,v)}{\partial v^k}\Bigg|_{v=0}\right] \cr && +
3395: O(uv) ~, \label{Mpexp}
3396: \end{eqnarray}
3397: %
3398: where the normalization relation $\bar{\mathcal{F}}(0,0)=1$ has
3399: been taken advantage of. Similarly, one may consider the following
3400: representation of the product $\bar{\mathcal{F}}(u,0)$
3401: $\bar{\mathcal{F}}(0,v)$ also obtained by means of a Taylor
3402: expansion:
3403: 
3404: \begin{eqnarray}
3405: \bar{\mathcal{F}}(u,0)\bar{\mathcal{F}}(0,v) &=&
3406: \left[\bar{\mathcal{F}}(0,0)+ \sum_{k=1}^{\infty} \frac{u^k}{k!}
3407: \frac{d^k \bar{\mathcal{F}}(u,0)}{du^k}\Bigg|_{u=0}\right]\cr
3408: &\times& \left[\bar{\mathcal{F}}(0,0)+ \sum_{j=1}^{\infty}
3409: \frac{v^j}{j!} \frac{d^j
3410: \bar{\mathcal{F}}(0,v)}{dv^j}\Bigg|_{v=0}\right] \cr &=& 1 +
3411: \sum_{n=1}^{\infty} \frac{1}{n!} \left[u^n \frac{d^n
3412: \bar{\mathcal{F}}(u,0)}{du^n}\Bigg|_{u=0} \right.\cr && \left.
3413: +v^n \frac{d^n \bar{\mathcal{F}}(0,v)}{dv^n}\Bigg|_{v=0}\right] +
3414: O(uv) ~,\label{Mpexp2}
3415: \end{eqnarray}
3416: %
3417: having used $\bar{\mathcal{F}}(0,0)=1$. Comparison of the last
3418: equality in (\ref{Mpexp2}) with the right hand side of Eq.
3419: (\ref{Mpexp}) shows that $\bar{\mathcal{F}}(u,v)\simeq
3420: \bar{\mathcal{F}}(u,0)\bar{\mathcal{F}}(0,v)$ up to corrections of
3421: order $uv \sim 1/\sqrt{\max[{N}_y,1]\max[{D}_y,1]}$, that is the
3422: quasi-homogeneous accuracy. Using this approximation in Eq.
3423: (\ref{gen2sepa}) yields
3424: 
3425: \begin{eqnarray}
3426: \hat{G}(0,\bar{y},\Delta {y}) &=& \frac{1}{4\sqrt{2}\pi^2}
3427: \exp\left[-\frac{{D}_y \Delta {y}^2}{2} \right] \bar{\mathcal{F}}
3428: (\Delta {y}, 0) \cr &&\times \int_{-\infty}^{\infty} d {u}
3429: \exp\left[-i \frac{{u}\bar{y}}{2}\right] \exp\left[-\frac{ {N}_y
3430: {u}^2}{2}\right] \bar{\mathcal{F}} (0,{u})~.\label{gen2sepabis}
3431: \end{eqnarray}
3432: %
3433: Finally, recalling the definitions of $\gamma_S$ and
3434: $\mathcal{I}_S$ we can write Eq. (\ref{gen2sepabis}) as
3435: 
3436: \begin{eqnarray}
3437: \hat{G}(0,\bar{y},\Delta {y}) &=& \sqrt{\frac{\pi}{N_y
3438: }}\frac{1}{2\mathcal{K}_S }\exp\left[-\frac{{D}_y \Delta {y}^2}{2}
3439: \right] \gamma_S(\Delta {y}) \cr && \times \int_{-\infty}^{\infty}
3440: d {\phi} \exp\left[-\frac{(\bar{y}+{\phi})^2}
3441: {2{N}_y}\right]\mathcal{I}_S({\phi})~.\label{gen2qhapp}
3442: \end{eqnarray}
3443: %
3444: Eq. (\ref{gen2qhapp}) is valid in any quasi-homogeneous case.
3445: 
3446: Note that Eq. (\ref{gen2qhapp}) accounts for diffraction effects
3447: through the universal functions $\gamma_S$ and $\mathcal{I}_S$.
3448: This may be traced back to the use of the inhomogeneous wave
3449: equation to calculate the cross-spectral density for the virtual
3450: source, from which Eq. (\ref{gen2qhapp}) follows. Deriving Eq.
3451: (\ref{gen2qhapp}), we assume a large number of modes, and this
3452: justifies the use of phase space representation as an alternative
3453: characterization of the source, in place of the cross-spectral
3454: density (i.e. Eq. (\ref{gen2qhapp}) itself).
3455: 
3456: Setting $\Delta {y}=0$, Eq. (\ref{gen2qhapp}) gives the exact
3457: intensity distribution at the virtual source, i.e. Eq.
3458: (\ref{Inty0}). The spectral degree of coherence on the virtual
3459: source is then recovered using the definition of quasi-homogeneous
3460: source $\hat{G} = \hat{I}(\bar{y}) g(\Delta {y})$. Since the
3461: source is quasi-homogeneous, the Fourier transform of the spectral
3462: degree of coherence $g(\Delta {y})$ yields the intensity in the
3463: far zone. Remembering that $\gamma_S$ and $\mathcal{I}_F$ form a
3464: Fourier pair, we conclude that, starting from Eq.
3465: (\ref{gen2qhapp}) it is possible to reproduce the exact result for
3466: the intensity in the far zone, Eq. (\ref{Intyf}). Quite
3467: remarkably, Eq. (\ref{gen2qhapp}), which is derived under the
3468: quasi-homogeneous approximation and is related to an accuracy
3469: $1/\sqrt{\max[{N}_y,1]\max[{D}_y,1]}$, yields back two results,
3470: Eq. (\ref{Inty0}) and Eq. (\ref{Intyf}) which are valid regardless
3471: the fact that the source is quasi-homogeneous or not.
3472: 
3473: Let us now consider the quasi-homogeneous case when both
3474: $\hat{N}_y \gg 1$ and $\hat{D}_y \gg 1$. Worsening the accuracy in
3475: the calculation of the cross-spectral density of the source, we
3476: may reduce Eq. (\ref{gen2qhapp}) to
3477: 
3478: \begin{eqnarray}
3479: \hat{G}_y\left(0,{\bar{y}},\Delta {y}\right) &=&
3480: \sqrt{\frac{\pi}{{N_y}}} \exp \left[-\frac{(\Delta y)^2D_y
3481: }{2}\right] \exp\left[-\frac{\bar{y}^2}{2 N_y}\right]
3482: ~,\label{gen2qhappworse}
3483: \end{eqnarray}
3484: %
3485: that is Eq. (\ref{Gnorgauss0}). Note that neglecting the product
3486: with the $\gamma_S$ function can be done with an accuracy
3487: $1/\sqrt{{D}_y}$, while extraction of the exponential function in
3488: $\bar{y}$ from the convolution product with the $\mathcal{I}_S$
3489: function can be done with an accuracy $1/\sqrt{{N}_y}$. In our
3490: study case when ${D}_y \gg 1$ and ${N}_y \gg 1$, the overall
3491: accuracy of Eq. (\ref{gen2qhappworse})  can be estimated as
3492: $\max(1/\sqrt{{D}_y},1/\sqrt{{N}_y})$, that is the accuracy of the
3493: Gaussian approximation. Such accuracy is much worse than that of
3494: the quasi-homogeneous assumption in Eq. (\ref{gen2qhapp}), that is
3495: $1/\sqrt{{N}_y {D}_y}$.
3496: 
3497: When ${N}_y \gg 1$ and ${D}_y \simeq 1$ the accuracy of the
3498: quasi-homogeneous approximation becomes $1/\sqrt{{N}_y
3499: \max[1,{D}_y]}$. When ${N}_y \simeq 1$ and ${D}_y \gg 1$ it
3500: becomes, instead, $1/\sqrt{\max[1,{N}_y]\hat{D}_y }$. In these
3501: cases, the accuracy of the quasi-homogeneous approximation is
3502: comparable to the accuracy of the Gaussian approximation. To be
3503: specific, when ${N}_y \gg 1$ and ${D}_y \simeq 1$ the accuracy of
3504: the quasi-homogeneous approximation is of order $1/\sqrt{{N}_y}$
3505: and Eq. (\ref{gen2qhapp}) can be substituted with
3506: 
3507: \begin{eqnarray}
3508: \hat{G}(0,\bar{y},\Delta {y}) &=& \sqrt{\frac{\pi}{N_y
3509: }}\exp\left[-\frac{\bar{y}^2}{2{N}_y}\right]\exp\left[-\frac{{D}_y
3510: \Delta {y}^2}{2} \right] \gamma_S(\Delta {y})~.\label{gen2qhappN}
3511: \end{eqnarray}
3512: %
3513: without loss of accuracy, because the relative accuracy of the
3514: convolution is of order $1/\sqrt{{N}_y}$ as the accuracy of the
3515: quasi-homogenous approximation. Eq. (\ref{gen2qhappN}) is just Eq.
3516: (\ref{nong10}). A similar reasoning can be done when ${D}_y \gg 1$
3517: and ${N}_y \simeq 1$. In this case the accuracy of the
3518: quasi-homogeneous approximation is of order $1/\sqrt{{D}_y}$, and
3519: Eq. (\ref{gen2qhapp}) can be substituted with
3520: 
3521: \begin{eqnarray}
3522: \hat{G}(0,\bar{y},\Delta {y}) &=& \sqrt{\frac{\pi}{N_y
3523: }}\frac{1}{2\mathcal{K}_S }\exp\left[-\frac{{D}_y \Delta {y}^2}{2}
3524: \right]\int_{-\infty}^{\infty} d {\phi}
3525: \exp\left[-\frac{(\bar{y}+{\phi})^2}
3526: {2{N}_y}\right]\mathcal{I}_S({\phi})~.\label{gen2qhappD}
3527: \end{eqnarray}
3528: %
3529: without loss of accuracy. Eq. (\ref{gen2qhappD}) is just Eq.
3530: (\ref{nong20}).
3531: 
3532: 
3533: 
3534: 
3535: \subsection{\label{sub:geop} Quasi-homogeneous sources in terms of phase space}
3536: 
3537: 
3538: The cross-spectral density at the virtual source can be written as
3539: in Eq. (\ref{introhdopo}) that we rewrite here for convenience in
3540: terms of coordinates $\bar{r}_{x,y}$ and $\Delta {{r}_{x,y}}$:
3541: 
3542: 
3543: \begin{eqnarray}
3544: \hat{G}_o({\bar{r}_x},\bar{r}_y,\Delta {{r}_x},\Delta {{r}_y}) =
3545: \hat{I}\left(\bar{r}_x,\bar{r}_y\right) {g}(\Delta {{r}_x},\Delta
3546: {{r}_y}) ~.\label{introhdopob}
3547: \end{eqnarray}
3548: %
3549: For notational simplicity we use notation $\hat{G}_o$ to indicate
3550: $\hat{G}$ at $\hat{z}=0$. The Fourier transform of Eq.
3551: (\ref{introhdopob}) with respect to all variables will be
3552: indicated with
3553: 
3554: \begin{eqnarray}
3555: \hat{\mathcal{G}}_o( {\bar{\theta}_x}, {\bar{\theta}_y},{\Delta
3556: {\theta}_x},\Delta {\theta}_y) &=& \int_{-\infty}^{\infty} d
3557: \Delta { {r}_x'}\int_{-\infty}^{\infty}d \Delta { {r}_y'}
3558: \int_{-\infty}^{\infty} d { \bar{r}_x'}\int_{-\infty}^{\infty}d
3559:  { \bar{r}_y'} ~ \hat{G}_o({ \bar{r}_x'},{ \bar{r}_y'},
3560: \Delta{{r}'_x},\Delta{{r}'_y}) \cr &&\times\exp [2i (
3561: \bar{\theta}_x \Delta{{r}'_x}+\bar{\theta}_y \Delta{{r}'_y} )]\exp
3562: [i ( \Delta {\theta}_x { \bar{r}'_x}+\Delta {\theta}_y {
3563: \bar{r}'_y} )]~.\cr && \label{ftgdeffor}
3564: \end{eqnarray}
3565: %
3566: The two quantities
3567: $\hat{I}(\bar{r}_x,\bar{r}_y)=\hat{G}_o({\bar{r}_x},\bar{r}_y,0,0)$
3568: and $\hat{\Gamma}({\bar{\theta}_x}, {\bar{\theta}_y})
3569: =\hat{\mathcal{G}}_o( {\bar{\theta}_x}, {\bar{\theta}_y},0,0)$ are
3570: always positive because, by definition of $\hat{G}_o$, they are
3571: ensemble averages of quantities under square modulus.
3572: 
3573: Let us now introduce the Fourier transform of Eq.
3574: (\ref{introhdopo}) with respect to $\Delta {r_{x,y}}$:
3575: 
3576: \begin{eqnarray}
3577: \hat{\Phi}_o({\bar{r}_x},\bar{r}_y, {\bar{\theta}_x},
3578: {\bar{\theta}_y}) &=& \int_{-\infty}^{\infty} d \Delta {
3579: {r}_x'}\int_{-\infty}^{\infty}d \Delta {r_y'}~ \hat{G}_o({
3580: \bar{r}_x},{ \bar{r}_y}, \Delta{{r}'_x},\Delta{{r}'_y}) \cr
3581: &&\times\exp [i ( \bar{\theta}_x \Delta{ {r}'_x}+\bar{\theta}_y
3582: \Delta{{r}'_y} )]~. \label{wigdef}
3583: \end{eqnarray}
3584: %
3585: Accounting for Eq. (\ref{introhdopob}), i.e. in the particular
3586: case of a (virtual) quasi-homogeneous source, Eq. (\ref{wigdef})
3587: can be written as
3588: 
3589: \begin{eqnarray}
3590: \hat{\Phi}_o({\bar{r}_x},\bar{r}_y, {\bar{\theta}_x},
3591: {\bar{\theta}_y}) &=& \hat{I}\left(\bar{r}_x,\bar{r}_y\right)
3592: \hat{\Gamma}({\bar{\theta}_x}, {\bar{\theta}_y})~,
3593: \label{wigdefdef}
3594: \end{eqnarray}
3595: %
3596: having recognized that $\hat{\Gamma}({\bar{\theta}_x},
3597: {\bar{\theta}_y}) =\hat{\mathcal{G}}_o( {\bar{\theta}_x},
3598: {\bar{\theta}_y},0,0)$ is the Fourier transform of the spectral
3599: degree of coherence ${g}$. The distribution $\hat{\Phi}_o$, being
3600: the product of two positive quantities, never assumes negative
3601: values. Therefore it may always be interpreted as a phase space
3602: distribution\footnote{Physically, in the quasi-homogeneous case,
3603: $\hat{\Gamma}$ can be identified with the radiant intensity of the
3604: virtual source. This follows from a statement similar to the van
3605: Cittert-Zernike theorem for quasi-homogeneous sources (see
3606: \cite{MAND}). Note that the intensity and the Fourier transform of
3607: the spectral degree of coherence are obtained back from the phase
3608: space distribution, Eq. (\ref{wigdefdef}), by integration over
3609: coordinates $\bar{\theta}_{x,y}$ and $\bar{r}_{x,y}$
3610: respectively.}. This analysis shows that quasi-homogeneous sources
3611: can always be characterized in terms of Geometrical Optics. It
3612: also shows that, in this particular case, the coordinates in the
3613: phase space, $\bar{r}_{x,y}$ and $\bar{\theta}_{x,y}$, are
3614: separable.
3615: 
3616: Eq. (\ref{wigdef}) is the definition of a Wigner distribution. In
3617: the case of quasi-homogenous sources, as we have just seen, the
3618: Wigner distribution is never negative and, therefore, can always
3619: be interpreted as a phase space distribution.  In the case of non
3620: quasi-homogeneous sources one may still define a Wigner
3621: distribution using Eq. (\ref{wigdef}). However the Wigner function
3622: itself is not always a positive function. As a consequence it
3623: cannot always be interpreted as a phase space distribution. On the
3624: one hand, quasi-homogeneity is a sufficient condition for the
3625: Geometrical Optics approach to be possibly used in the
3626: representation of the source. On the other hand though, necessary
3627: and sufficient conditions for $\hat{\Phi}_o$ to be a positive
3628: function are more difficult to find.
3629: 
3630: 
3631: Note that in the case of quasi-homogeneous non-Gaussian sources
3632: one should account to diffraction effects when calculating source
3633: properties. However, using a Wigner function approach, one can
3634: still use a phase-space representation. In this case diffraction
3635: effects have the effect of complicating the structure of the
3636: phase-space describing the source. Also note that Gaussian-Schell
3637: model cannot be applied in all generality even to
3638: quasi-homogeneous sources, as it does not properly describe the
3639: case of non-Gaussian sources, which is most natural for third
3640: generation facilities in the vertical direction.
3641: 
3642: As we have seen before, third generation light sources are
3643: characterized, up to the VUV range, by ${N}_x \gg 1$ and ${D}_x
3644: \gg 1$, which implies factorization of the cross-spectral density.
3645: Thus, considering the vertical direction separately, equivalent
3646: condition for quasi-homogeneity requires that either (or both)
3647: ${N}_y \gg 1$ and ${D}_y \gg 1$.
3648: 
3649: An intuitive picture in the real space is given by a (virtual)
3650: quasi-homogeneous source with characteristic (normalized) square
3651: sizes of order $\max[{N}_{y},1]$ and $N_x$, and characteristic
3652: (normalized) correlation length square sizes of order
3653: $\min[1/{D}_{y},1]$ and $1/D_x$, in the vertical and horizontal
3654: directions. Since horizontal and the vertical directions can be
3655: treated separately, we have a large number of independently
3656: radiating sources given by the product
3657: 
3658: \begin{equation}
3659: M_{y}= \max[{N}_{y},1]\max[{D}_{y},1]~ \label{modesappr}
3660: \end{equation}
3661: %
3662: in the vertical direction, and
3663: 
3664: 
3665: \begin{equation}
3666: M_{x}= {N}_{x}{D}_{x}~ \label{modesappr2}
3667: \end{equation}
3668: %
3669: in the horizontal direction. The number $M_{x,y}$ is, in other
3670: words, an estimation of the number of coherent modes in the
3671: horizontal and in the vertical direction\footnote{This is in
3672: agreement with an intuitive picture where the photon-beam phase
3673: space reproduces the electron-beam phase space up to the limit
3674: imposed by the intrinsic diffraction of undulator radiation.
3675: Imagine to start from a situation with ${N}_{x,y}\gg 1$ and
3676: ${D}_{x,y} \gg 1$ and to "squeeze" the electron-beam phase space
3677: in the vertical direction by diminishing ${N}_{y}$ and ${D}_{y}$.
3678: On the one hand the characteristic sizes of the phase space of the
3679: electron beam are always of order ${N}_{y}$ and ${D}_{y}$. On the
3680: other hand the characteristic sizes of the phase space of the
3681: photon beam are of order $\max[{N}_{y},1]$ and $\max[{D}_{y},1]$:
3682: diffraction effects limit the "squeezing" of the phase space of
3683: the photon beam. }. The number $M_{x,y}^{-1}$ is the accuracy of
3684: Geometrical Optics results compared with Statistical Optics
3685: results or, better, the accuracy of the quasi-homogeneous
3686: assumption. It should be noted that, as $M_{x,y}$ approaches
3687: unity, the accuracy of the quasi-homogeneous assumption becomes
3688: worse and worse and $M_{x,y}$ cannot be taken anymore as a
3689: meaningful estimation of the number of modes: it should be
3690: replaced by a more accurate concept based on Statistical Optics.
3691: To complete the previous statement we should add that $M_{x,y}$
3692: completely loses the meaning of "number of modes" when Geometrical
3693: Optics cannot be applied. For instance when both ${N}_y$ and
3694: ${D}_y$ are of order unity (or smaller), one can state that the
3695: Geometrical Optics approach fails in the vertical direction
3696: because the phase space area is getting near to the uncertainty
3697: limit. In this case it is not possible to ascribe the meaning of
3698: "number of modes" to the number $M_y$ simply because the
3699: Geometrical Optics approach in the vertical direction fails.
3700: However, when ${N}_y$ and ${D}_y$ are of order unity (or smaller),
3701: but both ${N}_x \gg 1$ and ${D}_x \gg 1$, the cross-spectral
3702: density admits factorization in the horizontal and in the vertical
3703: direction and the source in the horizontal direction can be still
3704: described, independently, with the help of Geometrical Optics.
3705: 
3706: We should remark that Statistical Optics is the only mean to deal,
3707: in general, with the stochastic nature of SR. Only in particular
3708: cases SR can be treated in terms of Geometrical Optics, where
3709: planning of experiments can take advantage of ray-tracing code
3710: techniques. One of these cases is constituted by second generation
3711: light sources, because ${N}_{x,y} \gg 1$ and ${D}_{x,y} \gg 1$.
3712: 
3713: Description in terms of quasi-homogeneous sources is also
3714: important for third generation facilities in the case of bending
3715: magnet beamlines (surprisingly, in both horizontal and vertical
3716: plane), horizontal plane in undulator beamlines, but it cannot be
3717: applied with accuracy to future sources (e.g. ERL-based sources).
3718: 
3719: 
3720: Quasi-homogenous sources can be described in terms of geometrical
3721: optics, the outcome being equivalent to description in terms of
3722: statistical optics. In order to decide whether Geometrical Optics
3723: or Wave Optics is applicable, in all generality one should
3724: separately compare the \textit{photon beam} size and divergence
3725: with the radiation diffraction size and diffraction angle, which
3726: are quantities pertaining the single electron radiation. Let us
3727: fix a given direction $x$ or $y$. The square of the diffraction
3728: angle is defined by $(\sigma_d')^2 \sim \lambda/(2 \pi L_f)$,
3729: $L_f$ being the formation length of the radiation at wavelength
3730: $\lambda$. The diffraction size of the source is given by
3731: $\sigma_d \sim \sigma'_d L_f$. In calculating the photon beam size
3732: and divergence one should always include diffraction effects. As a
3733: result, if $\sigma^2$ and $(\sigma')^{2}$ indicate the square of
3734: the electron beam size and divergence, the corresponding square of
3735: the photon beam size and divergence will be respectively of order
3736: $\max[\sigma^2, \sigma_d^2]$ and $\max[(\sigma')^{2},
3737: (\sigma_d')^2]$. These quantities can be rewritten in terms of the
3738: electron beam emittance  as $\max[\epsilon \beta, \sigma_d^2]$ and
3739: $\max[\epsilon/\beta, (\sigma'_d)^2]$, $\beta$ being the beta
3740: function value at the virtual source position for the radiator
3741: (undulator, bending magnet, or other). Dividing these two
3742: quantities respectively by $\sigma_d^2$ and $(\sigma'_d)^2$  give
3743: natural values, normalized to unity, for the photon beam size
3744: $\max[2 \pi \epsilon \beta/(L_f \lambda), 1]$ and divergence
3745: $\max[2 \pi \epsilon L_f/ (\beta \lambda), 1]$. When the product
3746: between these two quantities is much larger than unity one can use
3747: a Geometrical Optics approach. In this case, this product
3748: represents the normalized \textit{photon beam} emittance. When
3749: $\beta \sim L_f$, as in many undulator cases, one may compare, for
3750: rough estimations, the electron beam emittance and the radiation
3751: wavelength as we have done before. However, in the case of a
3752: bending magnet one may typically have $\beta$ of order $10$ m and
3753: $L_f\simeq (\rho^2 \lambdabar)^{1/3}$ ($\rho$ being the bending
3754: radius) of order $10^{-3} \div 10^{-2} $ m. The ratio $\beta/L_f
3755: \gg 1$ now constitutes an extra large parameter of the problem. In
3756: this case, even if the electron beam emittance is two order of
3757: magnitude smaller than the wavelength, due to diffraction effects
3758: one can still apply a Geometrical Optics approach, because $\max[2
3759: \pi \epsilon \beta/(L_f \lambda), 1] \cdot \max[2 \pi \epsilon
3760: L_f/ (\beta \lambda), 1] \gg 1$, i.e. the photon beam emittance is
3761: much larger than the wavelength. As a result, dimensional analysis
3762: suggests that bending magnet radiation may be treated exhaustively
3763: in the framework of Geometrical Optics even for third generation
3764: light sources.
3765: 
3766: 
3767: \section{\label{sec:conc} Conclusions}
3768: 
3769: This work presents a theory of transverse coherence properties
3770: from third generation light sources, valid while radiation evolves
3771: in free-space. Besides being important for experiments involving
3772: coherence that make no use of optical elements, it constitutes the
3773: first step towards the solution of the image formation problem for
3774: undulator sources (see \cite{OURP}) that will be a natural
3775: follow-up to the present article.
3776: 
3777: We considered Synchrotron Radiation (SR) as a random statistical
3778: process to be described using the language of statistical optics.
3779: Statistical optics developed in connection with Gaussian,
3780: stationary processes characterized by homogeneous sources.
3781: However, for SR, there is no a priori reason to hold these
3782: assumptions satisfied.
3783: 
3784: We showed that SR is a Gaussian random process. As a result,
3785: statistical properties of SR are described satisfactory by
3786: second-order field correlation functions. We focused, in
3787: particular, on undulator sources. It should be noted here that
3788: wiggler and bending magnets are still being used at third
3789: generation facilities. However, as has been remarked in the
3790: previous Section \ref{sub:geop}, these devices are characterized
3791: by a much shorter formation length, which allows one to apply a
3792: Geometrical Optics approach to describe them. Thus, in this case,
3793: the formalism developed for second generation facilities can be
3794: satisfactory taken advantage of. In other words, use of wigglers
3795: and bending magnets is mainly related with applications requiring
3796: higher photon flux (but not high coherent flux), compared with
3797: analogous devices installed in second generation facilities. Our
3798: choice of considering undulator sources is justified by the fact
3799: that we are interested in transverse coherence properties of
3800: radiation.
3801: 
3802: With this in mind, a frequency-domain analysis was used to
3803: describe undulator sources from a mathematical viewpoint. As a
3804: consequence of the frequency domain analysis we could study the
3805: spatial correlation for a given frequency content using the
3806: cross-spectral density of the system. This can be used to extract
3807: information even if the process is non-stationary, and
3808: independently of the spectral correlation function.
3809: 
3810: We gave a general expression for the cross-spectral density
3811: dependent on six dimensionless parameters. Subsequently we assumed
3812: small normalized detuning from resonance, thus obtaining a
3813: simplified expression of practical relevance.
3814: 
3815: We simplified our expressions further in the case of third
3816: generation light sources, based on a large horizontal emittance
3817: (compared with the wavelength). In this case, the cross-spectral
3818: density can be factored in the product of a horizontal and a
3819: vertical factor.
3820: 
3821: Attention was subsequently drawn on the vertical cross-spectral
3822: density $\hat{G}_y$, without loss of generality. We expressed
3823: $\hat{G}_y$ in terms of one-dimensional convolutions between
3824: universal functions and Gaussian functions, and studied different
3825: asymptotic cases of interest.
3826: 
3827: In the case of a vertical emittance much smaller than the
3828: radiation wavelength we derived the counter-intuitive result that
3829: radiation is not fully coherent in the vertical direction. This
3830: effect can be interpreted as an influence of a large horizontal
3831: emittance on the vertical plane, and is related with the
3832: particular non-Gaussian nature of the single particle field.
3833: 
3834: Subsequently, we studied  quasi-homogeneous cases of interest,
3835: discussing the applicability of the VCZ theorem for Gaussian and
3836: non-Gaussian quasi-homogeneous sources. Finally, we discussed the
3837: accuracy of the quasi-homogeneous model.
3838: 
3839: It is interesting to spend a few words about relation of our
3840: theory with numerical techniques. Computer codes have been written
3841: (see e.g. \cite{CHUB,BAHR}) that deal with beamline design, based
3842: on wave-optics techniques. Codes also begin to be used to treat
3843: the case of partially coherent radiation: one of their final goals
3844: is to solve the image formation problem starting from first
3845: principles. Results may be obtained using numerical techniques
3846: alone, starting from the Lienard-Wiechert expressions for the
3847: electromagnetic field and applying the definition of the field
3848: correlation function without any analytical work. Calculation of
3849: the intensity at a single point on the image plane can be
3850: approached by propagating wavefronts from different
3851: macro-particles through the entire optical system, calculating
3852: intensities and summing them up. This method is well-suited for
3853: parallel processing, and relatively easy to implement. Yet, a
3854: first-principle calculation of the field correlation function
3855: between two generic points at the image plane involves very
3856: complicated and time-expensive numerical evaluations. To be
3857: specific, one needs to perform two integrations along the
3858: undulator device and four integrations over the electron-beam
3859: phase space distribution to solve the problem in free space. Then,
3860: even in the simple case when the optical beamline can be modelled
3861: as a single focusing lens, other four integrations are needed to
3862: characterize coherence properties on the image plane, for a total
3863: of ten integrations. The development of a universal code for any
3864: experimental setup is then likely to be problematic. A more
3865: conservative approach may suggest the use of computer codes based
3866: on some analytical transformation of first principle equations
3867: suited for specific experimental setups. From this viewpoint our
3868: most general expressions (or alternatively, as it is being done,
3869: expressions for the Wigner distribution function) may be used as
3870: reliable basis for the development of numerical methods. Our
3871: analytical theory allows treatment and physical understanding of
3872: many asymptotes of the parameter space and their applicability
3873: region with the help of a consistent use of dimensional analysis.
3874: This physical understanding, together with the possibility of
3875: using our asymptotic results as benchmarks for numerical methods,
3876: will be of help to code writers.
3877: 
3878: In closing, as has been remarked in \cite{HOWE}: "[...] it is very
3879: desirable to have a way to model the performance of undulator
3880: beamlines with significant partial coherent effects, and such
3881: modelling would, naturally, start with the source. The calculation
3882: would involve the knowledge of the partial coherence properties of
3883: the source itself and of how to propagate partially coherent
3884: fields through space and through the optical components used in
3885: the beamline. [...] it is important to recognize that, although
3886: most of these calculations are, in principle, straightforward
3887: applications of conventional coherence theory (Born and Wolf,
3888: 1980; Goodman, 1985), there is not much current interest in the
3889: visible optics community.". These statements were formulated more
3890: than ten years ago, when operation of third generation light
3891: sources started. While it was immediately recognized that usual SR
3892: theory was not adequate to describe them, no theoretical progress
3893: was made in that direction. Our paper finally answers the call in
3894: \cite{HOWE}.
3895: 
3896: 
3897: 
3898: \newpage
3899: 
3900: \section*{Acknowledgements}
3901: 
3902: The authors wish to thank Hermann Franz and Petr Ilinski for many
3903: useful discussions, Massimo Altarelli, Jochen Schneider and Edgar
3904: Weckert for their interest in this work.
3905: 
3906: \begin{thebibliography}{99}
3907: 
3908: \bibitem{EDGA} D. Bilderback, P. Elleaume and E. Weckert, J. Phys.
3909: B: At. Mol. Opt. Phys. 38, 773 (2005)
3910: \bibitem{PETR} K. Balewski, W. Brefeld, W. Decking, H. Franz, R. R\"{o}hlsberger and E. Weckert (ed) 2004 PETRA III:
3911: a low emittance synchrotron radiation source, Technical Design
3912: Report DESY 2004-035, Online available at
3913: http://www-hasylab.desy.de/facility/upgrade/petra$_-$tdr.htm
3914: \bibitem{GOOD} J. W. Goodman, Statistical optics, John Wiley $\&$
3915: Sons, Inc., 1985
3916: \bibitem{MAND} L. Mandel and E. Wolf, Optical Coherence and
3917: Quantum optics, Cambridge University Press, 1995
3918: \bibitem{NEIL} E. O'Neill, Introduction to Statistical optics,
3919: Dover Publications Inc., Mineola, New York (1991)
3920: \bibitem{HOWE} M.R. Howells and M.M. Kincaid, "The properties of
3921: undulator radiation" in New Directions in Research with
3922: Third-Generation Soft X-Ray Synchrotron Radiation Sources, Kluwer
3923: Academic Publishers, 315-358 (1994)
3924: \bibitem{TAKA} Y. Takayama et al., Nucl. Instr. Meth.
3925: in Phys. Res. A 441, 565 (2000)
3926: \bibitem{COI1} R. Co\"{i}sson, Applied Optics 34, No. 5, 904-908
3927: (1995)
3928: \bibitem{COI2} R. Co\"{i}sson and S. Marchesini, J. Synchrotron
3929: Radiat. 4, 263 (1997)
3930: \bibitem{COI3} R. Co\"{\i}sson and S. Marchesini, "Spatial
3931: Coherence of Synchrotron Radiation", Recent research developments
3932: in Optics, 3, 213 (2003), and report UCRL-JRNL-200688
3933: http://library.llnl.gov/uhtbin/cgisirsi/0/0/0/60/55/X
3934: \bibitem{KIM2} K-J. Kim, Nucl. Instr. Meth., A246, p. 71, 1986
3935: \bibitem{KIM3} K-J. Kim, SPIE Proceedings. vol. 582, p. 2, 1986.
3936: \bibitem{OURU} G. Geloni, E. Saldin, E. Schneidmiller and M.
3937: Yurkov, "Understanding transverse coherence properties of X-ray
3938: beams in third generation light sources", DESY 05-109 (2005)
3939: \bibitem{WOLF} E. Wolf, Nature, 172, 535 (1953)
3940: \bibitem{NOST} M. Yabashi, K. Tamasku et al., Phys. Rev. Lett.,
3941: 88, 24 (2000)
3942: \bibitem{MAN2} L. Mandel, J. Opt. Soc. Am. 51,1342 (1961)
3943: \bibitem{OURF} G. Geloni, E. Saldin, E. Schneidmiller and M.
3944: Yurkov, Opt. Commun 276, 1, 167 (2007)
3945: \bibitem{XFEL} M. Altarelli et al. (ed.), XFEL: The European X-Ray Free-Electron
3946: Laser. Technical Design Report, DESY 2006-097, DESY, Hamburg
3947: (2006) (See also http://xfel.desy.de)
3948: %\bibitem{OURS} G. Geloni, E. Saldin, E. Schneidmiller and M.
3949: %Yurkov, "Paraxial Green's functions in Synchrotron Radiation
3950: %theory", DESY 05-032, ISSN 0418-9833 (2005)
3951: \bibitem{YAB1} M. Yabashi, K. Tamasaku and T. Ishikawa, Phys. Rev.
3952: A 69, 023813 (2004)
3953: \bibitem{PFEI} F. Pfeiffer et al., Phys. Rev. Lett. 94, 164801
3954: (2005)
3955: \bibitem{YAB2} M. Yabashi, K. Tamasaku S. Goto and T. Ishikawa, J.
3956: Phys D: Appl. Phys. 38 (2005) A11-A16
3957: \bibitem{OURP} G. Geloni, E. Saldin, E. Schneidmiller and M.
3958: Yurkov "Statistical Optics Approach to the Design of Beamlines for
3959: Synchrotron Radiation", DESY 06-037, ISSN 0418-9833 (2006)
3960: \bibitem{CHUB} O.Chubar and P.Elleaume, in Proc. of the 6th European Particle Accelerator Conference EPAC-98, 1177-1179 (1998)
3961: \bibitem{BAHR} J. Bahrdt, Phys. Rev. ST-AB 10, 060701 (2007)
3962: \end{thebibliography}
3963: 
3964: 
3965: 
3966: 
3967: \end{document}
3968: