-
Notifications
You must be signed in to change notification settings - Fork 6
/
Copy pathleapx.tex
3048 lines (2712 loc) · 133 KB
/
leapx.tex
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
34
35
36
37
38
39
40
41
42
43
44
45
46
47
48
49
50
51
52
53
54
55
56
57
58
59
60
61
62
63
64
65
66
67
68
69
70
71
72
73
74
75
76
77
78
79
80
81
82
83
84
85
86
87
88
89
90
91
92
93
94
95
96
97
98
99
100
101
102
103
104
105
106
107
108
109
110
111
112
113
114
115
116
117
118
119
120
121
122
123
124
125
126
127
128
129
130
131
132
133
134
135
136
137
138
139
140
141
142
143
144
145
146
147
148
149
150
151
152
153
154
155
156
157
158
159
160
161
162
163
164
165
166
167
168
169
170
171
172
173
174
175
176
177
178
179
180
181
182
183
184
185
186
187
188
189
190
191
192
193
194
195
196
197
198
199
200
201
202
203
204
205
206
207
208
209
210
211
212
213
214
215
216
217
218
219
220
221
222
223
224
225
226
227
228
229
230
231
232
233
234
235
236
237
238
239
240
241
242
243
244
245
246
247
248
249
250
251
252
253
254
255
256
257
258
259
260
261
262
263
264
265
266
267
268
269
270
271
272
273
274
275
276
277
278
279
280
281
282
283
284
285
286
287
288
289
290
291
292
293
294
295
296
297
298
299
300
301
302
303
304
305
306
307
308
309
310
311
312
313
314
315
316
317
318
319
320
321
322
323
324
325
326
327
328
329
330
331
332
333
334
335
336
337
338
339
340
341
342
343
344
345
346
347
348
349
350
351
352
353
354
355
356
357
358
359
360
361
362
363
364
365
366
367
368
369
370
371
372
373
374
375
376
377
378
379
380
381
382
383
384
385
386
387
388
389
390
391
392
393
394
395
396
397
398
399
400
401
402
403
404
405
406
407
408
409
410
411
412
413
414
415
416
417
418
419
420
421
422
423
424
425
426
427
428
429
430
431
432
433
434
435
436
437
438
439
440
441
442
443
444
445
446
447
448
449
450
451
452
453
454
455
456
457
458
459
460
461
462
463
464
465
466
467
468
469
470
471
472
473
474
475
476
477
478
479
480
481
482
483
484
485
486
487
488
489
490
491
492
493
494
495
496
497
498
499
500
501
502
503
504
505
506
507
508
509
510
511
512
513
514
515
516
517
518
519
520
521
522
523
524
525
526
527
528
529
530
531
532
533
534
535
536
537
538
539
540
541
542
543
544
545
546
547
548
549
550
551
552
553
554
555
556
557
558
559
560
561
562
563
564
565
566
567
568
569
570
571
572
573
574
575
576
577
578
579
580
581
582
583
584
585
586
587
588
589
590
591
592
593
594
595
596
597
598
599
600
601
602
603
604
605
606
607
608
609
610
611
612
613
614
615
616
617
618
619
620
621
622
623
624
625
626
627
628
629
630
631
632
633
634
635
636
637
638
639
640
641
642
643
644
645
646
647
648
649
650
651
652
653
654
655
656
657
658
659
660
661
662
663
664
665
666
667
668
669
670
671
672
673
674
675
676
677
678
679
680
681
682
683
684
685
686
687
688
689
690
691
692
693
694
695
696
697
698
699
700
701
702
703
704
705
706
707
708
709
710
711
712
713
714
715
716
717
718
719
720
721
722
723
724
725
726
727
728
729
730
731
732
733
734
735
736
737
738
739
740
741
742
743
744
745
746
747
748
749
750
751
752
753
754
755
756
757
758
759
760
761
762
763
764
765
766
767
768
769
770
771
772
773
774
775
776
777
778
779
780
781
782
783
784
785
786
787
788
789
790
791
792
793
794
795
796
797
798
799
800
801
802
803
804
805
806
807
808
809
810
811
812
813
814
815
816
817
818
819
820
821
822
823
824
825
826
827
828
829
830
831
832
833
834
835
836
837
838
839
840
841
842
843
844
845
846
847
848
849
850
851
852
853
854
855
856
857
858
859
860
861
862
863
864
865
866
867
868
869
870
871
872
873
874
875
876
877
878
879
880
881
882
883
884
885
886
887
888
889
890
891
892
893
894
895
896
897
898
899
900
901
902
903
904
905
906
907
908
909
910
911
912
913
914
915
916
917
918
919
920
921
922
923
924
925
926
927
928
929
930
931
932
933
934
935
936
937
938
939
940
941
942
943
944
945
946
947
948
949
950
951
952
953
954
955
956
957
958
959
960
961
962
963
964
965
966
967
968
969
970
971
972
973
974
975
976
977
978
979
980
981
982
983
984
985
986
987
988
989
990
991
992
993
994
995
996
997
998
999
1000
\section{LEAPR}
\label{sLEAPR}
\index{LEAPR|textbf}
\hypertarget{sLEAPRhy}{The}
LEAPR\index{LEAPR} module is used to prepare the scattering law
$S(\alpha,\beta)$\index{$S(\alpha,\beta)$}, which describes thermal
scattering\index{thermal scattering} from bound moderators, in the
ENDF-6 format used by
\hyperlink{sTHERMRhy}{THERMR}\index{THERMR}. The original ENDF thermal
scattering data\cite{GAreport} was prepared by General Atomics (GA)
using the GASKET code\cite{GASKET}. This code has difficulty working
with the very large energy and momentum transfers encountered for
large incident energy or very low temperatures. For these reasons,
LEAPR is based on the British code LEAP+ADDELT originally written
by McLatchie\index{McLatchie} at Harwell\cite{McLatchie}, then
implemented by Butland\index{Butland} at Winfrith\cite{Butland},
and finally modified to work better for cold moderators as part of
the Ph.D. Thesis of D.~J.~Picton\cite{Picton}\index{Picton}. The
code that Dr. Picton provided
was modified extensively to fit better into the NJOY environment
and to take advantage of modern large computers. This involved
massive rearrangement of storage and routines, updating for first
FORTRAN-77, then Fortran-90, and extensive editing of the comment
cards. The original Edgewood expansion and Short Collision
Time (SCT) treatments were removed in favor of using more terms
in the phonon expansion and using the simple ENDF SCT
formula\cite{ENDF102}.\index{short collision time} In
addition, output in ENDF-6 format\cite{ENDF102} was provided, the
capability to include either coherent or incoherent elastic scattering
was added\index{coherent elastic}\index{incoherent elastic}, a major
speedup for the diffusion calculation was provided, and a capability
to produce a mixed scattering law for materials like BeO\index{BeO}
and benzine\index{benzine} was developed. In order to improve the numerics
on short-word computers, the code was changed to use the asymmetric
scattering function, the normalization scheme for the phonon
expansion was revised, and the discrete-oscillator calculation
was rebuilt to take better advantage of the convolution properties
of the delta function and to use better Bessel Functions. A complete
discussion of the theory used in the code is given below.
LEAPR was used\cite{New} to reevaluate several of the thermal moderator
materials from ENDF/B-VI\index{ENDF!ENDF/B-VI} using the original GA
physics\cite{GAreport} but extending the alpha and beta ranges for
incoherent inelastic\index{incoherent inelastic} scattering. The
energy range for coherent scattering was also increased. The code
was also used to prepare scattering-law files for several cold
moderators\index{cold moderators} of interest for experimental
cold-neutron sources\cite{ICANS,coldworkshop}. More recently, some
of these materials were updated for ENDF/B-VII.0\index{ENDF!ENDF/B-VII}
and additional materials were added. Several of the ENDF/B-VII.0 cases
are discussed below as examples of how to run LEAPR.
Acknowledgements are due to Gary Russell\index{Russell} of the
Los Alamos Neutron Science Center (LANSCE) for motivating this
work, to Drew Kornreich\index{Kornreich} of the University of Arizona
for his careful study of the water problem, and to Max Lazo\index{Lazo}
(University of New Mexico, SAIC, and Sandia National Laboratory) for
testing and reviewing the code and documentation. In addition, we
thank D. J. Picton\index{Picton} for sending us his cold-moderator
version of LEAPR, Rolf Neef\index{Neef} of Julich and Guy
Robert\index{Robert} of ILL-Grenoble for sending us the initial
version of the cold hydrogen treatment, and M. Mattes\index{Mattes}
and J. Keinert\index{Keinert} of the University of Stuttgart for their
help with the liquid hydrogen and deuterium models.
This chapter describes the LEAPR module in NJOY2016.0.
\subsection{Theory}
\label{ssLEAPR_theory}
The following discussion of the theories used in the the code is based
on the original British documentation and the presentation in a standard
reference\cite{Parks}.
\paragraph{Coherent and Incoherent Scattering.}
In practice, the scattering of neutrons from a system of N particles
with a random distribution of spins or isotope types can be expressed
as the sum of a coherent part and an incoherent part. The coherent
scattering includes the effects from waves that are able to interfere
with each other, and the incoherent part depends on a simple sum of
noninterfering waves from all the N particles. (The spin correlations
in ortho and para hydrogen violate the assumption of randomness, so
liquid hydrogen\index{liquid hydrogen} does not fit into the model
described here. A method for treating them will be described below.)
The cross sections for coherent and incoherent scattering can be
considered to be characteristic properties of the materials.
\index{coherent elastic}\index{incoherent elastic} As examples,
the scattering from hydrogen is almost completely incoherent, and the
scattering from carbon and oxygen is almost completely coherent.
Furthermore, the coherent and incoherent scattering include both
elastic and inelastic parts. The elastic scattering takes place with
no energy change. It should not be confused with the elastic
scattering from a single particle that is familiar for higher neutron
energies where the neutron loses energy; thermal elastic scattering can
be considered to be scattering from the entire lattice, thus the
effective mass of the target is very large, and the neutron does not
lose energy in the scattering process. Thermal inelastic scattering
results in an energy loss (gain) for the neutron with a corresponding
excitation (deexcitation) of the target. The excitation may correspond
to the production of one or more phonons in a crystalline material, to
the production of rotations or vibrations in molecules, or to the
initiation of atomic or molecular recoil motions in a liquid or gas.
In addition, the coherent inelastic part of the scattering contains both
interference effects between waves scattered by different particles and
direct terms. It turns out that the direct part for gases, liquids and
solids consisting of randomly oriented crystallites has approximately the
same form as the incoherent term. The interference is usually neglected.
Therefore, we can usually divide the thermal scattering cross section into
three different parts:
\begin{itemize}
\begin{singlespace}
\item {\it Coherent elastic.} Important for crystalline solids
like graphite or beryllium.\index{coherent elastic}
\item {\it Incoherent elastic.} Important for hydrogenous solids
like solid methane, polyethylene, and zirconium
hydride.\index{incoherent elastic}
\item {\it Inelastic.} Important for all materials
(this category includes both incoherent and coherent
inelastic).\index{incoherent inelastic}
\end{singlespace}
\end{itemize}
The absence of interference in incoherent scattering and the neglect of
interference in coherent inelastic scattering allows us to construct
thermal scattering laws for ``hydrogen in water" or ``hydrogen in
solid methane'' or ``oxygen in beryllium oxide". However, this
simplification is not possible in general for coherent elastic scattering
in materials with more that one type of atom in the unit cell;
for coherent elastic scattering, beryllium oxide must be considered as
a unit.
\paragraph{Inelastic Scattering}
It is shown in the standard references\cite{Parks} that the double
differential scattering cross section for thermal neutrons for gases,
liquids or solids consisting of randomly ordered microcrystals can be
written as
\begin{equation}
\sigma(E{\rightarrow}E',\mu)=\frac{\sigma_b}{2kT}
\sqrt{\frac{E'}{E}}\,{\cal S}(\alpha,\beta)\,,
\end{equation}
\noindent
where $E$ and $E'$ are the incident and secondary neutron energies in the
laboratory system, $\mu$ is the cosine of the scattering angle in the
laboratory, $\sigma_b$ is the characteristic bound scattering cross
section for the material, $kT$ is the thermal energy in eV, and ${\cal S}$
is the asymmetric form of the scattering law. The scattering law depends
on only two variables: the momentum transfer
\begin{equation}
\alpha=\frac{E'+E-2\mu\sqrt{E'E}}{AkT}\,,
\label{alph}
\end{equation}
\noindent
where $A$ is the ratio of the scatterer mass to the neutron mass, and
the energy transfer
\begin{equation}
\beta=\frac{E'-E}{kT}\,.
\end{equation}
\noindent
Note that $\beta$ is positive for energy gain and negative for energy
loss. Working in the incoherent approximation and the Gaussian approximation,
the scattering law\index{scattering law} can be written
\begin{equation}
{\cal S}(\alpha,\beta)=\frac{1}{2\pi}\intall\exx{i\beta\hat{t}}
\,\exx{-\gamma(\hat{t})}\,d\hat{t}\,,
\label{sincoh}
\end{equation}
\noindent
where $\hat{t}$ is time measured in units of $\hbar/(kT)$ seconds.
The function $\gamma(\hat{t})$ is given by
\begin{equation}
\gamma(\hat{t})=\alpha\intall P(\beta)
\left[\,1-\exx{-i\beta\hat{t}}\,\right]\,\exx{-\beta/2}\,d\beta\,,
\label{gamma}
\end{equation}
\noindent
where
\noindent
\begin{equation}
P(\beta)=\frac{\rho(\beta)}{2\beta\sinh(\beta/2)}\,,
\label{pbeta}
\end{equation}
\noindent
and where $\rho(\beta)$ is the frequency spectrum\index{frequency spectrum}
of excitations in the system expressed as a function of $\beta$. The
spectrum must be normalized as
follows:
\begin{equation}
\int_0^\infty \rho(\beta)\,d\beta=1\,.
\label{rnorm}
\end{equation}
\noindent
The function $P(\beta)$ defined by Eq.(\ref{pbeta}) is used directly in
LEAPR, and $\rho(\beta)=\rho(\epsilon/kt)$ is given as a function of
$\epsilon$ in eV. For systems in thermal equilibrium, there is a
relationship between upscatter and downscatter called ``detail balance''
\index{detail balance} that is a consequence of microscopic reversibility.
It requires that
\begin{equation}
{\cal S}(\alpha,\beta)=\exx{-\beta}{\cal S}(\alpha,-\beta)\,.
\end{equation}
\noindent
Liquid hydrogen\index{liquid hydrogen} and deuterium violate this
condition, as will be described below. In addition, the scattering
law satisfies two other important constraints; namely, normalization,
\begin{equation}
\intall {\cal S}(\alpha,\beta)\,d\beta=1\,,
\label{ssnorm}
\end{equation}
\noindent
and the sum rule
\begin{equation}
\intall \beta\,{\cal S}(\alpha,\beta)\,d\beta=-\alpha\,.
\label{sssum}
\end{equation}
\noindent
Actually, ENDF works with the so-called ``symmetric'' $S(\alpha,\beta)$,
\begin{equation}
S(\alpha,\beta)=\exx{\beta/2}\,{\cal S}(\alpha,\beta)\,
\end{equation}
\noindent
which (except for liquid hydrogen or deuterium) satisfies the condition
\begin{equation}
S(\alpha,\beta)=S(\alpha,-\beta)\,.
\end{equation}
\noindent
Note that ${\cal S}(\alpha,-\beta)$ for positive $\beta$ describes
the downscatter side of the function, and because it is basically
proportional to the cross section, it can be represented by reasonable
numbers (say $10^{-8}$ to 1) for all $\beta$. The symmetric
$S(\alpha,\beta)$, on the other hand, can easily be smaller than ${\cal
S}(\alpha,-\beta)$ by factors like ${\rm e}^{-\beta/2} \sim {\rm e}^{-80}
\sim 10^{-35}$, which can cause trouble on short-word machines.
This kind of numerical problem is even more severe for cold moderators,
where dynamic ranges on the order of $10^{100}$ occur for
$S(\alpha,\beta)$. (The user will have to use some caution reading this
report, because the typographic symbols for S and script-S are very
similar.) By working with ${\cal S}(\alpha,-\beta)$, LEAPR can do all
of its calculations using single-precision variables, even on
a short-word machine.
The next step is to decompose the frequency spectrum into a sum of
simple spectra
\begin{equation}
\rho(\beta)=\sum_{j=1}^K\rho_j(\beta),
\end{equation}
\vspace{0.5 pt}
\noindent
where the following possibilities are allowed:
\begin{eqnarray}
\rho_j(\beta) &=& w_j\delta(\beta_j)\;\; \hbox{discrete oscillator}\\
\rho_j(\beta) &=& \rho_s(\beta) \;\;\;\; \hbox{solid-type spectrum}\\
\rho_j(\beta) &=& \rho_t(\beta) \;\;\;\; \hbox{translational spectrum}
\end{eqnarray}
\noindent
The solid-type spectrum must vary as $\beta^2$ as $\beta$ goes to zero,
and it must integrate to $w_s$, the weight for the solid-type law. The
translational spectrum can be either a free-gas law or a diffusion-type
spectrum represented with the approximation of Egelstaff and Schofield
that will be discussed later. In either case, the spectrum must
integrate to $w_t$, the translational weight. The sum of all the
weights of the partial spectra must equal 1. Defining
$\gamma_j(\hat{t})$ to be the value of $\gamma$ appropriate for
$\rho_j$, and ${\cal S}_j$ to be the corresponding partial scattering law,
and using the convolution theorem for Fourier transforms, leads to a
recursive formula for the scattering law:
\begin{equation}
{\cal S}(\alpha,\beta)={\cal S}^{(K)}(\alpha,\beta),
\end{equation}
\noindent
where
\begin{eqnarray}
{\cal S}^{(J)}(\alpha,\beta) &=&
\frac{1}{2\pi}\intall\exx{i\beta\hat{t}}
\prod_{j=1}^J\exx{-\gamma_j(\hat{t})}\,d\hat{t} \nonumber\\
&=&\intall{\cal S}_J(\alpha,\beta')\,
{\cal S}^{(J-1)}(\alpha,\beta{-}\beta')\,d\beta'\,.
\end{eqnarray}
\noindent
As an example of the use of this recursive procedure, consider a case
where the desired frequency spectrum is a combination of $\rho_s$ and
two discrete oscillators. First, calculate ${\cal S}^{(1)}{=}{\cal
S}_1$ using $\rho_s$. Then calculate ${\cal S}_2$ using
$\rho(\beta_1)$, the distribution for the first discrete oscillator,
and convolve ${\cal S}_2$ with ${\cal S}^{(1)}$ to obtain ${\cal
S}^{(2)}$, the composite scattering law for the first two partial
distributions. Repeat the process with the second discrete oscillator
to obtain ${\cal S}^{(3)}$, which is equal to ${\cal S}(\alpha,\beta)$
for the full distribution.
\paragraph{The Phonon Expansion.}\index{phonon expansion} Consider
first $\gamma_s(\hat{t})$, the Gaussian function for solid-type
frequency spectra. Expanding the time-dependent part of the exponential
gives
\begin{equation}
\exx{-\gamma_s(\hat{t})}=\exx{-\alpha\lambda_s}
\sum_{n=0}^{\infty}\frac{1}{n!}\left[\,\alpha
\intall P_s(\beta)\,\exx{-\beta/2}\exx{-i\beta\hat{t}}
\,d\beta\,\right]^n\,,
\end{equation}
\noindent
where $\lambda_s$ is the Debye-Waller coefficient
\begin{equation}
\lambda_s=\intall P_s(\beta)\,\exx{-\beta/2}\,d\beta\,.
\label{dbwf}
\end{equation}
\vspace{0.5 pt}
\noindent
The scattering function becomes
\begin{eqnarray}
{\cal S}_s(\alpha,\beta)&=&\exx{-\alpha\lambda_s}\sum_{n=0}^{\infty}
\frac{1}{n!}\alpha^n\nonumber\\
& &\times\,\frac{1}{2\pi} \intall\exx{i\beta\hat{t}}\left[
\intall P_s(\beta')\,\exx{-\beta'/2}\,\exx{-i\beta'\hat{t}}
\,d\beta'\right]^n d\hat{t}\,.
\end{eqnarray}
\noindent
For convenience, define the quantity in the second line of this equation
to be $\lambda_s^n {\cal T}_n(\beta)$. Then clearly,
\begin{equation}
{\cal S}_s(\alpha,\beta)=\exx{-\alpha\lambda_s}
\sum_{n=0}^{\infty}\frac{1}{n!}\,[\alpha\lambda_s]^n\,{\cal T}_n(\beta)\,,
\label{phoexp}
\end{equation}
\noindent
where
\begin{equation}
{\cal T}_0(\beta)=\frac{1}{2\pi}
\intall\exx{i\beta\hat{t}}\,d\hat{t}=\delta(\beta)\,,
\end{equation}
\noindent
and
\begin{equation}
{\cal T}_1(\beta)=\intall \frac{P_s(\beta')\,
\exx{-\beta'/2}}{\lambda_s}\left\{\frac{1}{2\pi}
\intall\exx{i(\beta{-}\beta')\hat{t}}\,d\hat{t}\right\}d\beta'
=\frac{P_s(\beta)\,\exx{-\beta/2}}{\lambda_s}\,,
\end{equation}
\noindent
In general,
\begin{equation}
{\cal T}_n(\beta)=\intall {\cal T}_1(\beta')\,
{\cal T}_{n-1}(\beta{-}\beta')\,d\beta'\,.
\label{phoconv}
\end{equation}
\noindent
The script-T functions obey the relationship
${\cal T}_n(\beta)=\exx{-\beta}{\cal T}_n(-\beta)$.
In addition, each of the ${\cal T}_n$ functions
obeys the following normalization condition:
\begin{equation}
\intall {\cal T}_n(\beta)\,d\beta = 1\,.
\end{equation}
\noindent
It guarantees that Eq.\ref{ssnorm} will be satisfied by the sum in
Eq.\ref{phoexp}. In LEAPR, the ${\cal T}_n(-\beta)$ functions are
precomputed on the input $\beta$ grid for $n$ up to some specified
maximum value, typically 100. It is then easy to compute the smooth
part of ${\cal S}_s(\alpha,-\beta)$ for any sufficiently small value of
$\alpha$ using Eq.\ref{phoexp}. The corresponding values of
${\cal S}_s(\alpha,\beta)$ can then be obtained by multiplying by
$\exx{-\beta}$. The delta function arising from the ``zero-phonon''
term is carried forward separately. The normalization in
Eq.\ref{phoexp} has better numerical properties than the one used in
LEAP.
\paragraph{The Short-Collision-Time Approximation.}\index{short collision time}
For larger values of $\alpha$, the expansion of Eq.\ref{phoexp} requires
too many terms. LEAPR uses the simple Short-Collision-Time (SCT)
approximation from ENDF to extend the solid-type scattering law.
\begin{equation}
{\cal S}_s(\alpha,-\beta)=\frac{1}{\sqrt{4\pi w_s\alpha
\overline{T}_s/T}}\,
\exp\left[-\frac{(w_s\alpha-\beta)^2}{w_s\alpha
\overline{T}_s/T}\right]\,,
\end{equation}
\noindent
and
\begin{equation}
{\cal S}_s(\alpha,\beta)=\exx{-\beta}{\cal S}_s(\alpha,-\beta)\,
\end{equation}
\noindent
where $\beta$ is positive, and
where the effective temperature\index{effective temperature} is given by
\begin{equation}
\overline{T}_s = \frac{T}{2 w_s}
\intall\beta^2 P_s(\beta)\,\exx{-\beta}\,d\beta\,.
\end{equation}
\noindent
As above, $w_s$ is the weight for the solid-type spectrum.
\paragraph{Diffusion and Free-Gas Translation.}
The neutron scattering from many important liquids, including water
and liquid methane, can be represented using a solid-type spectrum
of rotational and vibrational modes combined with a
diffusion\index{diffusion} term. Egelstaff and Schofield have
proposed an especially simple model for diffusion called the
``effective width model''. It has the advantage of having
analytic forms for both $S(\alpha,\beta)$ and the
associated frequency spectrum $\rho(\beta)$:
\begin{eqnarray}
{\cal S}_t(\alpha,\beta)&=&\frac{2 c w_t\alpha}{\pi}\,
\,\exp\left[2 c^2 w_t\alpha-\beta/2\right]\, \nonumber\\
& &\frac{\sqrt{c^2+.25}}{\sqrt{\beta^2+4 c^2 w_t^2\alpha^2}}\,
K_1\Bigl\{\sqrt{c^2+.25}\sqrt{\beta^2+4 c^2 w_t^2 \alpha^2}
\Bigr\}\,,
\end{eqnarray}
\vspace{0.5 pt}
\noindent
and
\begin{equation}
\rho(\beta)=w_t \frac{4 c}{\pi\beta}\,\sqrt{c^2+.25}\,
\sinh(\beta/2)\,K_1\Bigl\{\sqrt{c^2+.25}\,\beta\Bigr\}\,.
\end{equation}
\vspace{0.5 pt}
\noindent
In these equations, $K_1(x)$ is a modified Bessel function of the
second kind, and the translational weight $w_t$ and the
diffusion constant $c$ are provided as inputs.
An alternative for the translational part of the distribution is
the free-gas law.\index{free gas} It is clearly appropriate for
a gas of molecules, but it has also been used to represent the
translational component for liquid moderators like water\cite{GAreport}.
The scattering law is given by
\begin{equation}
{\cal S}_t(\alpha,-\beta)=\frac{1}{\sqrt{4\pi w_t\alpha}}
\,\exp\left[-\frac{(w_t\alpha-\beta)^2}{4w_t\alpha}\right]\,,
\end{equation}
\noindent
and
\begin{equation}
{\cal S}_t(\alpha,\beta)=\exx{-\beta}{\cal S}_t(\alpha,-\beta)\,,
\end{equation}
\noindent
with $\beta$ positive. The free-gas law is used in LEAPR if the
diffusion coefficient $c$ is input as zero.
In LEAPR, ${\cal S}_s(\alpha,\beta)$, the scattering law for the solid-type
modes, is calculated using the phonon expansion as described above.
The translational contribution ${\cal S}_t(\alpha,\beta)$ is then calculated
using one of the formulas above on a $\beta$ grid chosen to represent its
shape fairly well. The combined scattering law is then obtained
by convolution as follows:
\begin{equation}
{\cal S}(\alpha,\beta)={\cal S}_t(\alpha,\beta)\,\exx{-\alpha\lambda_s}
+\intall {\cal S}_t(\alpha,\beta')\,
{\cal S}_s(\alpha,\beta{-}\beta')\,d\beta'.
\end{equation}
\noindent
The first term arises from the delta function in Eq.\ref{phoexp},
which isn't included in the numerical results for the phonon series
calculation. The values for ${\cal S}_t(\beta)$ and
${\cal S}_s(\beta{-}\beta')$ are obtained from
the precomputed functions by interpolation. This makes LEAPR run
much faster than LEAP+ADDELT for diffusive cases, because the original
code did direct recalculations of the solid-type scattering law for
all the desired values of $\beta{-}\beta'$. It also had to take pains
to compute $S_t$ on a $\beta$ grid that was commensurate with the
input grid. This often resulted in more points for $S_t$ than were
necessary to obtain useful accuracy for the convolutions.
The effective temperature\index{effective temperature} for a
combination of solid-type and translation modes is computed using
\begin{equation}
\overline{T}_s = \frac{w_t T + w_s \overline{T}_s}{w_t+w_s}\,.
\end{equation}
\paragraph{Discrete Oscillators.}\index{discrete oscillators}
Polyatomic molecules normally contain a number of vibrational
modes that can be represented as discrete oscillators. The distribution
function for one oscillator is given by $w_i\delta(\beta_i)$, where
$w_i$ is the fractional weight for mode $i$, and $\beta_i$ is the
energy-transfer parameter computed from the mode's vibrational
frequency. The corresponding scattering law is given by
\begin{eqnarray}
{\cal S}_i(\alpha,\beta) &=& \exx{-\alpha\lambda_i}\sum_{n=-\infty}^\infty\,
\delta(\beta-n\beta_i) \, I_n \left[
\frac{\alpha w_i}{\beta_i\sinh(\beta_i/2)} \right]\,
\exx{-n\beta_i/2}\nonumber\\
& = & \sum_{n=-\infty}^\infty\,A_{in}(\alpha)\,
\delta(\beta-n\beta_i)\,,
\end{eqnarray}
\noindent
where
\begin{equation}
\lambda_i = w_i \frac{\coth(\beta_i/2)}{\beta_i}\,.
\end{equation}
\noindent
The combination of a solid-type mode (s) with discrete
oscillators (1) and (2) would give
\begin{eqnarray}
{\cal S}^{(0)}(\alpha,\beta)&=&{\cal S}_s(\alpha,\beta)\,,\\
{\cal S}^{(1)}(\alpha,\beta) &=& \intall {\cal S}_1(\alpha,\beta')\,
{\cal S}^{(0)}(\alpha,\beta{-}\beta')\,d\beta' \nonumber\\
&=& \sum_{n=-\infty}^\infty\,A_{1n}(\alpha)\,
{\cal S}^{(0)}(\alpha,\beta{-}n\beta_1)\,,\\
{\cal S}^{(2)}(\alpha,\beta)&=&\intall {\cal S}_2(\alpha,\beta')\,
{\cal S}^{(1)}(\alpha,\beta{-}\beta')\,d\beta' \nonumber\\
&=& \sum_{m=-\infty}^\infty\,A_{2m}(\alpha)\,
\sum_{n=-\infty}^\infty\,A_{1n}(\alpha)\,\,
{\cal S}^{(0)}(\alpha,\beta{-}n\beta_1{-}m\beta_2)\,.
\end{eqnarray}
\noindent
This process can be continued through ${\cal S}^{(3)}(\alpha,\beta)$,
${\cal S}^{(4)}(\alpha,\beta)$, etc., until all the discrete oscillators
have been included. The result has the form
\begin{equation}
{\cal S}(\alpha,\beta)=\sum_k W_k(\alpha)\,
{\cal S}_s(\alpha,\beta-\beta_k)\,,
\end{equation}
\noindent
where the $\beta_k$ and the associated weights $W_k$ are easily
generated recursively using a procedure that throws out small weights
at each step. The Debye-Waller $\lambda$ for the combined
modes is computed using
\begin{equation}
\lambda=\lambda_s+\sum_{i=1}^N \lambda_i\,.
\end{equation}
\noindent
The effective temperature\index{effective temperature} for the
combined modes is given by
\begin{equation}
\overline{T}_s=w_t T + w_s \overline{T}_s
+ \sum_{i=1}^N w_i \frac{\beta_i}{2}
\coth\left(\frac{\beta_i}{2}\right) T\,.
\end{equation}
If the starting-point scattering law ${\cal S}^{(0)}$ does not contain
a translational contribution (true for hydrogenous solids like
polyethylene and frozen methane), it is important to remember to
include the effects of the ``zero-phonon'' term
$\exp(-\alpha\lambda_s)\delta(\beta)$. The code does this by
adding in triangular peaks with the proper areas and with their apexes
at the $\beta$ value closest to the $\beta_k$ values. One of these
peaks is at $\beta{=}0$. This peak is not put into the scattering
law as a sharp triangle; instead, it is handled as
``incoherent elastic'' scattering in order to take full advantage of
the analytic properties of $\delta(\beta)$.
\paragraph{Incoherent Elastic Scattering.}\index{incoherent elastic}
In hydrogenous solids, there is an elastic (no energy loss) component of
scattering arising from the ``zero-phonon'', or $n{=}0$ term, of
Eq.\ref{phoexp}. In ENDF terminology, this is called the ``incoherent
elastic'' term.
Clearly,
\begin{equation}
S_{\rm iel}(\alpha,\beta)=\rm e^{-\alpha\lambda}\delta(\beta)\,.
\end{equation}
\noindent
The corresponding differential scattering cross section is
\begin{equation}
\sigma(E,\mu)=\frac{\sigma_b}{2}\,\rm e^{-2WE(1-\mu)}\,,
\end{equation}
\noindent
and the integrated cross section is
\begin{equation}
\sigma(E)=\frac{\sigma_b}{2}\left\{\frac{1-\rm e^{-4WE}}{2WE}\right\}\,.
\end{equation}
\noindent
In these equations, the Debye-Waller coefficient\index{Debye-Waller
coefficient} is given by
\begin{equation}
W=\frac{\lambda}{AkT}\,,
\end{equation}
\noindent
where $\lambda$ is computed from the input frequency spectrum as shown
by Eq.\ref{dbwf} and modified by the presence of discrete
oscillators (if any) as shown above. LEAPR writes the bound scattering
cross section $\sigma_b$ and the Debye-Waller coefficient $W$ as a
function of temperature into a section of the ENDF-6 output
with MF=7 and MT=2.
\paragraph{Coherent Elastic Scattering.}\index{coherent elastic}
In solids consisting of coherent scatterers -- for example, graphite -- the
zero-phonon term leads to interference scattering from the various planes
of atoms of the crystals making up the solid. Once again, there is no
energy loss, and the ENDF term for the process is ``coherent elastic
scattering''. The differential scattering cross section is given by
\begin{equation}
\sigma_{\rm coh}(E,\mu)=\frac{\sigma_c}{E}
\sum_{E_i<E}f_i\,\rm e^{-4WE_i}\delta(\mu-\mu_i)\,,
\end{equation}
\noindent
where
\begin{equation}
\mu_i=1-E_i/E\,,
\end{equation}
\noindent
and the integrated cross section is given by
\begin{equation}
\sigma_{\rm coh}=\frac{\sigma_c}{E}
\sum_{E_i<E}f_i\,\rm e^{-4WE_i}.
\label{coh1}
\end{equation}
\noindent
In these equations, $\sigma_c$ is the effective bound coherent scattering
cross section for the material, $W$ is the effective Debye-Waller coefficient,
$E_i$ are the so-called ``Bragg Edges'',\index{Bragg edges} and the
$f_i$ are related to the crystallographic structure factors.
It can be seen from Eq.\ref{coh1} that the coherent elastic cross section
is zero below the first Bragg Edge, $E_1$ (typically about 2 to 5 meV). It
then jumps sharply to a value determined by $f_1$ and the Debye-Waller term.
At higher energies, the cross section drops off as $1/E$ until $E{=}E_2$.
It then takes another jump, and resumes its $1/E$ drop off. The sizes
of the steps in the cross section gradually get smaller, and at high energies
there is nothing left but an asymptotic $1/E$ decrease (typically above
1 to 2 eV). LEAPR stores the quantity $E\sigma_{\rm coh}(E)$ as a function
of energy and temperature in a section of the ENDF-6 output with MF=7 and
MT=2. The cross section is easily recovered from this representation by
dividing by $E$ (this is done in the
\hyperlink{sTHERMRhy}{THERMR}\index{THERMR} module of NJOY).
The angular distribution of scattered neutrons can be calculated by
extracting the $f_i$ from the steps in $\sigma_{\rm coh}(E)$ file by
subtraction. This process is carried out automatically in the
\hyperlink{sGROUPRhy}{GROUPR}\index{GROUPR} module of NJOY.
The calculation of the $E_i$ and $f_i$ depends on a knowledge of the
crystal structure of the scattering material. The methods used are
borrowed from HEXSCAT\cite{HEXSCAT}.
In general, the energies of the Bragg edges are given by
\begin{equation}
E_i=\frac{\hbar^2\tau_i^2}{8m}\,,
\end{equation}
\noindent
where $\tau_i$ is the length of the vectors of one particular ``shell''
of the reciprocal lattice, and $m$ is the neutron mass. The $f_i$ factors
for a material containing a single atomic species are given by
\begin{equation}
f_i=\frac{2\pi\hbar^2}{4mNV}\sum_{\tau_i} |F(\tau)|^2\,,
\end{equation}
\noindent
where the sum extends over all reciprocal lattice vectors of the given
length, and the crystallographic structure factor is given by
\begin{equation}
|F(\tau)|^2=\Bigl|\sum_{j=1}^{N}\rm e^{i2\pi\phi_j}\Bigr|^2\,,
\end{equation}
\noindent
where $N$ is the number of atoms in the unit cell,
$\phi_j=\vec{\tau}\cdot\vec{\rho}_j$
are the phases for the atoms, and the $\vec{\rho}_j$ are their positions.
The situation is more complicated for materials containing different
atomic species, such as beryllium oxide. In these cases,
\begin{equation}
\sigma_c\rm e^{-2WE_i}f_i=\Bigl|\sum_{j=1}^{N}
\sqrt{\sigma_j}\,\rm e^{-W_jE_i}\rm e^{i2\pi\phi_j}\Bigr|^2\,,
\end{equation}
\noindent
where the coherent cross section and Debye-Waller factor can be different
for each site in the unit cell. The effective coherent cross section is
clearly given by
\begin{equation}
\sigma_c=\sum_{j=1}^{N}\sigma_j\,.
\end{equation}
\noindent
Since LEAPR only works with one material at a time, it doesn't have access
to different values of $W_j$ for the atoms in the unit cell. Therefore, it
assumes that either $W_jE_i$ is small, or the $W_j$ doesn't vary much from
site to site. This allows it to calculate the $f_i$ using
\begin{equation}
|F|^2=\Bigl|\sum_{j=1}^N\frac{\sqrt{\sigma_j}}{\sqrt{\sigma_c}}
\,\rm e^{-2\pi\phi_j}\Bigr|^2\,.
\end{equation}
\vspace{0.5 pt}
\noindent
For hexagonal materials, the lattice is described by the two constants
$a$ and $c$. The reciprocal lattice vector lengths are given by
\begin{equation}
\Bigl(\frac{\tau}{2\pi}\Bigr)=\frac{4}{3a^2}
\Bigl(\ell_1^2+\ell_2^2+\ell_1\ell_2\Bigr)
+\frac{1}{c^2}\ell_3^2\,
\end{equation}
\vspace{0.5 pt}
\noindent
where $\ell_1,\ell_2,$ and $\ell_3$ run over all the positive and negative
integers, including zero. The volume of the unit cell is
\begin{equation}
V=\sqrt{3}a_2c/2\,.
\end{equation}
\noindent
For graphite, there are four atoms in the unit cell at positions\cite{Wycoff}
$$
(0,0,0),\,(-\frac{1}{3},\frac{1}{3},0),\,
(-\frac{2}{3},-\frac{1}{3},\frac{1}{2}),\,
(-\frac{1}{3},\frac{1}{3},\frac{1}{2})\,.
$$
\noindent
These positions give the following phases:
\begin{eqnarray}
\phi_1 &=& 0\,,\\
\phi_2 &=& (-\ell_1+\ell_2)/3\,,\\
\phi_3 &=& -(2/3)\ell_1-(1/3)\ell_2+(1/2)\ell_3\,,\\
\phi_4 &=& -(1/3)\ell_1+(1/3)\ell_2+(1/2)\ell_3\,.
\end{eqnarray}
\noindent
The form factor for graphite becomes
\begin{equation}
|F|^2=\left\{
\begin{array}{ll}
6+10\cos[2\pi(\ell_1-\ell_2)/3] & \mbox{$\ell_3$ even}\\
4\sin^2[\pi(\ell_1-\ell_2)/3] & \mbox{$\ell_3$ odd}
\end{array}
\right.
\end{equation}
\vspace{0.5 pt}
\noindent
For the hexagonal close packed (hcp) structure, which includes beryllium,
there are two atoms per unit cell at
$$
(0,0,0),\,(\frac{1}{3},\frac{2}{3},\frac{1}{2})\,,
$$
\noindent
and the form factor for hcp lattices like beryllium becomes
\begin{equation}
|F|^2=2+2\cos\Bigl[2\pi(2\ell_1+4\ell_2+3\ell_3)/6\Bigr]\,.
\end{equation}
\vspace{0.5 pt}
The beryllium oxide lattice consists of two interpenetrating hcp lattices,
one for the beryllium atoms, and one for the oxygen. There are four atoms
per unit cell with positions
$$
(0,0,0),\,(\frac{1}{3},\frac{2}{3},\frac{1}{2}),\,
(0,0,u), (\frac{1}{3},\frac{2}{3},u+\frac{1}{2})\,,
$$
\vspace{0.5 pt}
\noindent
Using the approximation that the Debye-Waller factor doesn't vary from
position to position in the unit cell gives the following expression for
the structure factor:
\begin{equation}
|F|^2=\Bigl(1+\cos[2\pi(2\ell_1+4\ell_2+3\ell_3)/6]\Bigr)
\Bigl(r_1^2+2r_1r_2\cos(3\pi\ell_3/4)\Bigr)\,,
\end{equation}
\noindent
where $r_1^2$ and $r_2^2$ are the bound coherent cross sections for
beryllium and oxygen, respectively, and the effective coherent cross
section, $\sigma_c$, is to be taken as 1.
LEAPR contains similar logic to handle face-centered cubic (fcc),
{\it e.g.} aluminum,
and body-centered cubic (bcc), {\it e.g.} iron, lattices. More formulas
for the structure factor can be added to the code when needed.
\paragraph{Liquid Hydrogen and Deuterium.}\index{liquid hydrogen}
Materials containing hydrogen (H) or deuterium (D) molecules violate the
assumption that spins are distributed randomly that underlies the
incoherent approximation used for Eq.\ref{phoexp}, and an explicitly
quantum-mechanical formula is required to take account of the correlations
between the spins of two atoms in the same molecule. This problem was
considered by Young and Koppel\cite{YK}. Changing to our notation, the
formulas for the hydrogen molecule (neglecting vibrations) become
\begin{eqnarray}
{\cal S}_{\rm para}(\alpha,\beta)&=&\sum_{J=0,2,4,\dots}P_J \nonumber \\
&\times&\frac{4\pi}{\sigma_b}\Bigl[A_{\rm para}\sum_{J'=0,2,4,\dots}
+B_{\rm para}\sum_{J'=1,3,5,\dots}\Bigr](2 J'+1)\nonumber\\
&\times&{\cal S}_f(w\alpha,\beta{+}\beta_{JJ'})\nonumber\\
&\times&\sum_{\ell=|J'-J|}^{J'+J}4\,j_\ell^2(y)\,C^2(JJ'\ell;00)\,,
\label{spara}
\end{eqnarray}
\noindent
and
\begin{eqnarray}
{\cal S}_{\rm ortho}(\alpha,\beta)&=&\sum_{J=1,3,5,\dots}P_J\nonumber\\
&\times&\frac{4\pi}{\sigma_b}\Bigl[A_{\rm ortho}\sum_{J'=0,2,4,\dots}
+B_{\rm ortho}\sum_{J'=1,3,5,\dots}\Bigr](2 J'+1)\nonumber\\
&\times&{\cal S}_f(w\alpha,\beta{+}\beta_{JJ'})\nonumber\\
&\times&\sum_{\ell=|J'-J|}^{J'+J}4\,j_\ell^2(y)\,C^2(JJ'\ell;00)\,.
\label{sortho}
\end{eqnarray}
\noindent
The coefficients for the even and odd sums are given in the following table:
\begin{center}
\begin{tabular}{lcc}
Type & $A$(even) & $B$(odd) \\ \hline
H para & $a_c^2$ & $a_i^2$ \\
H ortho & $a_c^2/3$ & $a_c^2+2a_i^2/3$ \\
D para & $3a_i^2/4$ & $a_c^2+a_i^2/4$ \\
D ortho & $a_c^2+5a_i^2/8$ & $3a_i^2/8$ \\ \hline
\end{tabular}
\end{center}
\index{para hydrogen}
\index{ortho hydrogen}
\index{para deuterium}
\index{ortho deuterium}
\noindent
Here $a_c$ and $a_i$ are the coherent and incoherent scattering lengths
(note that the characteristic bound cross section is
$\sigma_b{=}4\pi[a_c^2{+}a_i^2]$), $P_J$ is the statistical weight
factor, $\beta_{JJ'}{=}(E_J'{-}E_J)/kT$ is the energy transfer for a
rotational transition, $j_\ell(x)$ is a spherical Bessel function of
order $\ell$, and $C(JJ'\ell;00)$ is a Clebsch-Gordan coefficient.
The parameter $y$ is given by $\kappa a/2{=}(a/2)\sqrt{4MkT\alpha/2}$,
where $a$ is the interatomic distance in the molecule. The translational
weight $w$ is 1/2 for H$_2$ and 1/4 for D$_2$. The sums over $J'$ are
treated as operators into order to keep the notation compact.
Young and Koppel assumed that the molecular translations were
free,\index{free gas}
so the equations contain
\begin{equation}
{\cal S}_f(\alpha,-\beta)=\frac{1}{\sqrt{4\pi\alpha}}
\,\exp\left[-\frac{(\alpha-\beta)^2}{4\alpha}\right],
\end{equation}
\noindent
and
\begin{equation}
{\cal S}_f(\alpha,\beta)=\rm e^{-\beta}{\cal S}_f(\alpha,-\beta)\,,
\end{equation}
\noindent
the free-atom scattering function (with $\beta$ positive). Note that
$\alpha$ is multiplied by a translational weight of 0.5 or 0.25 when
this equation is used in order to make the formula apply to a molecule
with mass ratio 2 or 4, respectively.
These formulas as stated are appropriate for a gas of hydrogen or
deuterium molecules. In a liquid, there are two additional effects to
be considered: interference between the neutron waves scattered from
different molecules, and the fact that the recoil of the hydrogen
molecule is not really free. First, we will consider the latter
effect. Experiments by Egelstaff, Haywood, and Webb at
Harwell\cite{EHW} and Schott at Karlsruhe\cite{Schott} showed
appreciable broadening of the quasi-elastic scattering peak for liquid
hydrogen, and both groups ascribed this to diffusive effects. Later,
Utsuro of Kyoto University constructed a simple analytic
model\cite{Utsuro} that included both diffusion and intermolecular
vibrations and showed good agreement with experiment. More recently,
Keinert and Sax of the University of Stuttgart proposed the
model\cite{Keinert} that we follow here.
They suggested that the free translation term in the Young and Koppel
formulas be replaced by the superposition of a solid-state like motion
and a diffusive law.\index{diffusion} One can picture a hydrogen
molecule bound in a cluster of about 20 molecules and undergoing
vibrations similar to those of a hydrogen molecule in a solid. These
clumps then diffuse through the liquid (hindered translations) according
to the Egelstaff-Schofield effective width model discussed above. The
details of the performance of this model will be discussed further
in a subsequent section.
As mentioned earlier, waves scattered from different molecules can also
interfere. Inter-molecular coherence results when there is a correlation
between the positions of nearby molecules.\index{intermolecular
interference} This kind of coherence is described by the
``static structure factor'' $S(\kappa)$.\index{static structure factor}
This quantity can be used in an approximation due to
Vineyard\index{Vineyard} as follows:
\begin{equation}
\frac{d^2\sigma}{d\Omega d\epsilon}
=\frac{d^2\sigma_{\rm coh}}{d\Omega d\epsilon}\,S(\kappa)
+\frac{d^2\sigma_{\rm incoh}}{d\Omega d\epsilon}\,\,.
\end{equation}
\noindent
This is equivalent to using Eqs.\ref{spara} and \ref{sortho} with
$a_c^2$ replaced by $S(\kappa)a_c^2$ in the calculation of the coefficients
$A$ and $B$. The effects of this procedure will be shown below.
\paragraph{Mixed Moderators.}\index{mixed moderators}
In some cases, thermal evaluations give the scattering for a
principal scatterer as bound in a moderator; for example, H in H$_2$O,
or Zr in ZrH. The other atoms in the molecule are represented by
an analytic law (free-gas O), or by another detailed scattering law
(H in ZrH). In other cases, the scattering from the entire molecule
is represented in one file. Examples are BeO and benzine in
earlier versions of ENDF/B, but only benzine remains as a mixed
moderator in ENDF/B-VII.0. The molecular scattering is renormalized