Abstract
We theoretically analyze the differential cross sections (DCSs) for the state-to-state reaction, H + HD(vi = 0, ji = 0, mi = 0) → H2(vf = 0, jf = 1,2,3, mf = 1,..,jf) + D, over the whole range of scattering angles, where v, j, and m are the vibrational, rotational, and helicity quantum numbers for the initial and final states. The analysis extends and complements previous calculations for the same state-to-state reaction, which had jf = 0,1,2,3 and mf = 0, as reported by Xiahou C.; Connor J. N. L.. Phys. Chem. Chem. Phys. 2021, 23, 13349–13369. Motivation comes from the state-of-the-art experiments and simulations of Yuan et al.. Nature Chem. 2018, 10, 653–65829686377 who have measured, for the first time, fast oscillations in the small-angle region of the degeneracy-averaged DCSs for jf = 1 and 3 as well as slow oscillations in the large-angle region. We start with the partial wave series (PWS) for the scattering amplitude expanded in a basis set of reduced rotation matrix elements. Then our main theoretical tools are two variants of Nearside-Farside (NF) theory applied to six transitions: (1) We apply unrestricted, restricted, and restrictedΔ NF decompositions to the PWS including resummations. The restricted and restrictedΔ NF DCSs correctly go to zero in the forward and backward directions when mf > 0, unlike the unrestricted NF DCSs, which incorrectly go to infinity. We also exploit the Local Angular Momentum theory to provide additional insights into the reaction dynamics. Properties of reduced rotation matrix elements of the second kind play an important role in the NF analysis, together with their caustics. (2) We apply an approximate N theory at intermediate and large angles, namely, the Semiclassical Optical Model of Herschbach. We show there are two different reaction mechanisms. The fast oscillations at small angles (sometimes called Fraunhofer diffraction/oscillations) are an NF interference effect. In contrast, the slow oscillations at intermediate and large angles are an N effect, which arise from a direct scattering, and are a “distorted mirror image” mechanism. We also compare these results with the experimental data.
1. Introduction
The H + H2 → H2 + H reaction and its isotopic variants are important benchmarks in the theory of chemical reaction dynamics. In particular, measurements and calculations of state-to-state differential cross sections (DCSs) can provide detailed information on the dynamics and mechanism of this class of reactions.
Recently, an important experimental advance has been reported by Yuan et al.1 for the H + HD → H2 + D reaction. They have measured, for the first time, fast oscillations in the small-angle region of the degeneracy-averaged DCSs (abbreviated as daDCSs). They reported daDCSs for the following two transitions.
![]() |
R1 |
In reaction (R1), vi, ji and vf, jf are the initial and final vibrational and rotational quantum numbers of the diatomic molecules, respectively. The experiment of Yuan et al.1 used a high-resolution molecular-beam apparatus, crossed at 150°, with a velocity map imaging product detection at a translational energy of 1.35 eV. These daDCS measurements are the current state-of-the-art. Related experimental work can be found in refs (2−5).
The purpose of the present paper is to analyze and quantitatively understand the daDCSs of Yuan et al.1 To do this, we start with the helicity (or body-fixed) representation for the scattering amplitude. We then have to consider the following 16 state-to-state DCSs
![]() |
R2 |
where mi and mf are the helicity quantum numbers for the initial and final states, respectively.
In our earlier paper6 (denoted XC1), which is a companion to this one, we studied theoretically the DCSs for the four state-to-state transitions in reaction (R2) with mf = 0, namely
![]() |
R3 |
In particular, we analyzed the dynamics of the angular scattering for reaction (R3) in order to understand the physical content of the structure in the four helicity-resolved DCSs.6 We discovered: glory scattering at small angles, broad or “hidden” nearside rainbows, Nearside-Farside (NF) interference effects (sometimes called Fraunhofer diffraction/oscillations), a “CoroGlo” test to distinguish corona and forward glory scattering, and a “distorted mirror image” mechanism present at intermediate and large angles.6
In this paper, we focus on the DCSs for state-to-state transitions with nonzero helicities. This reduces the number of DCSs in reaction (R2) to 12. A further reduction is possible because the DCSs for mf = −1, −2, −3 are equal to those for mf = +1, +2, +3, respectively. This leaves the following six DCSs to be analyzed.
![]() |
R4 |
We will often write 000 → 011, 000 → 021, 000 → 031, 000 → 022, 000 → 032, and 000 → 033 for the six transitions or, more simply, 011, 021, 031, 022, 032, and 033.
There is a fundamental difference between DCSs with mf = 0 and those with mf > 0 for reactions of the type (R3) and (R4). All DCSs with mf > 0 are identically equal to zero in the forward (θR = 0°) and backward (θR = 180°) directions in the center-of-mass reference frame, which is a consequence of the conservation of angular momentum. Furthermore, the partial wave series (PWS) for the scattering amplitude for mf = 0 uses a basis set of Legendre polynomials, whereas for mf > 0 the basis set consists of reduced rotation matrix elements (also called Wigner or little d functions), which simplify to associated Legendre functions when mi = 0. This means the theoretical analysis is more complicated and difficult for the mf > 0 case compared to mf = 0.
Now there has been one previous NF analysis of DCSs for chemical reactions with mf > 0, which was made more than 20 years ago.7 In this work, Dobbyn et al.7 made the following important observation (on page 1117):
“...although the PWS becomes more complicated for more general types of collisions, this has little impact on the physical insight provided by a NF analysis”.
Thus, in this paper, (two variants of) NF theory will be used to provide physical insight into the reaction dynamics. Note that the NF theory was used extensively in XC1.6 In particular, an NF analysis has the advantage that the semiclassical (asymptotic) picture is still evident, even though semiclassical techniques, such as the stationary phase or saddle point methods, are not applied. Note that Yuan et al.1 have conjectured on the role an NF analysis plays in explaining oscillatory structures in their DCSs. The two NF theories we use are
-
(1)
For a PWS with a basis set of Wigner functions, we use three NF decompositions: unrestricted (unresNF),8restricted (resNF),9,10 and restrictedΔ (resΔNF).7 The unresNF decomposition is a straightforward generalization8 of the NF decomposition for a Legendre PWS.8,11 The unresNF DCSs incorrectly diverge as θR → 0°, 180°. In contrast, the resNF and resΔNF DCSs correctly go to zero as θR → 0°, 180°.7,9,10
The properties of the caustics of Wigner functions as well as those for reduced rotation matrix elements of the second kind play an important role in the definitions of resNF and resΔNF.7,9,10 We also perform a resummation for a PWS of Wigner functions,12 since it is well-known that a resummation can improve the physical effectiveness of an NF decomposition.13−19 In fact, the present paper is the first time that resummation theory has been combined with the resNF and resΔNF decompositions.
The above remarks apply, in particular, to NF analyses of the full DCSs for the six transitions. We also report the results (including resummations) of the unresNF decomposition for the Local Angular Momentum (LAM), since this provides important additional insights into the reaction dynamics.13−16
-
(2)
A simple approximate N model, the Semiclassical Optical Model (SOM), which was originally introduced by Herschbach.20,21 It is particularly useful for understanding structures in a DCS at intermediate and large angles for direct reactions.6,7
This paper is organized as follows: Section 2 summarizes the partial wave theory and explains our conventions and definitions for the DCSs and LAMs. This section also includes a discussion of the caustic properties that we need and summarizes the unresNF, resNF, and resΔNF decompositions. Section 3 outlines the resummation for a PWS of Wigner functions. The properties of the input scattering matrices for the six transitions are presented in Section 4; we use the same accurate scattering matrix elements employed by Yuan et al.1 in a simulation of their experiments. In Section 5, we discuss in detail the behavior of the unresNF, resNF, and resΔNF DCSs at small and large angles, as this has not been done before. Our results for the full and NF DCSs and LAMs, including resummations, are presented and discussed in Sections 6 and 7, respectively. The SOM DCSs at intermediate and large angles are presented and discussed in Section 8. We report daDCSs in Section 9, where we make comparisons with the experimental data. Our conclusions are in Section 10. Most of our results are presented graphically.
Appendix A proves that the state-to-state mi = 0 DCSs for mf = −1, −2, −3 and mf = +1, +2, +3 are equal, respectively. In applications of the NF theory, it is essential to use unambiguous and consistent definitions for the special functions (of the first and second kinds) employed in the various NF decompositions. In Appendix B, we gather together the precise mathematical definitions that we use, since there is often more than one definition in the literature.
We also emphasize the following: This paper complements and extends XC1,6 where additional discussions and references can be found. These two papers illustrate the potency of the NF theory for divers applications.
2. Partial Wave Theory
2.1. Partial Wave Series
We start with the helicity (or body-fixed) PWS representation of the scattering amplitude for reaction (R4) at a fixed translational (or total) energy22,23
![]() |
1 |
where k0,0 is the translational wavenumber for relative motion in the initial channel, J is the total angular momentum quantum number, S̃000→0jfmfJ is a modified scattering matrix element, and dmf,0(θR) is a reduced rotation matrix element (also called a Wigner function or “little d” function) as defined by Edmonds.24 The reactive scattering angle θR is the angle between the incoming H atom and the outgoing H2 molecule in the center-of-mass reference frame. Thus, θR = 0° and θR = 180° define the forward and backward directions, respectively. In practice, the upper limit of J = ∞ in the PWS is replaced by a finite value, J = Jmax. This assumes that all partial waves with J > Jmax can be neglected. In our applications, there are ∼40 numerically significant coherent partial waves, which makes the direct physical interpretation of the PWS very difficult or impossible. In addition, a constant phase has been omitted from eq 1.
The corresponding state-to-state DCS is given by
![]() |
2 |
The PWS representation (1) is also valid for mf = 0, as further analyzed in detail for the H + HD reaction in XC1;6 however, DCSs with mf = 0 will only be needed in Section 9, when we discuss daDCSs. In passing, we note that the PWS (1) remains valid for mf < 0, provided the starting value of the summation is replaced by J = |mf|. This is only needed in Appendix A. In the remainder of this paper, we will often drop the channel labels from f, k, S̃, σ, etc. to keep the notation simple. We will also write S̃J ≡ S̃J.
To provide additional insight into the reaction dynamics, we also perform a Local Angular Momentum analysis.13−16 The LAM analysis provides information on the total angular momentum variable that contributes to the scattering at an angle θR under semiclassical conditions. It is defined by
![]() |
3 |
Note that the arg in eq 3 is not necessarily the principal value in order that the derivative be well-defined.
Next we describe the unrestricted NF decomposition for the full PWS (1), which is the simplest NF decomposition, and we point out a limitation when mf > 0.
2.2. Unrestricted Nearside-Farside Decomposition (unresNF)
We exactly decompose f(θR) by writing it as the sum of two subamplitudes N and F, namely8
![]() |
4 |
This is accomplished by exactly decomposing the dmf,0J(θR) in eq 1 into traveling angular functions of degree J and order mf
![]() |
5 |
where, for θR ≠ 0,π
![]() |
6 |
In eq 6, the emf,0J(θR) are reduced rotation matrix elements of the second kind (also known as “little e” functions) and defined in Appendix B. We see in eq 6 that the dmf,0(θR) are linear combinations of reduced rotation matrix elements of the first and second kinds or, equivalently, from Appendix B, a linear combination of Jacobi functions of the first and second kinds.
Using eqs 4–(6), the N and F subamplitudes are given by (θR ≠ 0,π)
![]() |
7 |
which is called the unrestricted NF decomposition. The adjective “unrestricted” is added because eq 7 can be used for all θR ∈ (0,π) with no restriction on the sum over J. We also sometimes write unresNF. The corresponding N and F DCSs are given by
![]() |
8 |
With the help of eqs 2 and (8), we obtain
![]() |
9 |
Equation 9 is the Fundamental Identity for Full and NF DCSs and is exact.25
Similarly, we can define (unrestricted) N and F LAMs
![]() |
10 |
There is an exact Fundamental Identity for Full and NF LAMs analogous to eq 9, although more complicated in form.25 As before, the args in eq 10 are not necessarily principal values in order that the derivatives be well-defined.
However, there is a problem with the unrestricted decomposition for mf > 0. Although the unresNF decomposition (5)–(7) is mathematically exact, its physical usefulness requires that the dmf,0J(θR) and emf,0(θR) oscillate as θR varies in the range of (0°,180°). Figures 1 and 2 examine this point by showing plots of the little d and little e functions, respectively, versus θR/ deg for J = 10, mi = 0, and (a) mf = 0 (the Legendre case), (b) mf = 1, (c) mf = 2, (d) mf = 3. Note that the little d function has J – mf zeros and the little e function has J – mf + 1 zeros. We make the following observations about Figures 1 and 2:
We see that the little e function diverges as θR → 0°, 180°, which means that the N,F components of eq 6 also diverge. Then we have the unfortunate situation in the interesting forward and backward regions that σ(N,F)(θR)→∞, whereas σ(θR) → 0; i.e., although the NF decomposition (5)–(7) is mathematically exact, it is not physically meaningful at small and large angles.
We see there are angular regions where the little d and little e functions are oscillatory (which can be called classically allowed regions) separated from two nonoscillatory regions (classically forbidden regions) when θR is close to 0°, 180°. We can distinguish between these regions using the notion of caustics, which are discussed next.
Figure 1.
Plots of dmf, miJ (θR) vs θR/deg for J = 10, mi = 0 (a) mf = 0, (b) mf = 1, (c) mf = 2, (d) mf = 3. The vertical pink lines indicate the caustic angles at (a) 0°, 180°, (b) 5.7°, 174.3°, (c) 11.5°, 168.5°, (d) 17.5°, 162.5°.
Figure 2.
Plots of emf, miJ (θR) vs θR/deg for J = 10, mi = 0 (a) mf = 0, (b) mf = 1, (c) mf = 2, (d) mf = 3. The vertical pink lines indicate the caustic angles at (a) 0°, 180°, (b) 5.7°, 174.3°, (c) 11.5°, 168.5°, (d) 17.5°, 162.5°.
2.3. Caustic Properties of dmf,0J(θR) and emf,0(θR)
The boundaries between the two classically forbidden regions and the classically allowed region in Figures 1 and 2 can be conveniently characterized by two caustic angles,9,10 denoted θR min(J,mf) and θR max, for given values of J and mf. They are defined by the divergence of the primitive Wentzel-Kramers-Brillouin (WKB) approximation for two linearly independent solutions of the second-order differential equation satisfied by the little d and little e functions, which become the associated Legendre differential equation because mi = 0.
The caustic angles can be found from eq (5.13) of ref (26), and they are determined by sin θR = mf/J, which results in
![]() |
11 |
For J = 10, the caustics occur at: θR min(10,mf) = 0° (mf = 0), 5.7° (mf = 1), 11.5° (mf = 2), 17.5° (mf = 3). The corresponding values for θR max are 180° (mf = 0), 174.3° (mf = 1), 168.5° (mf = 2), 162.5° (mf = 3). These caustic angles are marked on Figures 1 and 2 as vertical pink lines. Note that the caustics for the Legendre case (mf = 0) are always at 0° and 180° for all values of J ≥ 1. The caustic angles are shown in a different way in Figure 3 on a (θR/deg, J) plot9,10 for mf = 1,2,3 and J = 1(1)30. This figure shows clearly that θ R min(J,mf) → 0 and θ R max → π as J increases for a fixed value of mf.
Figure 3.
Values of θR min(J, mf)/deg and θR max/deg (black solid circles) on a (θR/deg, J) plot for mi = 0 and mf = 1,2,3. Passing through the black solid circles are the curves, J = mf/sin θR, colored red for mf = 1, orange for mf = 2, and green for mf = 3.
The above discussion implies that the dmf,0J(N,F)(θR) values only exhibit an oscillatory behavior (for a given J and mf) in the angular range
![]() |
12 |
This in turn implies that the NF decomposition (7) should work best when θR satisfies the inequality (12).
An inspection of Figure 3 shows, for given values of θR, J, and mf, that there is a minimum value of J, denoted Jmin(mf)(θR), such that θR satisfies the inequality (12). For mf > 0, we have
![]() |
13 |
where int(x) ≡ integer part of x. Sometimes, +1 is added to the right-hand-side of eq 13 to exclude the case where J = mf. In practice, it makes little difference whether +1 is added or not.9,10 We confirmed this is the case in our calculations for all six transitions. The physical reason is that the PWS (1) receives its main numerical contribution from partial waves with J ≫ mf, helped by the (2J + 1) factor.
The comments just given lead us to introduce the restricted nearside-farside decomposition, denoted resNF, which we discuss next.
2.4. Restricted Nearside-Farside Decomposition (resNF)
The decomposition in which partial waves with J < Jmin(mf)(θR) are omitted from eqs 1 and (7) defines resNF.9,10 The restricted N,F subamplitudes are given by
![]() |
14 |
And the corresponding resNF DCSs are
![]() |
15 |
Note that resNF is an approximate decomposition because it omits partial waves from classically forbidden regions of θR; that is, it neglects the following terms in the PWS (1)
![]() |
16 |
The contribution of each partial wave in eq 16 is nonoscillatory and small in magnitude. Notice that eqs 13 and (14) are, in general, discontinuous functions of θR. As a result, the corresponding DCSs also exhibit discontinuities, although, as we shall see in Section 5, they are usually small and confined to small and large angles. In addition, there is no global LAM for resNF because of the discontinuities. Having identified the Δf(θR) term of eq 16, we can include it in the resNF decomposition—this gives rise to the restrictedΔ nearside-farside decomposition, denoted resΔNF, which we discuss next.
2.5. RestrictedΔ Nearside-Farside Decomposition (resΔNF)
The resΔNF decomposition is obtained when we combine eq 16 with eq 14 to obtain an improved version of resNF. We have for the subamplitudes7
![]() |
17 |
And the corresponding resΔNF DCSs are
![]() |
18 |
Notice that resΔNF is an exact NF decomposition, unlike eq 14, which is approximate. Similar to resNF, eq 17 is also a discontinuous function of θR, although the discontinuities are usually small in the corresponding DCSs and confined to small and large angles—examples are provided in Section 5. In addition, there is again no global LAM for resΔNF because of these discontinuities.
Note: If, for a given θR, we have that Jmin(mf)(θR) is equal to mf, then unresNF, resNF, and resΔNF become equivalent.
Practical Remark
It can often happen that Jmin(mf)(θR), for particular values of θR and mf, can exceed Jmax, that is, Jmin(θR) > Jmax. For example, Jmin(mf = 3)(θR = 0.1°) = 1718, but Jmax = 40. Then many computer programs applied directly to eqs 14–(18) will crash as they attempt to use values of S̃J that are undefined for J > Jmax, when the upper limit of J = ∞ has been replaced by J = Jmax. This problem can be avoided by adding sufficient S̃J ≡ 0 to the PWS for J > Jmax.
3. Resummation of the Partial Wave Series
It is well-established that a resummation of a Legendre PWS can significantly improve the physical effectiveness of an NF decomposition.13−19 Totenhofer et al.19 have provided an extensive discussion of the Legendre case. This same improvement in NF physical effectiveness occurs for a basis set of little d functions, although this has only been studied for a single example, namely, Ar + HF rotationally inelastic scattering.12
It has been found previously that the biggest effect for cleaning the N,F DCSs and N,F LAMs of unphysical oscillations occurs on going from resummation order, r = 0 (no resummation, i.e., eq 1) to resummation order, r = 1. There is usually a smaller cleaning effect for further resummations, r = 1 to r = 2 and r = 2 to r = 3.
Whiteley et al.12 have resummed the PWS (1), which we now write as fr=0(θR), from r = 0 to r = 1. We do not repeat the derivation here, which exploits the recurrence relation obeyed by the little d functions; rather, we simply write down the final result for the resummed representation for fr=1(θR). From eq (3.9) of ref (12) with mi = 0, we have
![]() |
19 |
where β ≡ β1 ≡ β1(r=1) is the resummation parameter, and
![]() |
20 |
with
![]() |
21 |
and
![]() |
22 |
![]() |
23 |
Equation 19 also assumes that (1 + β cos θR) ≠ 0. Notice that eq 20 is valid for J = mf, as proven in the Appendix of ref (12). For this case, we see from eq 22 that gmfmf = 0.
An NF decomposition of eq 19 can now be made
![]() |
24 |
with the corresponding N,F subamplitudes given by (θR ≠ 0,π)
![]() |
25 |
and the corresponding N,F r = 1 DCSs are
![]() |
26 |
Similar equations apply to the resNF and resΔNF decompositions. For the unrestricted NF decomposition, the NF r = 1 LAMs are given by
![]() |
Notice that the full amplitudes, fr=0(θR) and fr=1(θR), are independent of β and numerically the same for a given value of θR. This is also true for the full LAMs, LAMr=0(θR) and LAMr=1(θR).
In our applications, we need to choose a value for the resummation parameter β. We extend the prescription used by Anni et al.13 and solve the linear equation
![]() |
27 |
This results in
![]() |
These values are used in the resummation calculations of Sections 5–7. The general result is
![]() |
4. Properties of the Input Scattering Matrix Elements
We use the same S matrix elements that were computed by Yuan et al.1 and used for the mf = 0 analyses in XC1.6 The Boothroyd-Keogh-Martin-Peterson potential energy surface number two (BKMP2) was employed.27 Converged S matrix elements were obtained for translational energies Etrans up to 3.5 eV. All our results are for Etrans = 1.35 eV, which is the same translational energy as that employed in the molecular-beam experiments. The masses used are mH = 1.0078 u and mD = 2.0141 u, with the initial translational wavenumber being k = 11.692 a0–1. For each transition, Jmax is ∼40.
Figure 4 shows graphs of |S̃J| versus J for the three transitions 000 → 011, 000 → 021, 000 → 031, while Figure 5 shows plots for the remaining three transitions 000 → 022, 000 → 032, 000 → 033. Figures 6 and 7 display the corresponding plots for arg S̃J/rad versus J. Note that all the curves start at J = mf. A perusal of Figures 4–7 reveals the following:
For five of the transitions, the global maximum in an |S̃J| plot is at the first peak as J increases from J = mf. The exception is the 031 case, where the maximum occurs at the second peak. The peaks are then followed by subsidiary local maxima; these play an important role in the interpretation of the intermediate- and large-angle scattering using the SOM in Section 8. The overall shapes of the mf = 1,2,3 curves in Figures 4 and 5 are similar to those for the four mf = 0 transitions, with the exception that the global maxima of the |S̃J| curves are always at J = 0 when mf = 0 (see Figure 1 of XC1 (i.e., ref (6))).
Figures 6 and 7 show that the plots of arg S̃J/rad versus J are roughly quadratic in shape. The kinks in some of the curves are seen to correspond to near-zeros in |S̃J|, where the phase of S̃J varies more rapidly with J. The curves in Figures 6 and 7 have similar properties to the arg S̃J/rad plots for mf = 0 (see Figure 2 of XC1 (i.e., ref (6))).
Note that, in the NF analysis, only the values of S̃J at J = 0,1,2,... are used. To help guide the eye, the points (black solid circles) in Figures 4–7 have been joined by straight lines. This was also done in Figures 1 and 2 of XC1.6 When we want a smooth continuation of the {S̃J} to real values of J, for example, for use in an asymptotic (semiclassical) analysis, we would typically use a cubic B-spline interpolation.6 Notice also that the kinks do not affect the NF analysis nor the asymptotic analysis, as explained in XC1.6
Figure 4.
Plots of | S̃J | vs J at Etrans = 1.35 eV. The black solid circles are the numerical S matrix data, {| S̃J |}, at integer values of J, which have been joined by straight lines. The transitions are (a) 000 → 011, (b) 000 → 021, and (c) 000 → 031.
Figure 5.
Plots of | S̃J | vs J at Etrans = 1.35 eV. The black solid circles are the numerical S matrix data, {| S̃J |}, at integer values of J, which have been joined by straight lines. The transitions are (a) 000 → 022, (b) 000 → 032, and (c) 000 → 033.
Figure 6.
Plots of arg S̃J/rad vs J at Etrans = 1.35 eV. The black solid circles are the numerical S matrix data, {arg S̃J/rad}, at integer values of J, which have been joined by straight lines. The transitions are (a) 000 → 011, (b) 000 → 021, and (c) 000 → 031.
Figure 7.
Plots of arg S̃J/rad vs J at Etrans = 1.35 eV. The black solid circles are the numerical S matrix data, {arg S̃J/rad}, at integer values of J, which have been joined by straight lines. The transitions are (a) 000 → 022, (b) 000 → 032, and (c) 000 → 033.
We next consider in more detail the properties of the unresNF, resNF, and resΔNF decompositions.
5. Properties of the Unrestricted, Restricted and RestrictedΔ Nearside-Farside Decompositions Including Resummations
In Sections 2.2, 2.4, and 2.5, we developed the theory for the unresNF, resNF, and resΔNF decompositions, respectively, for r = 0; the extension of the theory to r = 1 was given in Section 3. In the present section, we investigate in detail how these three decompositions (including resummations) influence the corresponding N,F DCSs at small and large angles, as this has not been investigated before. In Figure 8, we plot four N DCSs for the 000 → 011 transition at large angles, namely, for θR = 140◦–180◦. The upper panel shows DCSs for r = 0, and the lower panel shows DCSs for r = 1.
Figure 8.
Plots of four PWS N DCSs in the large-angle region from θR = 140° to θR = 180° for the transition 000 → 011 at Etrans = 1.35 eV for (a) r = 0 and (b) r = 1. Purple long-dashed curve: PWS/N/unres. Lilac dashed curve: PWS/N/res. Red solid curve: PWS/N/resΔ. Red dashed curve: Least-squares-fit to PWS/N/resΔ.
We begin our discussion with the resN DCS (lilac dashed curve) and the resΔN DCS (red solid curve) in Figure 8. We note the following:
By construction, the resN and resΔN DCSs tend to zero as θR →180◦. Their discontinuities are clearly visible on the scale of the drawings. The density of jumps increases as θR →180◦; this is expected from Figure 3.
The resΔN DCS is usually smaller than the resN DCS for both r = 0 and r = 1. Now both the resN subamplitude and the Δf(θR)/2 term in eq 17 are complex-valued quantities, which means that destructive interference can occur, resulting in the resΔN DCS being smaller than the resN DCS.
-
The first discontinuity for increasing θR occurs at θR ≈ 150° for r = 0 but at θR ≈ 161◦ for r = 1. This behavior can be understood because Jmin(mf=1)(θR), which is equal to int(mf/sin θR) by eq 13, jumps from J = 1 at θR ≈ 149.9° to J = 2 at θR ≈ 150.0°, causing a discontinuity in the PWS (14) and in the resulting resΔN and resN DCSs.
In contrast, for r = 1, the J = mf = 1 term is put equal to zero by the choice of β in eq 27, resulting in the PWS (25) starting at J = mf + 1 = 2. Then the first jump occurs for J = 2 at θR ≈ 160.5° to J = 3 at θR ≈ 160.6°.
Because of congestion in the graphs, it is difficult for the eye to follow the jumps in the resΔN and resN DCSs in Figure 8, especially as θR → 180°. However, it is the general trend in these DCSs that is of interest. We therefore made least-squares-fits to the resΔN DCSs. These are shown as red dashed curves for r = 0 and r = 1 in Figure 8a,b, respectively.
We also plotted the unresN DCSs (purple long-dashed curves) for r = 0 and r = 1 in Figure 8. As expected, they tend to infinity as θR → 180°. The beneficial effect of cleaning can be seen because the unresN r = 0 DCS starts to diverge at θR ≈ 150°, whereas for r = 1, the unresN DCS diverges at a larger angle, namely, θR ≈ 170°.
The results discussed above for the 000 → 011 transition have all been for N DCSs at large angles. We also did a similar analysis for the N DCSs at small angles and obtained analogous results (not shown). In addition, we also calculated unresF, resF, and resΔF DCSs at large and small angles and obtained comparable results (also not shown). Finally, we performed unresNF, resNF, and resΔNF analyses for the remaining five transitions at large and small angles, again finding similar results (not shown) to the 011 case.
The least-squares-fits to the resΔNF DCSs are used in the next section, where we report NF analyses of the full DCSs for all the transitions.
6. Full and Nearside-Farside DCSs Including Resummations
Figure 9 shows logarithmic plots of the full and resΔN, resΔF r = 1 DCSs versus θR for the 000 → 011, 000 → 021, and 000 → 031 transitions. The corresponding DCSs for the 000 → 022, 000 → 032, and 000 → 033 transitions are displayed in Figure 10. For clarity of viewing, notice that, at large and small angles, least-squares-fits to the resΔN and resΔF DCSs are plotted, as explained in Section 5. We use the following color conventions for the DCSs in Figures 9 and 10 as well as in some other figures.
Full PWS: black solid, with the label, “PWS”.
resΔN r = 1 PWS: red solid, with the label, “PWS/N/resΔ”.
Fit to resΔN r = 1 PWS: red dashed, with the label, “Fit to PWS/N/resΔ”.
resΔF r = 1 PWS: blue solid, with the label, “PWS/F/resΔ”.
Fit to resΔF r = 1 PWS: blue dashed, with the label, “Fit to PWS/F/resΔ”.
Figure 9.
Plots of log σ(θR) vs θR/deg at Etrans = 1.35 eV for r = 1. Black curve: PWS. Red solid curve: PWS/N/resΔ. Blue solid curve: PWS/F/resΔ. Red dashed curves: least-squares-fits to PWS/N/resΔ in the small- and large-angle regions. Blue dashed curves: Least-squares-fits to PWS/F/resΔ in the small- and large-angle regions. The transitions are (a) 000 → 011, (b) 000 → 021, and (c) 000 → 031.
Figure 10.
Plots of log σ(θR) vs θR/deg at Etrans = 1.35 eV for r = 1. Black curve: PWS. Red solid curve: PWS/N/resΔ. Blue solid curve: PWS/F/resΔ. Red dashed curves: least-squares-fits to PWS/N/resΔ in the small- and large-angle regions. Blue dashed curves: least-squares-fits to PWS/F/resΔ in the small- and large-angle regions. The transitions are (a) 000 → 022, (b) 000 → 032, and (c) 000 → 033.
We first examine the full DCS for the 000 → 011 transition in Figure 9a. As θR increases from 0° to 180°, we observe the following.
The DCS = 0 a02 sr–1 at θR = 0° followed by the next observation listed here.
Fast oscillations in an angular range extending up to θR ≈ 50°, accompanied by a decreasing DCS. This behavior merges into the next observation listed here.
An increasing DCS with slow oscillations, which extend into the large-angle region.
The DCS = 0 a02 sr–1 at θR = 180°.
The full DCSs for the remaining five transitions exhibit similar properties to the 011 case and are not discussed separately. We can also compare with the four full DCSs for the mf = 0 case shown in Figure 3 of XC1.6 We see that the mf = 0 and mf > 0 DCSs are alike, the main difference being (a) the mf = 0 DCSs are nonzero at θR = 0°,180° unlike the mf > 0 DCSs, (b) the angular regions separating the fast and slow oscillations are slowly varying for mf = 0, whereas there are pronounced minima when mf > 0.
Next, we examine the resΔN, resΔF r = 1 DCSs in Figures 9 and 10, making use of the exact Fundamental Identity for Full and N,F DCSs given by eq 9, which is also valid for the r = 1 case.25 In angular regions where there are fast oscillations, we see that the resΔN and resΔF r = 1 DCSs are varying relatively slowly with θR, which tells us that the fast oscillations in the full DCSs arise from NF interference. Another name for the fast oscillations is Fraunhofer diffraction/oscillations. In contrast, the slow oscillations are seen to be resΔN-dominated. Thus, we have the important result from the NF analysis that the fast and slow oscillations arise from different physical mechanisms. This is also the case for the mf = 0 DCSs.6
We can extract useful information from the periods ΔθR of the fast oscillations. A simple NF model shows that these oscillations are analogous to the interference pattern from the well-known “Young’s double-slit experiment”, as explained in a molecular scattering context in Appendix A of ref (28). This analogy was also used in XC1,6 and it yields the simple relation
![]() |
28 |
where Jeff is an effective total angular momentum variable characteristic of the NF oscillations. For example, for the mf = 0 DCSs, we have Jeff = Jg + 1/2, where Jg is the glory angular momentum variable, defined as the position of a local maximum in a plot of arg S̃(J)/rad versus J (see Figure 2 of XC16). Figures 9 and 10 show that ΔθR usually lies in the range of ΔθR = 6°–7°, which is similar to the mf = 0 DCSs. Then eq 28 gives Jeff = 30.0–25.7. An inspection of Figures 6 and 7 shows that these values for Jeff are also close to a local maximum in the arg S̃(J)/rad plots.
7. Full and Nearside-Farside LAMs Including Resummations
A full and N,F LAM analysis provides information on the value of the total angular momentum variable that contributes to the scattering at an angle θR, under semiclassical conditions. An important tool16,25 for interpreting a LAM plot is the exact Fundamental Identity for Full and N,F LAMs, which is also valid for r = 1 and is analogous to the identity for DCSs given by eq 9.
Figure 11 shows a full and N,F LAM plot for the 000 → 011 transition using the unresNF decomposition for r = 0 and r = 1. We first make the following observations on the full LAM.
The full LAM shows oscillations at small angles. At intermediate and larger angles it becomes monotonic and increases except for θR around 140°.
The full LAM changes from F dominance to N dominance as θR increases in the small-angle region. This is the same behavior shown by the F and N r = 1 DCSs in Figure 9a.
The spike at θR ≈ 46.2° corresponds to the minimum in the full DCS—see Figure 9a. Thus, the full LAM plot provides a clear indication of a change in the mechanism for the reaction as θR increases.
Figure 11.
Plots of LAM (θR) vs θR/deg at Etrans = 1.35 eV for the 000 → 011 transition, showing results for both r = 0 and r = 1. Black curve: PWS. Red solid curve: PWS/N/r = 1. Red dashed curve: PWS/N/r = 0. Blue solid curve: PWS/F/r = 1. Blue dashed curve: PWS/F/r = 0. The fainter blue solid and dashed curves show where the F LAM(θR) is not physically significant.
Next, we examine the N and F r = 0,1 LAMs in Figure 11 and note the following.
The effect of cleaning the N,F r = 0 LAMs is very striking, with nonphysical oscillations for θR ≲ 50° being replaced by slower variations in the N,F r = 1 LAMs.
It is known that N and F LAMs that are nearly zero, or oscillate about zero, are nonphysical.13,15,16 It can be seen in Figure 11 that this occurs for both the r = 0 and r = 1 F LAMs when θR ≳ 50°. The corresponding curves are drawn in a fainter blue compared to the F LAMs for θR ≲ 50°.
An averaging of the N and F r = 1 LAMs for 10° ≤ θR ≤ 45° gives −28.1 and 28.2, respectively. The value J ≈ 28 is consistent with the information obtained from the periodicity of the fast oscillations in Section 6. An inspection of Figure 6a shows that J ≈ 28 is close to a local maximum in the arg S̃(J)/rad plot.
The N r = 1 LAM for θR ≳ 50° is close to the full N LAM. And both of them monotonically increase (except for θR ≈ 140°) and are similar to the LAM for a hard-sphere collision.13 This implies that the SOM model, which is an approximate N theory incorporating hard-sphere dynamics, should be approximately valid at intermediate and large scattering angles. This point is confirmed in Section 8.
The properties of the full and N,F r = 0,1 LAMs for the other transitions are similar to those for the 011 case and are not shown separately. Overall, we can say that the information given by the LAM analysis is consistent and complementary to that in the DCS plots of Figures 9 and 10.
8. Semiclassical Optical Model (SOM) DCSs at Intermediate and Large Angles
The SOM is a simple procedure, introduced by Herschbach,20,21 for calculating the DCSs of state-to-state reactions. In XC1,6 we applied the SOM to the four mf = 0 transitions and showed that the SOM provided valuable insights into structures in the DCSs at intermediate and large angles. In particular, we found that the SOM and PWS DCSs are distorted mirror images of the corresponding PJ ≡ |S̃J|2 versus J plots, with J = 0,1,2,.... The theory for the SOM has been given in XC1,6 and below we just state the working equations when mf > 0.
The SOM DCS is given by
![]() |
29 |
with PJ ≡ P(J) and
![]() |
30 |
where J = mf, mf + 1,.... In eqs 29 and (30), d is the sum of the radii of two hard spheres representing the reactants and is the only adjustable parameter in the theory. The above equations assume that J ≤ kd; otherwise, σSOM (θR) ≡ 0. Notice that the SOM only depends on the value of the modulus |S̃J| and is independent of arg S̃J. In NF terminology, the SOM is an approximate N theory, which should work best for direct rebound reactions, in particular, at intermediate and backward angles in the DCS.
The SOM and PWS DCSs are compared in Figures 12 and 13 for the six transitions in the range of θR = 50°–180°. Now, for the mf = 0 case, we obtained values of d by fitting the SOM DCS to the PWS DCS at, or close to, θR = 180°. This does not work for mf > 0 because the PWS DCSs are equal to 0 a02 sr–1 at θR = 180°. Instead, we obtained d by fitting the SOM DCS at, or close to, the PWS peak nearest to θR = 180°. An exception is the 000 → 031 transition in Figure 12c, for which the second nearest peak was used (note, this DCS exhibits the most detailed structure out of the six transitions). The values we used for d are given in the figures.
Figure 12.
Plots of σ (θR) vs θR/deg at Etrans = 1.35 eV for the angular range from θR = 50° to θR = 180°. Black curve: PWS. Red curve: SOM. The transitions are (a) 000 → 011, (b) 000 → 021, and (c) 000 → 031.
Figure 13.
Plots of σ (θR) vs θR/deg at Etrans = 1.35 eV for the angular range from θR = 50° to θR = 180°. Black curve: PWS. Red curve: SOM. The transitions are (a) 000 → 022, (b) 000 → 032, and (c) 000 → 033.
It can be seen in Figures 12 and 13 that the SOM reproduces the main features in the PWS DCSs, with larger deviations as the PWS DCSs become more structured. This is encouraging, considering the simplicity of the SOM, and, like the mf = 0 case, it tells us that the SOM and PWS DCSs are distorted mirror images of the corresponding PJ versus J plots. As expected, the SOM does not reproduce the NF interference (or Fraunhofer) oscillations in the PWS DCSs for θR ≲ 50° (not shown). Finally, we note that the values for d lie in the range of 1.44–1.97a0, which are much less than the sum of the radii at the saddle point for the BKMP2 potential energy surface, which is d‡ = rHH‡ + rHD‡ = 3.514 a0. This tells us, as was also found for mf = 0, that the scattering at intermediate and large angles arises mainly from small values of J or, equivalently, from small-impact parameters.
9. Degeneracy Averaged Differential Cross Sections (daDCSs)
In this section, we calculate degeneracy averaged DCSs (daDCSs) and compare with the experimental daDCSs for the two transitions vi = 0, ji = 0 → vf = 0, jf = 1 and vi = 0, ji = 0 → vf = 0, jf = 3. The usual definition of a daDCS is
![]() |
31 |
In our applications, we have a single initial state, namely, vi = 0, ji = 0, mi = 0, so eq 31 simplifies to
![]() |
32 |
A further simplification is possible when mi = 0 because, as shown in Appendix A, DCSs for mf = −1,–2,–3 are equal to those for mf = +1,+2,+3, respectively. We can write eq 32 in the form
![]() |
33 |
where the sum is zero if jf = 0. Equation 33 can also be written in a more compact way, namely
![]() |
34 |
where the prime on the Σ sign means “multiply the first term in the sum by 1/2”.
We can also substitute eqs 1 and (2) into eq 34, obtaining
![]() |
35 |
Equation 35 is a more explicit version of eq 1 of Yuan et al.1
Figure 14a compares the calculated daDCS using eq 35 for the transition 00 → 01 with the experimental data. The corresponding results for the 00 → 03 transition are in Figure 14b. A single scaling factor has been applied to the experimental data to compare with the calculations.1 The results in Figure 14a,b are an extension of the corresponding figures of Yuan et al.1 because we included estimated experimental uncertainties. These are a 10% error in the measurements and an angular uncertainty of 1.5°.1 It can be seen that the agreement between the calculated and experimental daDCSs is very good, in particular, for the NF interference (Fraunhofer) oscillations at θR ≲ 40°; these are shown in more detail in the insets. In the experiments, also note that ∼97% of the HD molecules in the molecular beam are in their ground state, and the translational energy uncertainty is ∼1.2%.1
Figure 14.
Plots of degeneracy averaged σ(θR) (daDCS) vs θR/deg at Etrans = 1.35 eV. (a) The transition 00 → 01 (red), together with experimental results and their estimated errors (blue). (b) The transition 00 → 03 (purple), together with experimental results and their estimated errors (blue). (c) Black curve: Degeneracy averaged, σ(θR), for the 00 → 0 transition, which is summed over jf = 0,1,2,3. The four colored curves show the degeneracy averaged σ(θR) for the transitions 00 → 00 (orange), 00 → 01 (red), 00 → 02 (green), and 00 → 03 (purple).
If an experiment cannot resolve individual jf states, then it is necessary to sum over these states. We have
![]() |
36 |
Figure 14c shows a plot of σ00→0(θR) versus θR over the whole angular range. It can be seen that the structure in σ00→0(θR) (black curve) is largely washed out, even though the individual σ00→0jf(θR) in eq 36 (colored curves) possess distinct fast and slow oscillations, although less pronounced than the helicity-resolved DCSs in Figures 9 and 10. We can also relate our results and notations to the figures in the Supporting Information (SI) and main text of ref (1), with the following clarifications noted.
-
(a)
Figure 6(c) (SI) is a dimensionless plot, as is Figure 5c (main text).1
-
(b)
The three curves for K′ = 1,2,3 in Figure 5b (main text) show 2 × σ000→03K′(θR).1
-
(c)
The units for the labels on the ordinates of Figures 6(a), 6(b), and 7 (SI) are a02 sr–1.1
-
(d)
The red curve for K′ = 1 in Figure 6(b) (SI) shows 2 × σ000→011(θR).1
-
(e)
The red curve for K′ = 1 in Figure 7(c) (SI) shows 2 × 103× [σ000→011(Jmax,θR) – σ000→011(Jmax −1,θR)] versus Jmax at θR = 4°.1 Here Jmax makes explicit the finite upper value for the PWS when used in eqs 1 and (2).
-
(f)
The blue curve for K′ = 0 in Figure 7(c) (SI) shows 103× [σ000→010(Jmax,θR) – σ000→010(Jmax −1,θR)] versus Jmax at θR = 0.4°, not 4°.1
-
(g)
The three curves for K′ = 1,2,3 in Figure 7(d) (SI) show 2 × 103× [σ000→03K′(Jmax,θR) – σ000→03K′(Jmax – 1,θR)] versus Jmax at θR = 6°.1
-
(h)
The blue curve for K′ = 0 in Figure 7(d) (SI) shows 103 × [σ000→030(Jmax,θR) – σ000→030(Jmax – 1,θR)] versus Jmax at θR = 6°.1
10. Conclusions
We have theoretically analyzed structures in the DCSs of the ground-state reaction H + HD → H2 + D for the product states 011, 021, 031, 022, 032, and 033. The calculations extend and complement our previous analyses in XC16 for the cases 000, 010, 020, and 030, making 10 DCSs in all. The motivation comes from the experiments and simulations of Yuan et al.,1 who have measured for the first time fast oscillations in the small-angle region of the daDCSs for jf = 1 and 3 as well as slow oscillations in the large-angle region.
Our main theoretical tools were two variants of Nearside-Farside theory: (1) We applied unrestricted, restricted, and restrictedΔ NF decompositions, including resummations, to the helicity PWS, which is expanded in a basis set of little d functions. We analyzed in detail the properties of restricted and restrictedΔ NF DCSs and showed that they correctly go to zero in the forward and backward directions when mf > 0, unlike the unrestricted NF DCSs, which incorrectly go to infinity. We also calculated LAMs to obtain further insights into the reaction dynamics. Properties of little e functions played an important role in the NF analysis, as do the caustics associated with the little d and little e functions. (2) We applied an approximate N theory at intermediate and large angles, namely, the Semiclassical Optical Model.
We showed that the fast oscillations at small angles (sometimes called Fraunhofer diffraction or oscillations) arise from an NF interference effect. In contrast, the slow oscillations at large angles are an N effect and arise in the DCS as a distorted mirror image of the corresponding PJ versus J plot. We also compared with the experimental daDCSs, obtaining very good agreement.
Our analyses confirm the earlier insight of Dobbyn et al.7 that as the PWS increases in complexity, this has little impact on the physical insight provided by an NF analysis.
Acknowledgments
We thank Professor Z. Sun (State Key Laboratory of Molecular Reaction Dynamics, Dalian Institute of Chemical Physics, Chinese Academy of Sciences, Dalian 116023, Liaoning, People’s Republic of China) for his help and for supplying scattering matrix elements. Support of this research is gratefully acknowledged by the P.R. China Project of Shandong Province Higher Educational and Technology Program, 2017, Project No. J17KB068, and the Zibo School and City Integration Platform Project–Applied Pharmaceutical Innovation Platform (Project No. 2018ZBC423).
Appendix A
This Appendix proves the following identity for PWS DCSs when mf = 1,2,3,...,jf (the identity is trivially true for mf = 0).
![]() |
A1 |
That is, the DCSs for mf and −mf are equal when mi = 0. Now the vibrational quantum numbers vi and vf do not change in the following, so they will be omitted from now on to simplify the notation. Note that all quantum numbers are integers.
We begin by writing the helicity or body-fixed S matrix element S̃jimi → jfmfJ as a linear combination of space-fixed S matrix elements, which are labeled by J, ji, jf and by li, lf, the initial and final orbital angular momentum quantum numbers, respectively. We have9,22,23
![]() |
A2 |
where ⟨j1m1, j2m2| jm⟩ is a Clebsch-Gordan coefficient with m = m1 + m2.
For mi = 0, eq (A2) simplifies to
![]() |
A3 |
We next replace mf > 0 by −mf < 0 in eq (A3) to obtain
![]() |
A4 |
Now we have the relation [ref (24), page 42, eq (3.5.17)]
![]() |
![]() |
A5 |
Next we note, from ref (24), page 46, eq (3.7.3), and page 49, eq (3.7.14), that ⟨ji 0, J 0 | li0 ⟩ is zero unless ji + J + li = an even number, which means we can introduce a factor of +1 = (−1)ji+J+li into eq (A5). In addition, there is the conservation of parity relation, (−1)ji+li = (−1)jf+lf, so for the phase factor we find (−1)jf+J+lf = (−1)ji+J+li = +1 and obtain
![]() |
A6 |
Comparing the right-hand sides of eq A3 and (A6) we see they are the same, so
![]() |
A7 |
Now the PWS (1) for the scattering amplitude remains valid for negative helicities, provided the starting value of the summation is replaced by J = |mf|. When mf > 0 is replaced with −mf < 0, in eq 1 we get
![]() |
Finally with the help of the relation [ref (24), page 60, eq (4.2.5)]
![]() |
together with the result, eq (A7), we find that the moduli of the scattering amplitudes for mf and −mf are equal, which then gives us the identity, eq (A1).
Appendix B
In the development and application of NF theory (NFology), it essential to use unambiguous and consistent definitions for the special functions (of the first and second kinds) used in the various NF decompositions. Here we give the precise mathematical definitions of the functions that we use, since there is often more than one convention in the literature.
For the little d function, which is also known as a reduced rotation matrix element (of the first kind) or Wigner function, we use the definition from ref (24), page 58, eq (4.1.23).
![]() |
B1 |
The definition of PJ–mf(mf,mf)(cos θR), a Jacobi polynomial, has become standardized (ref (24), page 57). Note that the Jacobi polynomial in eq (B1) is a special case of a Jacobi function of the first kind.(29) Important special values are
![]() |
where δmf 0 is a Kronecker delta function.
For the little e function, which is also called a reduced rotation matrix element of the second kind, we use10
![]() |
B2 |
where QJ–mf(mf, mf)(cos θR) is a Jacobi function of the second kind, which is a second independent solution of the Jacobi differential equation. We use Szegö’s definition as given in ref (29), page 78, eq (4.62.9). Note that QJ–mf(cos θR) is defined “on the cut”, −1 < cos θR < + 1.
Since mi = 0 in our applications, we also have the following important simplifications. From ref (24), page 59, eq (4.1.24), we get
![]() |
B3 |
and from section 2 in the Appendix of ref (10), we have
![]() |
B4 |
In eqs B3 and (B4), PJmf(cos θR) and QJ(cos θR) are Ferrers’ associated Legendre functions of the first and second kinds, respectively.30 They are defined by (ref (24), page 22, eq (2.5.10)).
![]() |
B5 |
The PJ(cos θR) in eq (B5) is a Legendre polynomial of degree J, which is a special case of a Legendre function of the first kind of degree J. The QJ(cos θR) in eq (B5) is a Legendre function of the second kind of degree J. Note: the other definition in common use, but not used in this paper, is that of Hobson, in which the right-hand-side of eq (B5) is multiplied by (−1)mf.
We also have the following simple results
![]() |
and
![]() |
The following asymptotic approximations are frequently useful.7,10 They are valid for a fixed value of mf, J → ∞, and 0 < θR< π.
![]() |
and
![]() |
The authors declare no competing financial interest.
Special Issue
Published as part of The Journal of Physical Chemistry virtual special issue “125 Years of The Journal of Physical Chemistry”.
References
- Yuan D.; Yu S.; Chen W.; Sang J.; Luo C.; Wang T.; Xu X.; Casavecchia P.; Wang X.; Sun Z.; Zhang D. H.; Yang X. Direct Observation of Forward-scattering Oscillations in the H + HD → H2 + D Reaction. Nat. Chem. 2018, 10, 653–658. 10.1038/s41557-018-0032-9. [DOI] [PubMed] [Google Scholar]
- Sang J.; Yuan D.; Chen W.; Yu S.; Luo C.; Wang S.; Wang T.; Yang X.; Wang X. High Resolution Crossed Molecular Beams Study of the H + HD → H2 + D Reaction. Chin. J. Chem. Phys. 2019, 32, 123–128. 10.1063/1674-0068/cjcp1901010. [DOI] [Google Scholar]
- Yuan D.; Guan Y.; Chen W.; Zhao H.; Yu S.; Luo C.; Tan Y.; Xie T.; Wang X.; Sun Z.; Zhang D. H.; Yang X. Observation of the Geometric Phase Effect in the H + HD → H2 + D Reaction. Science 2018, 362, 1289–1293. 10.1126/science.aav1356. [DOI] [PubMed] [Google Scholar]
- Xie Y.; Zhao H.; Wang Y.; Huang Y.; Wang T.; Xu X.; Xiao C.; Sun Z.; Zhang D. H.; Yang X. Quantum Interference in H + HD → H2 + D Between Direct Abstraction and Roaming Insertion Pathways. Science 2020, 368, 767–771. 10.1126/science.abb1564. [DOI] [PubMed] [Google Scholar]
- Yuan D.; Huang Y.; Chen W.; Zhao H.; Yu S.; Luo C.; Tan Y.; Wang S.; Wang X.; Sun Z.; Yang X. Observation of the Geometric Phase Effect in the H + HD → H2 + D Reaction Below the Conical Intersection. Nat. Commun. 2020, 11, 3640. 10.1038/s41467-020-17381-4. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Xiahou C.; Connor J. N. L. Glories, Hidden Rainbows and Nearside-farside Interference Effects in the Angular Scattering of the State-to-state H + HD → H2 + D Reaction. Phys. Chem. Chem. Phys. 2021, 23, 13349–13369. . This paper is denoted XC1 in the main text 10.1039/D1CP00942G. [DOI] [PubMed] [Google Scholar]
- Dobbyn A. J.; McCabe P.; Connor J. N. L.; Castillo J. F. Nearside-farside Analysis of State-selected Differential Cross Sections for Reactive Molecular Collisions. Phys. Chem. Chem. Phys. 1999, 1, 1115–1124. 10.1039/a809498e. [DOI] [Google Scholar]
- McCabe P.; Connor J. N. L. Nearside-farside Analysis of Differential Cross Sections: Diffraction and Rainbow Scattering in Atom-atom and Atom-molecule Rotationally Inelastic Sudden Collisions. J. Chem. Phys. 1996, 104, 2297–2311. 10.1063/1.470925. [DOI] [Google Scholar]
- McCabe P.; Connor J. N. L.; Sokolovski D. Nearside-farside Analysis of Differential Cross Sections: Ar + N2 Rotationally Inelastic Scattering Using Associated Legendre Functions of the First and Second Kinds. J. Chem. Phys. 1998, 108, 5695–5703. 10.1063/1.475979. [DOI] [Google Scholar]
- McCabe P.; Connor J. N. L.; Sokolovski D. Nearside-farside Analysis of Differential Cross Sections Using Jacobi Functions of the First and Second Kinds: Application to Ar + N2 Rotationally Inelastic Scattering. J. Chem. Phys. 2001, 114, 5194–5206. 10.1063/1.1335658. [DOI] [Google Scholar]
- Connor J. N. L.; McCabe P.; Sokolovski D.; Schatz G. C. Nearside-farside Analysis of Angular Scattering in Elastic, Inelastic and Reactive Molecular Collisions. Chem. Phys. Lett. 1993, 206, 119–122. 10.1016/0009-2614(93)85527-U. [DOI] [Google Scholar]
- Whiteley T. W. J.; Noli C.; Connor J. N. L. Nearside-farside Analysis of Differential Cross Sections: Ar + HF Rotationally Inelastic Scattering. J. Phys. Chem. A 2001, 105, 2792–2802. 10.1021/jp0100046. [DOI] [Google Scholar]
- Anni R.; Connor J. N. L.; Noli C. Improved Nearside-farside Method for Elastic Scattering Amplitudes. Phys. Rev. C: Nucl. Phys. 2002, 66, 044610. 10.1103/PhysRevC.66.044610. [DOI] [Google Scholar]
- Noli C.; Connor J. N. L.. Nearside-farside Analysis of Differential Cross Sections: Resummation of a Partial Wave Series Involving Legendre Polynomials. Russ. J. Phys. Chem. 2002, 76, S77–S89. Online at https://arXiv.org/abs/physics/0301054. [Google Scholar]
- Anni R.; Connor J. N. L.; Noli C.. Improved Nearside-farside Decomposition of Elastic Scattering Amplitudes. Khim. Fiz. 2004, 23, (2), , 6–12. Online at https://arXiv.org/abs/physics/0410266.
- Connor J. N. L.; Anni R. Local Angular Momentum–Local Impact Parameter Analysis: A New Tool for Understanding Structure in the Angular Distributions of Chemical Reactions. Phys. Chem. Chem. Phys. 2004, 6, 3364–3369. 10.1039/B402169J. [DOI] [Google Scholar]
- Hollifield J. J.; Connor J. N. L. Theory of Elastic Scattering in Strongly Absorptive Collisions: Application of the Fuller Nearside-farside Method. Phys. Rev. A: At. Mol. Opt. Phys. 1999, 59, 1694–1696. 10.1103/PhysRevA.59.1694. [DOI] [Google Scholar]
- Hollifield J. J.; Connor J. N. L. Nearside-farside Theory of Elastic Angular Scattering for Strongly Absorptive Collisions. Mol. Phys. 1999, 97, 293–304. 10.1080/00268979909482830. [DOI] [Google Scholar]
- Totenhofer A. J.; Noli C.; Connor J. N. L. Dynamics of the I + HI → IH + I Reaction: Application of Nearside-farside, Local Angular Momentum and Resummation Theories using the Fuller and Hatchell Decompositions. Phys. Chem. Chem. Phys. 2010, 12, 8772–8791. 10.1039/c003374j. [DOI] [PubMed] [Google Scholar]
- Herschbach D. R. Molecular Beam Studies of Internal Excitation of Reaction Products. Appl. Opt. 1965, 4 (S1), 128–144. 10.1364/AO.4.S1.000128. [DOI] [Google Scholar]
- Herschbach D. R. Reactive Scattering in Molecular Beams. Adv. Chem. Phys. 1966, 10, 319–393. 10.1002/9780470143568.ch9. [DOI] [Google Scholar]
- Jacob M.; Wick G. C. On the General Theory of Collisions for Particles with Spin. Ann. Phys. (N.Y.) 1959, 7, 404–428. 10.1016/0003-4916(59)90051-X. [DOI] [Google Scholar]
- Goldberger M. L.; Watson K. M.. Collision Theory ;Wiley: New York, 1967; Appendix E. The Jacob-Wick Expansion of the Scattering Matrix. [Google Scholar]
- Edmonds A. R.Angular Momentum in Quantum Mechanics ,2nd ed.; Princeton University Press: Princeton, NJ, 1974. (third printing with corrections). [Google Scholar]
- Monks P. D. D.; Xiahou C.; Connor J. N. L. Local Angular Momentum–Local Impact Parameter Analysis: Derivation and Properties of the Fundamental Identity, with Applications to the F + H2, H + D2, and Cl + HCl Chemical Reactions. J. Chem. Phys. 2006, 125, 133504. 10.1063/1.2210480. [DOI] [PubMed] [Google Scholar]
- Brussaard P. J.; Tolhoek H. A. Classical Limits of Clebsch-Gordan Coefficients, Racah Coefficients and Dmnl(φ,ϑ,ψ)– Functions. Physica 1957, 23, 955–971. 10.1016/S0031-8914(57)95547-7. [DOI] [Google Scholar]
- Boothroyd A. I.; Keogh W. J.; Martin P. G.; Peterson M. R. A Refined H3 Potential Energy Surface. J. Chem. Phys. 1996, 104, 7139–7152. 10.1063/1.471430. [DOI] [Google Scholar]
- Sokolovski D.; Connor J. N. L.; Schatz G. C. Complex Angular Momentum Analysis of Resonance Scattering in the Cl + HCl → ClH + Cl Reaction. J. Chem. Phys. 1995, 103, 5979–5998. 10.1063/1.470427. [DOI] [Google Scholar]
- Szegö G.Orthogonal Polynomials ,4th ed.; American Mathematical Society: Providence, RI, 1975. [Google Scholar]
- We follow ref 24, page 22, in calling eq (B5) Ferrers’ definition of Pjm (x) and Qjm (x), even though the Reverend N. M. Ferrers in his monograph An Elementary Treatise on Spherical Harmonics and Subjects Connected with Them ;MacMillan and Co.: London, England, 1877, only discussed in detail Pjm (x), which he denoted, TJ(m) (x). [Google Scholar]