Skip to main content

ORIGINAL RESEARCH article

Front. Chem., 18 October 2022
Sec. Inorganic Chemistry
Volume 10 - 2022 | https://doi.org/10.3389/fchem.2022.1005684

Triplet state harvesting and search for forbidden transition intensity in the nitrogen molecule

www.frontiersin.orgB. F Minaev1,2 www.frontiersin.orgO. O Panchenko1* www.frontiersin.orgV. A Minaeva1 www.frontiersin.orgH Ågren2
  • 1Department of chemistry and nanomaterial sciences, Bohdan Khmelnytsky National University, Cherkasy, Ukraine
  • 2Department of Physics and Astronomy, Uppsala University, Uppsala, Sweden

Triplet excited states of the N2 molecule play an important role in electric discharges through air or liquid nitrogen accompanied by various afterglows. In the rarefied upper atmosphere, they produce aurora borealis and participate in other energy-transfer processes connected with atmospheric photochemistry and nightglow. In this work, we present spin–orbit coupling calculations of the intensity of various forbidden transitions, including the prediction of the electric dipole transition moment of the new 13ΣgA3Σu+ band, which is strongly prohibited by the (+|−) selection rule, the new spin-induced magnetic B3ΣuA3Σu+ transition, magnetic and electric quadrupole transitions for the B3Πg X1Σg+ Wilkinson band, and the Lyman–Birge–Hopfield a1Πg ← X1Σg transition. Also, two other far-UV singlet–singlet quadrupole transitions are calculated for the first time, namely, the Dressler–Lutz a"1Σg+–X1Σg+ and the less studied z1Δg–X1Σg+ weak transitions.

Introduction

The great flux of solar energy through the upper atmosphere can be harvested by the rarefied gases of molecular and atomic components of the Earth’s mesosphere and lower thermosphere (MLT) regions (Minaev and Panchenko, 2020). The ground states of such abundant O2 (3Σg), O (3P), and N (4S) species of MLT possess high multiplicity, and thus their lowest excited states are metastable, having a low electronic spin and strongly forbidden radiative relaxation (Wilkinson and Mulliken, 1959; Brown and Winkle, 1970; Minaev and Panchenko, 2020). Their long-lived emission to the ground state provides the possibility to harvest visible and near-UV solar radiation engaged in various energy transfer processes, which determine the climate, meteorology, and weather conditions (Minaev and Panchenko, 2020). In contrast, the ground state of the nitrogen molecule possesses zero spin and several high-energy triplet excited states with deep potential wells. The lowest of them,A3Σu+, can harvest a stock of 6.22 eV energy, being a strongly metastable triplet state with a relatively long radiative lifetime (τr) of 2 s (Brown and Winkle, 1970; Minaev et al., 1995; Begley et al., 2022). Accounting for the short UV wavelength of the A3Σu+X1Σg+ transition, this τr value is indeed unusually large.

N2 is a very stable and inert molecule in the ground state XΣ1g+ with high dissociation energy (De = 9.76 eV). At the same time, N2 possesses a variety of quite stable valence excitations of the πu–πg and 3σg–πg types; these excited states have large De values (around 4–6 eV) and are mostly metastable since their emission to the ground state is strictly forbidden by the electric dipole selection rules (Wilkinson and Mulliken, 1959; Brown and Winkle, 1970; Begley et al., 2022; Minaev et al., 1995; Lewis et al., 2008; Lofthus. and Krupenie, 1977). In gaseous electric discharges, when a molecule is irradiated by an electron flux, N2 dissociates into the ground state N (4S) atoms; they can recombine forming the lowest singlet (X1Σg+), triplet (A3Σu+), and quintet (A′5Σg) basic states. The last two, shown in Figure 1, are involved in the so-called active nitrogen phenomenon detected by the characteristic “yellow afterglow” (Brown and Winkle, 1970; Begley et al., 2022). Its study together with aurora borealis involves a large number of metastable states and forbidden transitions in the N2 spectrum (Figure 2). The Lewis–Rayleigh afterglow (Brown and Winkle, 1970) in the discharge consists of the first positive system of the nitrogen molecule, extending from IR to the blue edge, being the triplet–triplet B3Πg → A3Σu+ transition (1+ system) (Lofthus. and Krupenie, 1977). The visible part of the 1+ system was already investigated in 1902 by Deslandres (1902); ab initio interpretation of its intensity was achieved by Werner et al. (1984) and a final form by Ni et al. (2017). It should be distinguished from the second positive system of the nitrogen molecule—the C3Πu → A3Σu+ transition (2+ system) and the infrared Hermann (HIR) band C″5Πu → A′5Σg+ (Figure 1). The main sources of emission of the first and second positive systems in N2 discharge are connected with the involvement of the N (2D) excited atom into a recombination reaction (Figure 1). The 2+ band system was observed as early as 1869 as it readily appears in ordinary air discharges (Deslandres, 1902), but its rovibronic assignment came much later (Lofthus and Krupenie, 1977). As opposed to the O2 molecule (Minaev and Panchenko, 2020), many visible and UV transitions between triplet excited states generated by electric discharge are possible in the nitrogen counterpart (Lofthus. and Krupenie, 1977; Minaev et al., 1995; Lewis et al., 2008). The quintet state A′5Σg+ and the HIR system of N2 have become clear only recently (Partridge et al., 1988; Hochlaf et al., 2010a). They are essentially important for the recombination of the N (4S) ground state atoms being the precursor of the Lewis–Rayleigh afterglow. The quintet A′5Σg+ can predissociate to the B3Πg state vibrational levels (v = 10–12, Figure 1), though the spin–orbit coupling (SOC) matrix element (ME) <A′5Σg+|Hso| B3Πg> is rather weak near the crossing in order to be efficient for generation of the spontaneous 1+ emission in the recombination of N (4S) atoms. At the same time, this SOC ME determines the high radiative probability (Einstein coefficient about 3∙104 s−1) of the newly predicted A′5Σg+ → A3Σu+ (0–6) transition, which borrows intensity from the 1+ system, as well as from the HIR band (Minaev et al., 2022). The latter source is attributed to a strong SOC between the A3Σu+ and C″5Πu states.

FIGURE 1
www.frontiersin.org

FIGURE 1. Potential energy curves of several spectroscopy important excited states of the N2 molecule. The first (1+) and second positive (2+) systems are denoted together with the Herman infrared (HIR) emission band.

FIGURE 2
www.frontiersin.org

FIGURE 2. Low-lying bound states, which contribute to the intensity of the studied transition.

The excited metastable N (2D) and N (2P) atoms with energies of 2.4 eV and 3.6 eV above the N (4S) ground state, respectively (Figure 1), are present with low concentration in the discharge. Their recombination leads to a huge number of excited N2 states with varying degrees of stability and spontaneous emission probabilities (Lofthus. and Krupenie, 1977). Several other important states of nitrogen are shown in Figure 2.

Energy harvesting by triplet states of nitrogen

The triplet excited manifold of the N2 molecule is well studied in far-UV absorption and emission spectra (Deslandres, 1902; Lofthus. and Krupenie, 1977; Werner et al., 1984; Partridge et al., 1988; Piper, 1993; Minaev et al., 1995; Lewis et al., 2008; Ndome et al., 2008; Hochlaf et al., 2010a; Ni et al., 2017). In 1932, Vegard detected 120 weak bands in the red-degraded phosphorescence of solid nitrogen through the wide region of 670–170 nm (Lofthus and Krupenie, 1977). Soon after, Kaplan observed similar bands in an N2 laboratory discharge (Minaev et al., 1995). The weak Vegard–Kaplan (VK) system was first detected by Wilkinson as absorption bands in a long-path spectrometer at 169 and 128 nm for highly excited vibronic levels (v’ = 6,7) (Wilkinson and Mulliken, 1959). Later on, the VK rovibronic intensity alternations were measured and analyzed very carefully (Lofthus and Krupenie, 1977; Piper, 1993) including ab initio calculations for the VK transition probability and many other inter-combination systems (Minaev et al., 1995). SOC calculations within the quadratic response theory (Minaev et al., 1995) explained why the Ogawa–Tanaka–Wilkinson system BΣ3uXΣ1g+ is much more intense (70 times) than the Vegard–Kaplan AΣ3u+XΣ1g+absorption and why the Tanaka transition CΠ3uXΣ1g+ is the most intense among all known triplet–singlet (T–S) absorption bands at that time (Lofthus and Krupenie, 1977; Minaev et al., 1995). The new T ← S transition DΣ3u+XΣ1g+ in the far-UV region predicted by Minaev et al. (1995) was later detected and analyzed by Lewis et al. (2008). The upper DΣ3u+ state has been observed earlier in the pure nitrogen condensed discharge afterglow through the DΣ3u+→B3Πg (0, v’‘) emission, which is now known as the fourth positive system (Lofthus and Krupenie, 1977; Minaev et al., 1995). The upper DΣ3u+ state was shown to be of Rydberg type (Minaev et al., 1995) converging to the ground state N2+ ion. At longer N–N distances, it avoids crossing with the bound Rydberg state and the valence 33Σu+ state potential energy curve (PEC), demonstrating a repulsive character (Minaev et al., 1995). All theoretical predictions of the inter-combination D ← X transition (Minaev et al., 1995) have mainly been supported by later experiments (Lewis et al., 2008). The predicted 0–0 transition is rather intense (f = 2×10−5), being the strongest inter-combination of a nitrogen molecule in agreement with measurements (Lofthus and Krupenie, 1977; Lewis et al., 2008). This far-UV region in N2 absorption is very dense, being covered by allowed transitions (b1Πu–X, for example) (Lofthus and Krupenie, 1977), but the D–X (0, 0) band has a clear location in a fortuitous region of the b1Πu ← X allowed spectrum, just above its (4,0) band head, enabling the D–X (0, 0) observation (Lewis et al., 2008). All three sublevels of the triplet D state provide four rotational branches in agreement with Minaev et al. (1995), according to rotational and parity selection rules of Hund’s case “b” (Lewis et al., 2008). The small negative zero-field splitting (λ = −0.036 cm−1 (Lewis et al., 2008)) of the DΣ3u+ state is in agreement with SOC and spin–spin coupling calculations (λ = −0.041 cm−1) within the response approach (Loboda et al., 2003; Hochlaf et al., 2010b; Qin et al., 2019; Minaev et al., 2022).

Thus, almost all important singlet–triplet transitions in the molecular nitrogen absorption spectra (up to the far-UV region) from the ground state XΣ1g+ to the triplet states of the “ungerade” symmetry—the AΣ3u+,BΣ3u, W3Δu, C3Πu, and DΣ3u+ states—have been calculated by the quadratic response theory within the multi-configuration approach (Minaev et al., 1995), giving results that are in good agreement with experimental intensity distributions (Lofthus. and Krupenie, 1977; Piper, 1993; Lewis et al., 2008). The present work aims to calculate new forbidden transitions in the nitrogen spectra which have not been observed so far but can influence the triplet state harvesting and total kinetic balance of the upper atmosphere.

The B3Πg state produced by the second and fourth positive systems (Lofthus and Krupenie, 1977) can further generate 1+ bands, and the lowest triplet AΣ3u+ state by the cascade in the positive column of electric discharge. We have to note that the B3ΠgXΣ1g+ phosphorescence was not calculated in Minaev et al. (1995), since even an account of SOC cannot overcome its parity prohibition in terms of electric dipole selection rules. The calculation of this transition intensity is an aim of the present work.

The VK transition satisfies the orbital electric dipole selection rule (EDSR) (Minaev et al., 1995), but being spin-forbidden it cannot be effectively induced by direct UV absorption. Thus, the N2 (A) state is primarily populated by collisions—in laboratory discharge and the upper atmosphere, this is accomplished through the electron impact and the cascade in the first positive system. The relatively long radiative lifetime enables N2 (AΣ3u+) to participate in collisions with the main background gases of the MLT region and to produce chemical reactions with N2, O2, N, and O species. In particular, the reactions

N2(A3Σu+)+O(3P)=NO(X2Π)+N2(D)
N2(A3Σu+)+O(3P)=N2(X1Σg+)+O(1S)

are the most important ones (Yonker and Bailey, 2019). A recent steady-state MLT model developed for the N2 (AΣ3u+) vibrational distribution in the terrestrial atmosphere is supported by comparison with the Vegard–Kaplan dayglow emission from atmospheric photochemistry and ionospheric spectroscopy measurements (Yonker and Bailey, 2019). The steady-state N2 (AΣ3u+,v) vibrational distribution in the MLT region is found to be shifted to higher (v > 6) levels. This is in agreement with the VK absorption (Minaev et al., 1995) and is important for our study. Direct excitation from the ground N2(X) state by the electron impact provides an essential contribution to populating the N2 (A, v > 6) sublevels, though their dominant excitation mechanism is the radiative cascade via the 1+ system (Bruna and Grein, 2009; Yonker and Bailey, 2019; Ajello et al., 2020). The efficiency of this cascade depends on the B3ΠgX1Σg+ transition intensity, which in turn is determined by the EDSR-forbidden a1ΠgX1Σg+ magnetic-dipole-allowed band system. Intensity calculations of these strongly forbidden transitions are also the purpose of our work.

The N2 molecule, the most common and abundant component of the air, plays a crucial role in many high-energy photochemical processes caused by solar radiation in the upper atmosphere (Yonker and Bailey, 2019; Ajello et al., 2020). The discovery of new N2 transitions forbidden by the spin-selection rule and induced by SOC perturbation is an important part of optical nitrogen monitoring at different altitudes. The intensity origin of the known emission bands that are forbidden by the electric dipole selection rules is also an important task of N2 spectroscopy (Deslandres, 1902; Wilkinson and Mulliken, 1959; Brown and Winkle, 1970; Lofthus and Krupenie, 1977; Werner et al., 1984; Partridge et al., 1988; Piper, 1993; Minaev et al., 1995; Lewis et al., 2008; Ndome et al., 2008; Hochlaf et al., 2010a; Ni et al., 2017; Begley et al., 2022). This work presents multi-reference configuration interaction (MRCI) calculations of the highly excited states of the nitrogen molecule and an explanation of the intensity origin of several forbidden optical transitions. With this aim and background, we have predicted the electric dipole transition moment (EDTM) of the unknown forbidden transition 13ΣgA3Σu+ and calculated its dependence on the internuclear distance. This is a triplet–triplet (T–T) band, the intensity of which is entirely determined by spin–orbit coupling perturbations between various spin sublevels of the T states as was preliminarily shown in a recent work (Minaev et al., 2022). The upper 13Σg state was earlier calculated by similar MRCI methods (Hochlaf et al., 2010b; Qin et al., 2019), but no experimental manifestations of its existence have been evidenced so far, although the 13Σgstate is predicted with a deep minimum (De = 1.23 eV) and high energy above the ground state (Te = 12.15 eV) (Qin et al., 2019). We believe that the 13Σg state can be produced by N (2P) + N (4S) recombination (Figure 1), and that its low vibrational levels can avoid pre-dissociation at low pressure. The N2(13Σg) state possesses a potential energy well located outside the Franck–Condon (FC) region, which is accessible from the metastable A3Σu+ state as well as from the ground state. This explains the difficulties with the observation of the corresponding absorption bands. Under these conditions, the emissive 13ΣgA3Σu+ transition from the lowest v’ = 0 sublevel will have the maximum FC factor for the v” = 7–8 vibronic levels of the A state. We provide evidence for the existence of this new band in the N2 molecule by calculating the transition probabilities through an account of SOC in the first order of the perturbation theory and comparing them with other known forbidden transitions to facilitate the validity of such a prediction. This would be a wide band of low intensity in the range of 209–450 nm with an approximate maximum at 328 nm; it is prohibited by the severe selection rule (+) → (–) but is allowed by spin-selection as a T–T transition (Minaev et al., 2022). Its spin-rovibronic structure would be analogous to the well-known Herzberg I band of molecular oxygen X3ΣgA3Σu+ (Herzberg, 1952; Minaev and Muldakhmetov, 1984; Klotz and Peyerimhoff, 1986).

Intensity borrowing mechanisms of the forbidden 13ΣgA3Σu+ transition

For planning intensity calculations of the new band in nitrogen, we first take into account the corresponding well-known and intense transitions of the N2 molecule, relevant for our purpose. According to SOC selection rules, the new N2 band 13ΣgA3Σu+ can be formed by spin–orbit coupling-induced mixing of the upper state 13Σg with the B3Πg state and by intensity borrowing from the first positive system A3Σu+B3Πg, see Eq. 1. To include the SOC effect, we have to add the Ω quantization, where Ω = Lz + Sz is the z-projection of the total electronic angular momentum and Lz and Sz are orbital and spin angular momenta projections on the molecular axis (Yonker and Bailey, 2019). The SOC operator can mix states with the same Ω; the EDTM selection allows transitions according to the rule ΔΩ = 0, ±1:

|1Σg,13=|1Σg,13+BΠg,13|HSO|1Σg,13E(1Σg3)E(BΠg3)|BΠg,13=|1Σg,13+СB,1|BΠg,13,
AΣu,0+3|er|1Σg,13=CB,1AΣu,0+3|e×|BΠg,13,(1)

Figure 3 presents this mechanism as the type “I SOC” mixing. By a similar SOC mechanism, the studied forbidden band can borrow EDTM intensity from the newly predicted 13ΣgC3Πu transition here, Eq. 2; this SOC mechanism of intensity borrowing refers to the “II SOC” type in Figure 3. Both these mechanisms have the perpendicular x, y polarization of EDTM; only the x component is shown in Eqs. 12 for one sublevel of the degenerate Π states (Qin et al., 2019; Minaev et al., 2022).

|AΣu,0+3=|AΣu,0+3+CΠu,03|HSO|AΣu,0+3E(AΣu+3)E(CΠu3)|CΠu,03=|AΣu,0+3+CC,A|CΠu,03,
AΣu,0+3|er|1Σg,13=CC,ACΠu,03|e×|1Σg,13.(2)

FIGURE 3
www.frontiersin.org

FIGURE 3. Scheme of intensity borrowing for the forbidden ΣgΣu+ transition. The first and second mechanisms provide perpendicular polarization, and the third one is responsible for the parallel polarization along the N–N axis.

Figure 3 provides a good explanation of the relevant intensity sources of the studied 13ΣgA3Σu+ transition, but it would be overloaded if all possible contributions are included. The type “II SOC” mechanism in Figure 3 includes also other states of the C′3Πu type (in total five 3Πu states are taken into account).

An additional source of intensity borrowing denoted as the type “III SOC” mechanism in Figure 3 includes parallel EDTM for the studied emission band (light polarization along the molecular z-axis). By symmetry arguments, the 13ΣgA3Σu+ transition in N2 is similar to the Herzberg I band of the O2 molecule, and its probability can be calculated by a similar scheme of intensity borrowing (Minaev and Muldakhmetov, 1984; Klotz and Peyerimhoff, 1986). In the oxygen molecule, the main contribution to the absorption intensity of the Herzberg I band X3ΣgA3Σu+ origins from the SOC mixing between the A3Σu,1+ state and the upper term B3Σu,1of the Schumann–Runge system X3ΣgB3Σu (Minaev and Muldakhmetov, 1984), which is the most intense valence transition in molecular oxygen (Herzberg, 1952). This provides a rather unusual (for the 3Σ3Σ+ band) type of Ω = 1–Ω = 1 parallel transition intensity, though the ΔΩ = 1 selection rule is more typical for such bands with prevailing perpendicular polarization (Herzberg, 1952).

Let us consider the type “III SOC” mechanism in more detail. The SOC-induced mixing between the lowest A3Σu+ state and the upper triplet BΣ3u of the Ogawa–Tanaka–Wilkinson system (B3ΣuX1Σg+) can be presented by the perturbation theory in the form:

|AΣu,1+3=|AΣu,1+3+BΣu,13|HSO|AΣu,1+3E(AΣu+3)E(BΣu3)|BΣu,13=|AΣu,1+3+CB,A|BΣu,13.(3)

We can also account for SOC perturbation for the 13Σg counterpart as follows:

|1Σg,13=|1Σg,13+EΣg,1+3|HSO|1Σg,13E(1Σg3)E(EΣg+3)|EΣg,1+3=|1Σg,13+CE,1|EΣg,1+3.(4)

The EDTM between the perturbed states (3) and (4) is equal to

1Σg,13|er|AΣu,1+3=CE,1EΣg,1+3|ez|AΣu,1+3+CB,A1Σg,13|ez|BΣu,1+3.(5)

This means that the (+|−) forbidden transition 13ΣgA3Σu+ can borrow intensity from ED-allowed E3Σg+A3Σu+ and 13ΣgB3Σu transitions. The latter contribution is a formal symmetry analog of the Schumann–Runge O2 transition. The SOC mixing mechanism shown in Eq. 3 is presented in Figure 3 by the intensity borrowing scheme “III-SOC”. The SOC-induced mechanism from Eq. 4 is not shown in Figure 3 to avoid overloading. The SOC matrix element (ME) in Eq. 3 is equal to zero in a semi-empirical approximation with the neglect of differential overlap:

HSOC=AςAilA,isi=iBisi,(6)

where ςA is the SOC constant for the valence shell of the A atom and lA,isi is a scalar product of the orbital and spin operators for the ith electron (Minaev and Muldakhmetov, 1984; Minaev et al., 1993). For the pure main configurations of the A3Σu+ and B3Σu states, the SOC ME is equal to 12(Bπu,xπu,yBπg,xπg,y); this expression is zero with the neglect of differential overlap since Bπu,xπu,y=Bπg,xπg,y (Minaev and Muldakhmetov, 1984), but the account of overlap in normalization of the πu and πg molecular orbitals in the r-centroid approach (1.282 Å) leads to the different estimations Bπu,xπu,y = 85 cm−1and Bπg,xπg,y = 60 cm−1. Thus, the SOC ME in Eq. 3 reaches a non-zero value of 12.5 cm−1, which is rather close to the MRCI result. This scrutinized analysis shows the importance of the contribution expressed by Eq. 3 and the analogy with the Herzberg I Schumann–Runge transition coupling in the O2 molecule (Minaev and Muldakhmetov, 1984). The denominator in Eq. 3 is rather small and homogeneously changes with r distance (Figure 3). Although the 13ΣgB3Σu transition is relatively weak in the N2 molecule (EDTM = 0.026 ea0 at r = 1.4 Å) (Qin et al., 2019), its contribution to the final EDTM of Eqs 15 is the largest. The EDTM of the E3Σg+A3Σu+ transition (the Herman–Kaplan band system (Lofthus. and Krupenie, 1977)) has a smaller value (0.017 and 0.0105 ea0 at r = 1.28 and 1.4 Å, respectively) (Qin et al., 2019), as well as the SOC ME in Eq. 4 at these distances (5.2 cm−1) (Hochlaf et al., 2010b).

We have stressed before the EDTM component of the studied intensity borrowing from the 13ΣgB3Σu band, which is a formal analog of the Schumann–Runge system of oxygen (Minaev and Panchenko, 2020). Thus, we can compare various contributions to the intensity of this so-far unknown transition with the well-known data for O2 and N2 spectra (Herzberg, 1952; Lofthus. and Krupenie, 1977; Minaev et al., 1995; Lewis et al., 2008). The intensity borrowing contribution from the first positive system B3ΠgA3Σu+ in Eq. 1 can be compared with the Vegard–Kaplan S–T transition intensity presented in Eq. 7 (Minaev et al., 1995), which explains an extremely low spontaneous emission of the VK system.

<XΣ1g+|er|AΣ3u+=CX,B<BΠ3g|er|AΣ3u+>+Cb,A<bΠ1u|er|X1Σg+>,(7)
CX,B=<XΣ1g+|Hso|BΠ3g>E(BΠ3g)E(XΣ1g+),Cb,A=<bΠ1u|Hso|AΣ3u+>E(AΣ3u+)E(bΠ1u).

As shown in Figure 2, the two denominators in Eq. 7 have opposite signs. The first denominator E(B)–E(X) decreases with r distance prolongation, whereas the second one, E(A)–E(b), increases by an absolute value with r. In the vicinity of the ground state equilibrium re distance (1.098 Å), both contributions tend to cancel each other, and the EDTM value crosses the zero point (Minaev et al., 1995). In the whole FC region, the EDTM is still close to zero, and the VK system has very low intensity both in absorption and emission. Although both the SOC ME values in the nominators of Eq. 7 are rather large (Bruna and Grein, 2009; Hochlaf et al., 2010b) as well as the transition moments of the 1+ and bΠ1uX1Σg+ systems (Qin et al., 2019), the cancellation of the two big terms in Eq. 7 is the only reason for the relatively large lifetime of the N2(A3Σu+) state. To a large extent, this is also the reason for the efficient solar energy harvesting by the triplet states of nitrogen molecules and the aurora borealis phenomena.

For the studied transition 13ΣgA3Σu+, only the first “I SOC” mechanism provides an essential sign change with the internuclear distance (Figure 4). In the FC region, no big cancellations of different sign contributions are shown. The deteriorating “I SOC” mechanism is rather weak in the FC region 1.28–1.62 Å. For the most intense 0–7 vibronic band, the calculated EDTM is equal to 1.41×10−4 ea0, which corresponds to the radiative rate constant of 2.48 s−1. The total radiative lifetime of the zero vibrational sublevel of the 13Σg state is estimated as 0.34 s.

FIGURE 4
www.frontiersin.org

FIGURE 4. Transition dipole moment contributions of the 13ΣgA3Σu+ band system of the N2 molecule according to three mechanisms shown schematically in Figure 3.

The 13Σg state can degrade much faster in the allowed T–T transitions (for example, through the 13ΣgB3Σu emission). Thus, our estimation of the emissive 13ΣgA3Σu+ transition is definitely negative. However, in absorption, the same A3Σu+(v=7)13Σg(v=0) transition can be observable since the calculated oscillator strength (f7–0 = 2.23*10−9) can be measured by modern techniques.

It is, at this point, relevant to estimate the other EDSR-forbidden inter-combination B3ΠgX1Σg+ transition of nitrogen (Wilkinson system) (Lofthus. and Krupenie, 1977), which so far has not been calculated by quantum chemical methods. This is a magnetic dipole transition that borrows intensity from the magnetic singlet–singlet counterpart a1ΠgX1Σg+ (Lofthus. and Krupenie, 1977).

Calculations of magnetic and electric quadrupole transition intensity

The Lyman–Birge–Hopfield (LBH) band system (a1ΠgX1Σg+) of the N2 molecule has been carefully studied in measurements of cascade-induced UV radiation to determine the intensity of this emission (Lofthus. and Krupenie, 1977). The LBH band has readily been seen in absorption as well as in emission though it is EDSR-forbidden by parity selection. Its magnetic and quadrupole transition moments are provided in Figure 5. They are calculated here at the level of the time-dependent density functional theory (TD DFT) using the B3LYP functional and 6-311G++(d, p) basis set with the Gaussian-09 package (Frisch et al., 2010). We have studied 40 singlet states and triplet excited states of N2 in the region 0.8–1.8 Å of the r distances. For the longer N–N bonds, the TD DFT approach produces untrustworthy PECs and cannot reproduce the proper dissociation limits. But for short r distances, all potential energy curves are quite reasonable and qualitatively reproduce MRCI results (Dahl and Oddershede, 1986; Qin et al., 2019). This DFT method provides equilibrium bond lengths of 1.205 and 1.598 Å for the triplet (B3Πg) and quintet (A′5Σg+) states of nitrogen, respectively. The latter is more realistic (Hochlaf et al., 2010a), whereas the former re value deviates slightly from the experimental value of 1.213 Å (Lofthus. and Krupenie, 1977).

FIGURE 5
www.frontiersin.org

FIGURE 5. (A) Square of magnetic dipole moment mx2+my2B is the Bohr magneton) and (B) electric quadrupole moment of the a1Πg–X1Σg+ transition in the N2 molecule (both in a. u.).

A similar approach has been successfully used for the permanent quadrupole moment calculations in N2 (Dahl and Oddershede, 1986). In addition to the LBH system, some other EDSR-forbidden bands are also calculated as quadrupole transitions, as shown in Figure 6. The Dressler–Lutz a"1Σg+–X1Σg+ quadrupole transition at 101 nm as well as the far-UV transition z1ΔgX1Σg+ (Figure 6) are calculated for the first time.

FIGURE 6
www.frontiersin.org

FIGURE 6. Electric quadrupole moment of the a”1Σg+X1Σg+ transition in the N2 molecule for all allowed components of the quadrupole tensor operator. The far-UV transition 1Δg– X1Σg+ for the QXY = QYX quadrupole tensor components are also presented (both in a. u.).

The growth of magnetic strength of the a1ΠgX1Σg+ transition (Figure 5A) and the decrease of its quadrupole moment are notable (Figure 5B). The aΣ1g+ X1Σg+ transition moment represents a complicated tensor with r-dependent anisotropy (Figure 6).

In the FC region (1.1–1.3 Å), our results in Figure 5 well coincide with the calculations of Dahl and Oddershede, (1986) using the random phase approximation (RPA). The magnetic dipole transition moment (MDTM) of the LBH system (Figure 5A) increases with r, showing a trend of saturation at r = 1.3 Å, whereas the electric quadrupole transition moment (EQTM) decreases along the whole r range. Accounting for experimental FC factors and transition frequencies, we have obtained the radiative lifetime for the 0–0 vibronic transition of the LBH system equal to 65 μs in a reasonable agreement with experimental values in the interval 80–120 μs (Lofthus. and Krupenie, 1977; Dahl and Oddershede, 1986). The calculated magnetic to quadrupole intensity ratio (m/eq) is equal to 92%, whereas experimental data are in the range of 67%–96% interval (Dahl and Oddershede, 1986). Emission from the higher vibrational levels has a lower probability of qualitative agreement with observations (Lofthus. and Krupenie, 1977; Dahl and Oddershede, 1986). At the same time, we cannot accept the idea that the a1Πg state can decay solely into the X1Σg+ ground state (Dahl and Oddershede, 1986). From Figure 2, one can see that the infrared a1Πga’1Σu emission is possible; its electric dipole transition moment is equal to 0.2 ea0 (Qin et al., 2019) using the r-centroid approach corresponding to the radiative lifetime for the 0–0 band of τr = 9 ms (FC factor is 0.219). We have also estimated a new quadrupole transition a1ΠgB3Πg,1. Accounting for SOC, in Eq. 8, this transition moment origins in the difference in the permanent quadrupole moments of these two states: Q (B3Πg) = 0.59 ea02 and Q (a1Πg) = 0.48 ea02. This difference is small as well as the quadrupole moment of transition a→B (4.9*10–4 ea02), but in principle, we could not disregard branching emission into other lower lying triplet states (B′, W, and A) in the calculation of the radiative lifetime of the LBH system. These S–T transitions are allowed in the EDSR approach with an account of spin–orbit coupling perturbation. Thus, we consider it more appropriate to present also the oscillator strength for the Lyman–Birge–Hopfield 0–0 band a1ΠgX1Σg+ in absorption: f0–0 = 7.24×10−6.

The Dressler–Lutz a"1Σg+X1Σg+ quadrupole transition in the far-UV absorption region (101 nm) is of the Rydberg type (Lofthus. and Krupenie, 1977); it is well reproduced by our TD DFT calculations. The triplet counterpart of the a"1Σg+ state is the known E3Σg+ Rydberg term, which was discussed previously when presenting our calculations of the Herman–Kaplan system (E3Σg+A3Σu+ transition). The Dressler–Lutz a"1Σg+X1Σg+ band was observed in absorption at high pressure, and its intensity is mainly induced by collisions (Lofthus. and Krupenie, 1977). In this aspect, it is similar to the quadrupole Noxon band of O2, which is very sensitive to collision-induced intensity enhancement (Minaev et al., 1994). Both 1Σg+ states have similar re distance (about 1.1 Å) and FC factor close to unit. The calculated oscillator strength of the 0–0 band of the quadrupole a"1Σg+X1Σg+ transition in nitrogen is equal to 1.5·10−7, and it is detectable even at low pressure.

Now, we can estimate the probability of the latter triplet–singlet B3ΠgX1Σg+ transition of the nitrogen molecule which, being strictly forbidden by ED selection, has not been included in previous calculations (Minaev et al., 1995). This Wilkinson band borrows intensity from the LBH band system (a1ΠgX1Σg+) of the N2 molecule because of the relatively strong spin–orbit coupling

<B3Πg,1|Hsox,y| a1Πg>=41.4 cm1,(8)

at the re distance and small energy gap between the B–a states. Only the Ω = 1 spin sublevel of the triplet B3Πg,1 state is active in the Wilkinson band absorption, and its rotational structure supports the magnetic transition nature (Lofthus. and Krupenie, 1977). The SOC of Eq. 8 and m1 magnetic moment (Figure 5A) provide the largest contribution (98.6%) to the B3ΠgX1Σg+ transition intensity. The other k1Πg state (1πu→3σu) shows a smaller magnetic moment for the k1ΠgX transition (m = 0.085 μB) and a much smaller SOC counterpart at the B state equilibrium. Although both parameters increase with r, their relative contributions remain rather small. The calculated magnetic transition moment for the 0–0 band of the Wilkinson absorption B3ΠgX1Σg+ is equal to 0.0073 μB. It corresponds to the oscillator strength f0–0 = 2.54∙10–10, and the magnetic intensity remains dominant for this transition. It is not strange that Wilkinson (1962) used an optical path as long as 20 m to detect this band.

Finally, we have estimated the spin-induced magnetic dipole moment for a new B3ΣuA3Σu+ transition of the N2 molecule. According to Eq. 3, the perturbed A state has a small B’ state admixture for the Ms = ±1 sublevels: |AΣu,1+3+CB,A|BΣu,13. Thus, the transition to the next Ms = 0+ spin sublevel of the B state B3Σu,0+A3Σu,1+ can borrow spin-current intensity from the microwave B3Σu,0B3Σu,1 absorption band with the standard spin-magnetic transition moment that equals 2 μB. For the 0–0 absorption band B3ΣuA3Σu+, we have obtained oscillator strength f = 1.67 . 10–12, which is probably possible for detection.

Conclusion

The presence of nitrogen atoms in the discharge afterglow classifies “active nitrogen” as a free-radical phenomenon. This is relevant to the aurora borealis’ bright light and the yellow–orange Lewis–Rayleigh afterglow in the N2 gas discharge. The spectrum consists of several triplet–triplet emission bands of the 1+ and 2 + nitrogen systems (B3Πg–A3Σu+ and C3Πu–B3Πg transitions) and the B3ΣuB3Πg infrared-visible afterglow system. The wide Wu–Benesh system B3Πg = W3Δu is another T–T transition of the afterglow (Lofthus. and Krupenie, 1977). One can see that many triplet states of the nitrogen molecule take part in discharge afterglow together with numerous T–S transitions and S–S cascades. The transitions allowed by the electric dipole selection rule are nowadays accurately calculated by sophisticated ab initio methods (Qin et al., 2019) including many T–S vibronic bands induced by SOC perturbation (Minaev et al., 1995). This is important for the kinetic balance of triplet harvesting in discharges and the Earth’s mesosphere and lower thermosphere regions. In the present work, we have calculated the probability of the magnetic and quadrupole Lyman–Berge–Hopfield transition a1ΠgX1Σg+, which is necessary for the intensity estimation of the Wilkinson B3ΠgX1Σg+ band (the only unknown intensity of a pure electronic T–S transition at zero pressure).

We have also calculated new transitions, 13ΣgA3Σu+ and B3ΣuA3Σu+, that can be observed during absorption. The reason for finding such transitions is that the first excited triplet state A3Σu+ of N2 possesses a relatively long radiative lifetime (about 2 s). Therefore, it is possible to excite the triplet–triplet transition from the A3Σu+ state by two-photon experiments or other methods of flash photolysis in discharge. We know that the Herzberg I transition was discovered in the oxygen molecule as an excitation from the ground state X3Σg, but in nitrogen, the situation is reversed since the 13Σg symmetry corresponds to the upper state.

The 13Σg state, non-observed so far, has an electronic wave function, which is mainly represented by the valence configuration (1πu)2 (1πg)2 in a form similar to a quintet A5Σg+ state. The quintet–triplet A5Σg+A3Σu+ transition, also induced by SOC in the electric dipole approach, is the most intense among all studied intercombinations. The spin-induced B3ΣuA3Σu+ transition in the visible region is interesting since it is rather unique in magnetic-origin borrowing intensity from the electron spin resonance in the B3Σu state. The transition intensity could be sensitive to the external magnetic field in solid nitrogen. The B3ΣuA3Σu+ band in N2 has common features with the visible A-band of molecular oxygen (Minaev et al., 1994; Minaev and Minaeva, 2001).

Thus, we have noted many important comparable features in N2 and O2 spectra and also calculated for the first time the intensity of the predicted forbidden transitions including some magnetic dipole and quadruple S–S transitions in the nitrogen molecule. The main new predicted results are summarized in the following table.

www.frontiersin.org

Data availability statement

The raw data supporting the conclusion of this article will be made available by the authors, without undue reservation.

Author contributions

OP: writing small fragments of the text, computer calculations of molecules, development of drawings, and correction of the text. BM: main author of the manuscript, writing most of the text, development of drawings, and selection and processing of literary sources. VM: writing text fragments and text correction. HÅ: writing text fragments and processing computer calculations.

Funding

This work was supported by the Ministry of Science and Education of Ukraine (project 0122U000760) and by the Swedish Wenner-Gren Foundations (project GFU 2022–0036).

Acknowledgments

The authors express gratitude to Ramon S. da Silva and Majdi Hochlaf for useful discussions. Boris Minaev acknowledges a grant from the Wennergren-Foundations through their program for support of international reserach, grant no. GFU2022-0036. The authors thank the Swedish National Infrastructure for Computing (SNIC 2021-3-22 and SNIC 2022-5-103) at the National Supercomputer Centre of Linköping University and High-Performance Computing Center North (Sweden) partially funded by the Swedish Research Council through grant agreement no. 2018-05973.

Conflict of interest

The authors declare that the research was conducted in the absence of any commercial or financial relationships that could be construed as a potential conflict of interest.

Publisher’s note

All claims expressed in this article are solely those of the authors and do not necessarily represent those of their affiliated organizations, or those of the publisher, the editors, and the reviewers. Any product that may be evaluated in this article, or claim that may be made by its manufacturer, is not guaranteed or endorsed by the publisher.

References

Ajello, J. M., Evans, J. S., Veibell, V., Malone, C. P., Holsclaw, G. M., Hoskins, A. C., et al. (2020). The UV spectrum of the Lyman‐Birge‐Hopfield band system of N2is induced by cascading from electron impact. J. Geophys. Res. Space Phys. 125, e2019JA027546. doi:10.1029/2019JA027546

CrossRef Full Text | Google Scholar

Begley, A. I., Shuman, N. S., Long, B. A., Kampf, R., Gyr, L., Viggiano, A. A., et al. (2022). Excited-state N atoms transform aromatic hydrocarbons into N-heterocycles in low-temperature plasmas. J. Phys. Chem. A 126 (10), 1743–1754. doi:10.1021/acs.jpca.1c10657

PubMed Abstract | CrossRef Full Text | Google Scholar

Brown, R., and Winkle, C. A. (1970). The chemical behavior of active nitrogen. Angew. Chem. Int. Ed. Engl. 9 (3), 181–196. doi:10.1002/anie.197001811

CrossRef Full Text | Google Scholar

Bruna, P. J., and Grein, F. (2009). Theoretical study of electric moments, polarizabilities, and fine and hyperfine coupling constants of the B3Πg, C3Πu, A’5Σg+, and C’5Πustates of N2//Can. Can. J. Phys. 87, 589–600. doi:10.1139/P09-011

CrossRef Full Text | Google Scholar

Dahl, F., and Oddershede, J. (1986). Radiative lifetime of the ”forbidden” a1Πg← X1Σg+ transition of N2. Phys. Scr. 33, 135–140. doi:10.1088/0031-8949/33/2/007

CrossRef Full Text | Google Scholar

Deslandres, H. (1902). On the band spectra of nitrogen//C. R. Acad. Sci. 134, 747–750.

Google Scholar

Frisch, M. J., Trucks, G. W., Schlegel, H. B., Scuseri, G. E., Robb, M. A., Cheeseman, J. R., et al. (2010). Gaussian, inc., Gaussian 09, revision C.01. Wallingford CT.

Google Scholar

Herzberg, H. (1952). Spectra of diatomic molecules. Van Nostrand.

Google Scholar

Hochlaf, M., Ndome, H., and Hammoutène, D. (2010). Quintet electronic states of N2. J. Chem. Phys. 132, 104310. doi:10.1063/1.3359000

PubMed Abstract | CrossRef Full Text | Google Scholar

Hochlaf, M., Ndome, H., Hammoutene, D., and Vervloet, M. (2010). Valence Rydberg electronic states of N2: Spectroscopy and spin-orbit couplings. J. Phys. B At. Mol. Opt. Phys. 43, 245101. doi:10.1088/0953-4075/43/24/245101

CrossRef Full Text | Google Scholar

Klotz, R., and Peyerimhoff, S. D. (1986). Theoretical study of the intensity of the spin- or dipole forbidden transitions between the c 1Σ-u, A′ 3Δ u, A 3Σ+u and X 3Σ-g, a1Δg, b1Σg states in O2. Mol. Phys. 57, 573–594. doi:10.1080/00268978600100421

CrossRef Full Text | Google Scholar

Lewis, B. R., Baldwin, G. H., Heays, A. N., Gibson, S. T., Sprengers, J. P., Ubachs, W., et al. (2008). Structure and predissociation of the 3pσuD3Σu+ Rydberg state of N2: First extreme-ultraviolet and new near-infrared observations, with coupled-channels analysis. J. Chem. Phys. 129, 204303. doi:10.1063/1.3023034

PubMed Abstract | CrossRef Full Text | Google Scholar

Loboda, O., Minaev, B., Vahtras, O., Schimmelpfennig, B., Ågren, H., Ruud, K., et al. (2003). Ab initio calculations of zero-field splitting parameters in linear polyacenes. Chem. Phys. 286 (1), 127–137. doi:10.1016/S0301-0104(02)00914-X

CrossRef Full Text | Google Scholar

Lofthus., A., and Krupenie, P. H. (1977). The spectrum of molecular nitrogen. J. Phys. Chem. Ref. Data 6, 113–307. doi:10.1063/1.555546

CrossRef Full Text | Google Scholar

Minaev, B. F., Knuts, S., Ågren, H., and Vahtras, O. (1993). The vibronically induced phosphorescence in benzene. Chem. Phys. 175, 245–254. doi:10.1016/0301-0104(93)85153-Y

CrossRef Full Text | Google Scholar

Minaev, B. F., Lunell, S., and Kobzev, G. I. (1994). Collision-Induced intensity of the b1Σg+−a1Δg transition in molecular oxygen: Model calculations for the collision complex O2 + H2. Int. J. Quantum Chem. 50 (4), 279–292. doi:10.1002/qua.560500405

CrossRef Full Text | Google Scholar

Minaev, B. F., and Minaeva, V. A. (2001). MCSCF response calculations of the excited states properties of the O2 molecule and a part of its spectrum. Phys. Chem. Chem. Phys. 3, 720–729. doi:10.1039/b006712l

CrossRef Full Text | Google Scholar

Minaev, B. F., and Muldakhmetov, Z. M. (1984). Influence of spin-orbit interaction on the intensity of optical doublet-doublet and triplet-triplet transitions in molecules. Opt. Spectrosc. 56 (1), 27–31.

Google Scholar

Minaev, B. F., Norman, P., Jonsson, D., and Ågren, H. (1995). Response theory calculations of singlet-triplet transitions in molecular nitrogen. Chem. Phys. 190, 11–29. doi:10.1016/0301-0104(94)00321-Z

CrossRef Full Text | Google Scholar

Minaev, B. F., Panchenko, A. A., and Hochlaf, M. (2022). Prediction of new emission band 13g-→A3u+ in the spectrum of nitrogen molecule. Sci. Tech. Today 4, 321–338. doi:10.52058/2786-6025-2022-4(4

CrossRef Full Text | Google Scholar

Minaev, B. F., and Panchenko, A. A. (2020). New aspects of the airglow problem and reactivity of the dioxygen quintet O2(5pg) state in the MLT region as predicted by DFT calculations. J. Phys. Chem. A 124, 9638–9655. doi:10.1021/acs.jpca.0c07310

PubMed Abstract | CrossRef Full Text | Google Scholar

Ndome, H., Hochlaf, M., Lewis, B. R., Heays, A. N., Gibson, S. T., and Lefebvre-Brion, H. (2008). Sign reversal of the spin-orbit constant for the C 3Πu state of N2. J. Chem. Phys. 129, 164307. doi:10.1063/1.2990658

PubMed Abstract | CrossRef Full Text | Google Scholar

Ni, C., Cheng, J., and Cheng, X. (2017). Ab initio calculations for the first-positive bands of N2. J. Mol. Spectrosc. 331, 17–22. doi:10.1016/j.jms.2016.10.013

CrossRef Full Text | Google Scholar

Partridge, H., Langhoff, S. R., Bauschlicher, C. W., and Schwenke, D. W. (1988). Theoretical study of the A’5Σ+g and C‘5Πu states of N2: Implications for the N2 afterglow. J. Chem. Phys. 88, 3174–3186. doi:10.1063/1.453962

CrossRef Full Text | Google Scholar

Piper, L. G. (1993). Reevaluation of the transition-moment function and Einstein coefficients for the N2 (A3Σ+u–X1Σ+g) transition. J. Chem. Phys. 99, 3174–3181. doi:10.1063/1.465178

CrossRef Full Text | Google Scholar

Qin, Z., Zhao, J., and Liu, L. (2019). Radiative transition probabilities between low-lying electronic states of N2. Mol. Phys. 117, 2418–2433. doi:10.1080/00268976.2018.1562579

CrossRef Full Text | Google Scholar

Werner, H.-J., Kalcher, J., and Reinsch, E.-A. (1984). Accurate ab initio calculations of radiative transition probabilities between the3Σu+, B3Πg, W 3ΔU, B'3Σu- and C 3Πu states of N2. J. Chem. Phys. 81, 2420–2431. doi:10.1063/1.447917

CrossRef Full Text | Google Scholar

Wilkinson, P. G., and Mulliken, R. S. (1959). Forbidden band systems in nitrogen. II. The a′1Σu ←X1Σg+ system in absorption. J. Chem. Phys. 31, 674–679. doi:10.1063/1.1730445

CrossRef Full Text | Google Scholar

Wilkinson, P. G. (1962). Some unsolved problems in the vacuum ultraviolet. J. Quant. Spectrosc. Radiat. Transf. 2, 343–348. doi:10.1016/0022-4073(62)90020-1

CrossRef Full Text | Google Scholar

Yonker, J. D., and Bailey, S. M. (2019). N2(A) in the terrestrial thermosphere. JGR. Space Phys. 125, e2019JA026508. doi:10.1029/2019JA026508

CrossRef Full Text | Google Scholar

Keywords: triplet–singlet transitions, nitrogen molecule, Vegard–Kaplan band, Wilkinson band, Herzberg I band analog

Citation: Minaev BF, Panchenko OO, Minaeva VA and Ågren H (2022) Triplet state harvesting and search for forbidden transition intensity in the nitrogen molecule. Front. Chem. 10:1005684. doi: 10.3389/fchem.2022.1005684

Received: 28 July 2022; Accepted: 22 September 2022;
Published: 18 October 2022.

Edited by:

Piotr Pander, Silesian University of Technology, Poland

Reviewed by:

Filippo Tamassia, University of Bologna, Italy
Sergey V. Krasnoshchekov, Lomonosov Moscow State University, Russia

Copyright © 2022 Minaev, Panchenko, Minaeva and Ågren. This is an open-access article distributed under the terms of the Creative Commons Attribution License (CC BY). The use, distribution or reproduction in other forums is permitted, provided the original author(s) and the copyright owner(s) are credited and that the original publication in this journal is cited, in accordance with accepted academic practice. No use, distribution or reproduction is permitted which does not comply with these terms.

*Correspondence: O. O Panchenko, panchenko9b@gmail.com

Download