Influence of a strong magnetic field on the hydrogen molecular ion using B-spline-type basis-sets
Zhang Yue-Xia†, Zhang Xiao-Long
Department of Physics, Chongqing University, Chongqing 400044, China

Corresponding author. E-mail: zyuex02@cqu.edu.cn

*Project supported by the National Natural Science Foundation of China (Grant No. 11204389) and the Natural Science Foundation Project of Chongqing (Grant Nos. CSTC2012jjA50015 and CSTC2012jjA00012).

Abstract

As an improvement on our previous work [ J. Phys. B: At. Mol. Opt. Phys.45 085101 (2012)], an accurate method combining the spheroidal coordinates and B-spline basis is applied to study the ground state 1 σg and low excited states 1 σu, 1 πg,u,1 δg,u,2 σg of the in magnetic fields ranging from 109 Gs (1 Gs = 10−4 T) to 4.414 × 1013 Gs. Comparing the one-center method used in our previous work, the present method has a higher precision with a shorter computing time. Equilibrium distances of the states of the in strong magnetic fields were found to be accurate to 3∼5 significant digits (s.d.) and the total energies 6∼11 s.d., even for some antibonding state, such as 1 πg, which is difficult for the one-center method to give reliable results while the field strength is B ≥ 1013 Gs. For the large disagreement in previous works, such as the equilibrium distances of the 1 πg state at B = 109 Gs, the present data may be used as a reference. Further, the potential energy curves (PECs) and the electronic probability density distributions (EPDDs) of the bound states 1 σg, 1 πu, 1 δg and antibonding states 1 σu, 1 πg, 1 δu for B = 1, 10, 100, 1000 a.u. (atomic unit) are compared, so that the different influences of the magnetic fields on the chemical bonds of the bound states and antibonding states are discussed in detail.

PACS: 31.10.+z; 31.15.–p; 31.15.ac
Keyword: magnetic field; B-spline; hydrogen molecular ion
1.Introduction

In our previous paper, [1] the one-center method is applied to study the ground state 1σ g and low-lying states 1σ u, 1π g, u, 1δ g, u, and 2σ g for the within the magnetic field region 109  Gs≤ B ≤ 4.414 × 1013  Gs. The goal of the study has two parts. One is to solve the slow convergence problem existing in the one-center method, which is solved successfully and high-precision energies can be comparable to those obtained by multicenter methods. The other is to study states in all fields regions, which is different from most methods referring to the partial region, [210] except the trial function method.[1113] The B-spline basis plays a critical role in the paper. While the results with high precision are at the cost of the large matrix of the basis. So for some antibonding state, such as 1π g, it is difficult to get results for B > 1000  a.u. using the one-center method. The difficulty of the calculation leads to that only the energies and the equilibrium distances in different magnetic fields can be studied.

Obviously, the results in Ref.  [1] are not sufficient for studying the influence of the magnetic fields for the chemical bond of the states of.[14, 15] The main problem that needs to be solved is to optimize the one-center method. As a test, we combined the spheroidal coordinates and B-spline basis to calculate the equilibrium distance and the total energies of the states with | m| ≤ 4 for B = 1  a.u. in Ref.  [16]. An accuracy from 10− 9– 10− 12 can be obtained for the energies when | m| ≤ 3.

The purpose of this paper is to optimize the method in Ref.  [1] and Ref.  [16] and investigate the influence of the magnetic field on the chemical bond. In this paper, the total energies and equilibrium distances of the ground states 1σ g, u, 1π g, u, 1δ g, u, 2σ g in all the field regions are recalculated. It is helpful for us to learn the advantage of the present method and narrow the gap between different methods. Next, the potential energy curves and the electronic probability density distributions of the states 1σ g, u, 1π g, u, 1δ g, u for a large region of B = 1, 10, 100, 1000  a.u. are plotted. Because the influence of the magnetic fields on the chemical bonds of the bound states and the antibonding states is different, so they are compared and discussed in detail.

2.Method

As shown in Fig.  1, r1 and r2 represent the distances from the electron to proton  1 and proton  2, respectively, R is the internuclear distance, and there is a constant magnetic field along the z axis. The electronic Hamiltonian of within the Born– Oppenheimer approximation is given by

where is the z component of angular momentum and the magnetic field strength γ = B/B0, B0 = 2.35 × 105  T. The atomic units (a.u.) are used in the present study unless specifically stated. Moreover, ρ is the transversal distance, [3] which can be written as

Fig.  1. Model of the in the magnetic fields along the z axis.

In the case of a parallel magnetic field, the Hamiltonian of is cylindrically symmetric, so that the magnetic quantum number m is conserved and the axial wave function can be factored out. The spheroidal coordinate has been proved as a good choice for the system with cylindrical symmetry in many literatures and will be used in the present paper. If we use two variables, u and v, which satisfy:

and

to replace r1 and r2 in Eq.  (1), then the Hamiltonian for a given m in spheroidal coordinates can be read as follows:

with

and

being the kinetic, Coulomb interaction, and magnetic field terms, respectively. Comparing the HC term with that in spherical coordinates, which is

in Ref.  [1], the expression in Eq.  (3) is much more simple. In other words, the present method avoids the complicated numerical calculation on the singularities at the nuclear positions of , which must be considered in the one-center method. Thus a large amount of time can be saved in the present computation.

The wavefunction of the system is expanded in symmetric B-spline type basis set as follows:[17, 18]

with

Here, ku and Nu denote the order and the number of B-spline basis function along the u axis. Similarly, kv and Nv along the v axis. π is the parity quantum number, and π = ± 1 represent even and odd states, respectively. It is worth mentioning that only half of the basis set is needed using the symmetric form in Eq.  (5) comparing with the general B-spline basis in Ref.  [19]. It is a very effective way to reduce the amount of calculation and speed up the convergence in the present method.

In addition, a proper knot sequence is also important for speeding up the convergence. In this project, the u axis knot sequence {uk} can be written as follows:

The knots in the domain [1, umax] are distributed exponentially as follows:

Here, subscript i values from [1, Nuku + 1] and α is an adjustable parameter.

The v axis knot sequence {vj} can be written as follows:

The knot distribution in the domain [− 1, 1] is regarded as a cosine function with

The knot distribution using the cosine function is more effective for speeding up the convergence than that in Ref.  [16], where it is uniform. For example the 1π g state at B = 1  a.u. shown in the following Table  4, 80 × 60(Nu × Nv) basis is needed in Ref.  [16], while 60 × 40(Nu × Nv) basis is enough to obtain the same results in this paper. For more detailed discussions about the knots and the properties of B-splines, readers can refer to Refs.  [1], [20]–   [23]. The eigenvalues and the corresponding wave functions of electronic states are obtained by diagonalization of the effective Hamiltonian  (2) in a space spanned by the basis functions  (4).

3.Results and discussion
3.1.The equilibrium distances and the total energies

In the following, the equilibrium distances and total energies for the states 1σ g, u, 1π g, u, 1δ g, u, and 2σ g of for 109  Gs≤ B ≤ 4.414 × 1013  Gs are given and compared to other results in several tables. In calculations, we fix the parameter umax = 40 and the number of B-spline basis along the u-axis Nu = 60. So in fact, only two parameters, α in Eq.  (6) and the number of B-spline basis along the v-axis Nv need to be adjusted. The detailed numbers Nu and Nv of the B-spline basis for every state in different magnetic fields have been listed in the corresponding tables. It is worth mentioning that the calculation time is significantly reduced compared to that taken by the one-center method in Ref.  [1]. For example, for the 1σ g state at B = 4.414 × 1013  Gs in Table  1, only 4 minutes are needed in the present paper, while one and a half hours were used in Ref.  [1]. Even for the 2σ g state at B = 4.414 × 1013  Gs in Table  7, the number of the basis Nu × Nv = 60 × 700 is the largest in the present paper, about 40 minutes are needed.

3.1.1 1σ g state (m = 0, even parity)

Table  1 lists the equilibrium distances Req and total energies ET for the ground state 1σ g of in different magnetic fields. Compared with the data given by Guan in Refs.  [3] and [24] in fields with strengths below 2.35 × 1010  Gs, the equilibrium distances and total energies obtained in the present method are in close agreement with the last digit listed by Guan. In Comparison with the results at B = 1, 10, 100, 1000  a.u. given by Vincke in Ref.  [8], which are the most accurate found in the literature, the equilibrium distances are consistent with each other at least 5  s.d. and the total energies at least 11  s.d. Compared with previous results obtained by the one-center method in Ref.  [1], the precision of the total energies is improved obviously, though the computation time is reduced largely. That shows the spheroidal coordinates are more suitable than the spherical coordinates for studying in all the field regions from 109  Gs to 4.414 × 1013  Gs. The accuracy of the equilibrium distances is estimated at least about 4  s.d. and that of the total energies about 7∼ 11  s.d. The case is the same as other literatures, the equilibrium distances decrease and the total energies increase as the magnetic field strength increases. This means that the states are increasingly stable as the strength of the magnetic field increases.

Table 1. The equilibrium distances Req and total energies ET for the ground state 1σ g of in different magnetic fields, are here compared to those calculated using other methods.
3.1.2 1σ u state (m = 0, odd parity)

As shown in Table  1, the equilibrium distances Req and the total energies ET for the 1σ u state in magnetic fields with different strengths are listed in Table  2. Most of the calculations on the 1σ u state are concentrated on the domain B ≤ 1010  Gs as shown in Table  2. For the cases of B > 1010  Gs, there are two calculations, i.e., one by Turbiner[12, 13] and the other by us using the one-center method, [1] available. In the same way as in the case of 1σ g, the calculated equilibrium distances and total energies of the state 1σ u in the present method are closer to the data listed by Guan[3, 24] for 109  Gs≤ B ≤ 1010  Gs and Vincke at B = 2.35 × 109  Gs than those obtained in the one-center method.[1] Results in the present method agree with the last digit given by Guan and Vincke. For 2.35 × 1010  Gs≤ B ≤ 1013  Gs, the results are closer to the data obtained by us in the one-center method[1] than Turbiner, [11] including the equilibrium distance and the total energy at B = 2.35 × 1010  Gs, where there is a larger difference between them. For the equilibrium distance from B = 2.35 × 1013  Gs, our data 2.143 is closer in agreement with 2.134 given by Turbiner than 2.02 by the one-center method. Also for B = 4.414 × 1013  Gs, there are no data available using the one-center method. However, the results of the present method are in good agreement with Turbiner’ s; this shows that this method is more suitable for than the one-center method even at magnetic field strengths reaching 4.414 × 1013  Gs. The accuracy of the equilibrium distances is estimated at about 2∼ 5  s.d. and the total energies about 7∼ 11  s.d.

Table 2. Equilibrium distances Req and total energies ET for the excited state 1σ u.

Comparing the numbers of the basis (Nu × Nv) for 1σ u and 1σ g, the amount of computation for 1σ u increases obviously; this may refer to the antibonding behavior of the 1σ u state. It can also explain why less works are found for the 1σ u state than the 1σ g state. Table  2 also shows the equilibrium distances Req decreasing and total energies ET of the 1σ u state increasing as the magnetic field grows from 109  Gs to 4.414 × 1013  Gs.

3.1.3 1π u state (m = − 1, even parity)

We compared the equilibrium distances Req and total energies ET for the excited state 1π u using different methods within a magnetic field ranging from 109  Gs to 4.414 × 1013  Gs as shown in Table  3. The present results are in perfect agreement with the data obtained by Guan in Refs.  [3] and [24], Vincke in Ref.  [11], and us using the one-center method in Ref.  [1]. Even for B = 4.414 × 1013  Gs, the results of the present method are very consistent with those in the one-center method. Turbiner also gave the equilibrium distances Req and total energies ET with the field strength from 109  Gs to 4.414 × 1013  Gs. Their values are less accurate, but they are also closely consistent with our data.

We compare the energies of the 1π u state in Table  3 with those of the 1σ u state in Table  2, and the energy crossing is implied in the range 1012  Gs< B < 2.35 × 1012  Gs.

3.1.4 1π g state (m = − 1, odd parity)

In Table  4, the results of the antibonding state 1π g in different literatures are listed. The equilibrium distances of the 1π g were found to be larger than those of the antibonding state 1σ u. As shown in Table  4, most works focus on the case for B = 2.35 × 109  Gs. Compared with the high-precision results given by Vincke in Ref.  [8], the equilibrium distance is in agreement within 5  s.d. and the total energy within 11  s.d. For B = 109  Gs, the equilibrium distances listed in Ref.  [1] and Refs.  [12] and [13] are 18.08 and 20.10, respectively. The result in the present paper is in good agreement with that using the one-center method in Ref.  [1]. For 1010  Gs≤ B ≤ 2.35 × 1012  Gs, two sets of data, i.e., ours using the one-center method and Turbiner’ s, can be used to compare. The equilibrium distances and the total energies in the present paper are closer to those obtained by the one-center method. The accuracy of the equilibrium distance is estimated at about 3∼ 4  s.d. and the total energy at about 7∼ 8  s.d. For 1013  Gs≤ B ≤ 4.414 × 1013  Gs, only Turbiner’ s data are available. They are also closely consistent with ours.

Table 3. Equilibrium distances Req and total energies ET for the excited state 1π u.
3.1.5 1δ g state (m = − 2, even parity)

In Table  5, the equilibrium distances Req and total energies ET for the excited state 1δ g in strong magnetic fields are compared with the results in other literatures. Two sets of results, i.e., Turbiner’ s and ours obtained by using the one-centermethod, can be found in various magnetic fields. They agree with each other very well, especially the present results and those obtained by the one-center method, where the equilibrium distances are consistent within 3∼ 5  s.d. and the total energies within 6∼ 9  s.d., except the case for B = 2.35 × 1012  Gs. For the equilibrium distance at B = 2.35 × 1012  Gs, the present data 0.3533 is closer to Turbiner’ s 0.353 in Refs.  [12] and [13] than 0.3522 obtained by the one-center method in Ref.  [1].

Table 4. Equilibrium distances Req and total energies ET for the excited state 1π g.

We compared the energies of the 1δ g state to those of the 1σ u state and 1π g state. Two energy crossings are found, one is between the states 1δ g and 1π g in the range from 2.35 × 1010  Gs to 1011  Gs, and the other between the states 1δ g and 1σ u in the range from 1013  Gs to 2.35 × 1013  Gs.

3.1.6 1δ u state (m = − 2, odd parity)

This is another antibonding state of the in the free-field case. The results in different magnetic fields are listed in Table  6. For B = 2.35 × 109  Gs, the present equilibrium distance and the total energy are in prefect agreement with those given by the one-center method in Ref.  [1] and Vincke in Ref.  [8]. It means the present method is also very suitable for the case of the large equilibrium distance. For 109  Gs≤ B ≤ 2.35 × 1012  Gs, the present results have better accordance with those obtained by the one-center method in Ref.  [1] than those given by Turbiner in Ref.  [12] and [13]. The equilibrium distance with at least four effective figures and the energy with at least nine effective figures were consistent with the results in Ref.  [1]. However, for 1013  Gs≤ B ≤ 4.414 × 1013  Gs, the equilibrium distances have large differences from each other. The present results are larger than in another three papers.[1, 12, 13]

Table 5. Equilibrium distances Req and total energies ET for the excited state 1δ g.
Table 6. Equilibrium distances Req and total energies ET for the excited state 1δ u.

Comparing the 1δ u state in Table  4 and the 1π u state in Table  3 at the same magnetic field strength, the equilibrium distances of the 1δ u state are larger than the 1π g state by using the present method. It is different with Turbiner’ s conclusion, where for 1013  Gs≤ B ≤ 4.414 × 1013  Gs, the equilibrium distances of the 1δ u state are smaller than those of the 1π g state.

3.1.7 2σ g state (m = 0, even parity)

In Table  7, the results of the first excited state 2σ g for m = 0 and even parity series are listed. The equilibrium distance of the 2σ g state is larger than that of the ground state 1σ g. Far less investigation of this state in the 1σ g can be found in the literature. Only Turbiner and our group used the one-center method for all magnetic fields from 109  Gs to 4.414 × 1013  Gs and Kappes for B = 2.35 × 109  Gs to investigate this state. The results obtained using different methods are consistent with each other. For example, for 109  Gs≤ B ≤ 1010  Gs, the present equilibrium distances coincide exactly with the results obtained by the one-center method in Ref.  [1] and the energies are in agreement with 5∼ 6  s.d. However, present equilibrium distances and total energies at B = 2.35 × 1013  Gs and 4.414 × 1013  Gs are closer to the data of Turbiner. On the other hand, the one-center method shows weakness while the magnetic field strength is very high, similar to the case of the other antibonding states 1σ u, 1π g, and 1δ u.

Table 7. Equilibrium distances Req and total energies ET for the excited state 2σ g.

We compared the energies of the 2σ g state in Table  7 to those of the 1δ u state in Table  6 and the 1δ g state in Table  5. The two energies were found to cross within the range 109  Gs< B < 2.35 × 109  Gs.

3.2.The potential energy curves for the states 1σ g, u, 1π g, u, and 1δ g, u

To analyze the influence of the different magnetic fields strengths on the ion, the PECs of the 1σ g, u, 1π g, u, and 1δ g, u states in different fields strengths are compared in Fig.  2#cod#x2013; Fig.  4. For every state, the PECs for B = 1, 10, 100, 1000  a.u. are chosen from bottom to top. There solid lines depict the PECs of the bound states 1σ g, 1π u, and 1δ g in different magnetic fields strengths, and dotted lines depict the antibonding states 1σ u, 1π g, and 1δ u. As shown in Fig.  2#cod#x2013; Fig.  4, with increasing nuclear distance R, there is an overlap of the PECs between the bound state and the matching antibonding state with the same magnetic quantum number m. For example, in Fig.  2, the energies of the states 1σ g and 1σ u equal the energies of the state 1s of the hydrogen atom in corresponding fields strengths when the nuclear distance R → ∞ .

Fig.  2. The Pecs of the states 1σ g, u of the with B = 1, 10, 100, 1000  a.u.
Fig.  3. The Pecs of the states 1π g, u of the in different magnetic fields strengths.
Fig.  4. The Pecs of the states 1δ g, u of the in different magnetic fields strengths.

Let us first consider the PECs of the bound states, that is, 1σ g, 1π u, and 1δ g in different magnetic field strengths. A common feature can be easily found. With increasing field strength, the more pronounced potential well located corresponding equilibrium distance is shown. It means stronger attraction exists with increasing magnetic field strength. The depths of the potential well are also given quantitatively by the dissociation energies of Table  8. As expected, for every bound state, the dissociation energies increase with the fields strengths from 1  a.u. to 1000  a.u. If we estimate the influence of the magnetic field on the state according to the ratio of the dissociation energies ED(B)/ED(B = 1  a.u.), they are 2.97, 9.43, 25.9 for the state 1σ g in B = 10, 100, 1000  a.u., respectively. The ratios are 4.24, 15.2, and 45.4 for 1π u and 4.58, 17.2, and 53.3 for 1δ g. For the same magnetic field strength, the ratios of the three states 1σ g, 1π u, and 1δ g increase, such as for B = 1000  a.u., 25.9 < 45.4 < 53.3. It shows the influence of the magnetic field increases with | m| from 0 to 2.

Next, let us take a look at the influence of the magnetic field on the antibonding states 1σ u, 1π g, and 1δ u. As depicted in dotted lines in Fig.  2#cod#x2013; Fig.  4, a shallow potential well exists in the PECs. The depths of the potential wells, that is, the dissociation energies of Table  8, are the order of magnitude of 10− 4 for B = 1  a.u. and 10− 3 for B > 1  a.u. Same as the case of the bound states, dissociation energies of every antibonding state increase with the increasing magnetic field strength from 1  a.u. to 1000  a.u. The ratios of the dissociation energies ED(B)/ED(B = 1  a.u.) are 3.13, 4.88, and 6.17 for 1σ u state in B = 10, 100, 1000  a.u., respectively, and they are 3.74, 5.97, 7.48 for 1π g and 4.23, 6.96, 8.85 for 1δ u. In the same way as in the case of the bound states, the ratios increase with increasing | m| for the same magnetic field strength. It means the same magnetic field strength has a greater influence on the state with larger | m| . For the case of the states with the same m and B, the radios are 25.9 and 6.17 for 1σ g and 1σ u with B = 1000  a.u., respectively. It is not difficult to find that the ratio of the antibonding state is much smaller than that of the corresponding bound state. The conclusion can be drawn that the same change of magnetic field strength has a smaller influence on the dissociation energies of the antibonding states than the corresponding bound states with the same m.

However, as we all known, existence of the potential well is not a criteria to judge whether a molecular bond is stable with respect to dissociation . Only if a vibrational state can exist in the potential well, the molecular bond is stable. To the best of our knowledge, the vibrational energies are referred to in some works. For example, vibrational energies for 1σ g with 109  Gs≤ B ≤ 4.414 × 1013  Gs are given in Refs.  [12] and [13], and the vibrational frequencies for 1σ g, 1π u and 1δ g with B ≤ 104  a.u. are calculated based on the one-center method in Ref.  [30]. Comparing dissociation energies in Table  8 and the vibrational energies in Refs.  [12], [13], and [30], we can find that the dissociation energies are larger than the corresponding vibrational energies. In every state, the differences between these two energies are larger and larger with the increasing magnetic fields. For example, for the state 1σ g in B = 1  a.u., dissociation energy 0.28763869708 in Table  8 is larger than the vibrational energy 0.014 in Refs.  [12], [13], and 0.01328 in Ref.  [30]. It means that some vibrational levels can exist in the potential well of the 1σ g for B = 1  a.u. and the state 1σ g is stable. For the state 1σ g for B = 1000  a.u., the dissociation energy is 7.45682330, and the vibrational energy is 0.390 in Refs.  [12], [13], and 0.3588 in Ref.  [30], respectively. Obviously, the difference between these two energies for B = 1000  a.u. is much larger than that for B = 1  a.u. It means the state is more stable in stronger fields. However, no vibrational energy of the antibonding state in different magnetic fields is given because of the high accuracy, which is hard to achieve. The calculation of the vibrational energy is out of this article’ s scope; therefore, the stability of the antibonding states in strong fields cannot efficiently be discussed in the present paper.

Table 8. Dissociation energies ED of the states 1σ g, u, 1π g, u, and 1δ g, u of with B = 1, 10, 100, 1000  a.u. In the calculation, , and ET(H) is the energy of the hydrogen atom in corresponding magnetic field. The data in Ref.  [31] is used in the present paper. in different magnetic fields, are compared here to those calculated using other methods.
3.3.Electronic probability density distributions for the states 1σ g, u, 1π g, u, and 1δ g, u

The EPDDs in the corresponding equilibrium distance Req in the xz plane are plotted. The states from top to bottom are the bound states 1σ g, 1π u, and 1δ g in Fig.  5, and the antibonding states 1σ u, 1π g, and 1δ u in Fig.  6. The positions of the protons are labelled at the interaction points of the dotted lines along the x axis and the z axis. For every state, the density distributions for B = 1, 10, 100, 1000  a.u. are considered and compared from left to right in Fig.  5 and Fig.  6, which can better explain the change of the PECs with the magnetic fields strengths. For B = 100  a.u. and 1000  a.u., drawings of partial enlargement are placed at the upper right corner. We compare the density distributions of every state from left and right in Fig.  5 and Fig.  6, it is easy to find that the cylindrical symmetry is more obviously about the z axis with the increasing magnetic fields, and the electronic distribution is constricted in a smaller space. However, there is also different changing tendency between the bound states and the antibonding states.

First, let us discuss the change of the density distributions of the bound state 1σ g with the magnetic fields strengths. For B = 1  a.u. and 10  a.u., the electron is found with a great probability in the region around the two nuclei. When the | z | value between the two nuclei gets small, the probability becomes less and less and a saddle point appears at z = 0. When B is added to 100  a.u. and 1000  a.u., the saddle point disappears and a large density distribution of the electron occurs in the region between the two nuclei. The density distributions of the states 1π u and 1δ g are similar. A symmetrical density distribution occurs on both sides of the z axis for B = 1  a.u. With the magnetic field increasing from 1  a.u. to 1000  a.u., the distance of two regions gets closer to the internuclear axis. So for 1π u at B = 1000  a.u., the two regions come together. The electron can occur in the z axis with some probability. The density distribution is closer to the z axis and the increased probability between the two nuclei will screen and reduce the nucleus– nucleus repulsion. Thus, the bond length of the bound states of becomes shorter and shorter with the increasing magnetic fields, and the stronger attraction implies that the depth of the potential well will increase, as shown in the corresponding PECs.

Fig.  5. The electronic probability density distributions of the bound states in the plane xz in different magnetic fields. States from top to bottom are 1σ g, 1π u, and 1δ g. The drawings of partial enlargement are placed at the upper right corner.

Fig.  6. The electronic probability density distributions of the antibonding states in the plane xz in different magnetic fields. States from top to bottom are 1σ u, 1π g, and 1δ u. The drawings of partial enlargement are placed at the upper right corner.

The EPDDs of the antibonding states are different from those of the bound states. For example, the density distribution for 1σ u in B = 1  a.u., mainly concentrate in the range of approximately − 1 < x < 1 and 4 < | z | < 6. There is a long range defined by − 4 < z < 4 between the two nuclei. No electron is found. The symmetrical distribution of the electron about the positions z = ± Req/2 shows the influence on the electron from the farther nucleus is a lot smaller than the nearer nucleus. With the magnetic field increasing, the distance of the two nuclei becomes shorter. Obviously the compression is the result of the effect of the magnetic field being greater than the repulsion between the two nuclei. It also shows that a stronger attraction occurs with the increasing magnetic field. However for B = 1000  a.u., the electronic distribution of positions z = ± Req/2 becomes asymmetric. It reflects that for B = 1000  a.u., the effect on the electron from the nucleus farther away cannot be ignored when the distance of the two nuclei is 2.848  a.u. The density distributions of the 1π g and 1δ u are similar. The electron mainly occurs in four regions of both sides of the z axis. With the magnetic field increasing, the region of the distribution is shrinking along the direction of parallel and perpendicular fields. For the state 1π g in B = 1000  a.u., these four regions of distribution merge into two around the z axis. Similar to the case of the 1σ u in B = 1000  a.u., the symmetry about the positions z = ± Req/2 is also destroyed because of the effect from the nucleus farther away. The similar effect is also found for the state 1δ g in B = 100  a.u. and 1000  a.u. The more significant asymmetry of the density distribution for the states 1σ u, 1π g, and 1δ u for B = 1000  a.u. shows the relative effect of the other nucleus farther away increases with larger | m| from 0 to 2.

4.Summary

In conclusion, we successfully applied the method combining the spheroidal coordinate and B-spline to calculate the equilibrium distances and the total energies of the states with | m| ≤ 2 of for 109  Gs≤ B ≤ 4.414 × 1013  Gs. Compared with the one-center method used in our previous work, the present method avoids dealing with the singularities at the nuclear positions of . So the amount of computations and time are largely reduced in the present calculation. It speeds up about 40 minutes for calculating the results of 2σ g state of B = 4.414 × 1013  Gs, which is the largest base in the present calculation. However, the accuracy of the results is improved obviously in comparison with other high-precision data. The present equilibrium distances are accurate about 3  s.d.∼ 5  s.d. and the total energies about 6  s.d.∼ 11  s.d., even for the antibonding states 1π g, which is difficult to obtain the results for B ≥ 1013  Gs using the one-center method. Moreover, for several larger differences of the equilibrium distances between Ref.  [1], Refs.  [12] and [13], the present results may serve as a reference. Such as that for the 1π g state of B = 109  Gs, the present datum 18.08 is perfectly consistent with that in Ref.  [1]. However, for the 1δ u state of B = 4.414 × 1013  Gs, the present datum 2.38 is in closer agreement with 2.230 in Refs.  [12] and [13].

Further, the PECs and dissociation energies of the bound states and the anbibonding states for B = 1, 10, 100, 1000  a.u. show that the depth of the potential well increases with the increasing fields strengths. The stronger behavior of the attraction is reflected for a stronger magnetic field. Comparing the corresponding vibrational energies of the bound states in Refs.  [12], [13] and [30], we can find that many vibrational states can exist in the potential wells of the bound states. It means the bound states 1σ g, 1π u, 1δ g are more stable in stronger magnetic fields. However, we cannot draw a conclusion on whether antibonding states can exist stably in strong fields due the accurate vibrational energies of the antibonding states. The increasing magnetic fields can cause the electron clouds to contract along both the parallel and perpendicular magnetic fields. However, the reason is different for the bound states and the antibonding states. For the bound states, EPDDs show that more and more electrons distribute in the region between the two nuclei with increasing magnetic fields. The screening of the two nuclear charges can explain the bond lengths decreasing with stronger fields strengths. For the antibonding states, the effect of the strong magnetic field is far greater than the Coulomb repulsion between the two nuclei, which is the reason why the bond length of the antibonding states decreases with the increasing magnetic field. With the distance between two nuclei closing, the electron is also influenced by the other nucleus farther away, and the symmetry of the electron clouds on the positions z = ± Req/2 changes. It is obvious for the EPDDs of the states 1σ u, 1π g for B = 1000  a.u. and 1δ g for B = 100, 1000  a.u. The asymmetry shows also that the influence of the nucleus farther away from the electron increases with larger | m| at the same magnetic field strength.

The present method can provide satisfactory results with a small amount of computation and save much computing time. The results we obtained in the present paper encourage us to extend the present method to study the influence of magnetic field strengths with different inclinations on the equilibrium distances, total energies, and molecular bonds of . This may provide a way to better understand the spectrum of the isolated neutron star 1E1207.4-5209.

Reference
1 Zhang Y X, Liu Q and Shi T Y 2012 J. Phys. B: At. Mol. Opt. Phys. 45 085101 DOI:10.1088/0953-4075/45/8/085101 [Cited within:22]
2 Wille U 1988 Phys. Rev. A 38 3210 DOI:10.1103/PhysRevA.38.3210 [Cited within:1]
3 Guan X X, Li B W and Taylor K T 2003 J. Phys. B: At. Mol. Opt. Phys. 36 3569 DOI:10.1088/0953-4075/36/17/302 [Cited within:4]
4 Kappes U, Schmelcher P and Pacher T 1994 Phys. Rev. A 50 3775 DOI:10.1103/PhysRevA.50.3775 [Cited within:1]
5 Kappes U and Schmelcher P 1994 J. Chem. Phys. 100 2878 DOI:10.1063/1.466430 [Cited within:1]
6 Kappes U and Schmelcher P 1995 Phys. Rev. A 51 4542 DOI:10.1103/PhysRevA.51.4542 [Cited within:1]
7 Kappes U and Schmelcher P 1996 Phys. Rev. A 53 3869 DOI:10.1103/PhysRevA.53.3869 [Cited within:1]
8 Vincke M and Baye D 2006 J. Phys. B: At. Mol. Opt. Phys. 39 2605 DOI:10.1088/0953-4075/39/11/023 [Cited within:3]
9 Kravchenko Y P and Liberman M A 1997 Phys. Rev. A 55 2701 DOI:10.1103/PhysRevA.55.2701 [Cited within:1]
10 Baye D, Joos de ter Beerst A and Sparenberg J M 2009 J. Phys. B: At. Mol. Opt. Phys. 42 225102 DOI:10.1088/0953-4075/42/22/225102 [Cited within:1]
11 Turbiner A V and López Vieyra J C 2003 Phys. Rev. A 68 012504 DOI:10.1103/PhysRevA.68.012504 [Cited within:3]
12 Turbiner A V and López Vieyra J C 2004 Phys. Rev. A 69 053413 DOI:10.1103/PhysRevA.69.053413 [Cited within:12]
13 Turbiner A V and López Vieyra J C 2006 Phys. Rep. 424 309 DOI:10.1016/j.physrep.2005.11.002 [Cited within:13]
14 Schmelcher P and Schweizer W 1998 Atoms and Molecules in Strong External Fields New York Plenum [Cited within:1]
15 Sanwal D, Pavlov G G, Zavlin V E and Teter M A 2002 Astrophys. Lett. 574 L61 DOI:10.1086/342368 [Cited within:1]
16 Zhang Y X, Liu Q and Shi T Y 2013 Chin. Phys. Lett. 30 043101 DOI:10.1088/0256-307X/30/4/043101 [Cited within:4]
17 Yulian V V and Alejand ro S 2004 J. Phys. B: At. Mol. Opt. Phys. 37 4101 DOI:10.1088/0953-4075/37/20/005 [Cited within:1]
18 Zhang Y X, Kang S and Shi T Y 2008 Chin. Phys. Lett. 25 3946 [Cited within:1]
19 Shi T Y, Zhang Z J and Li B W 2004 Mod. Phys. Lett. B 18 113 DOI:10.1142/S0217984904006743 [Cited within:1]
20 Kang S, Li J and Shi T Y 2006 J. Phys. B: At. Mol. Opt. Phys. 39 3491 DOI:10.1088/0953-4075/39/17/007 [Cited within:1]
21 Qiao H X and Li B W 1999 Phys. Rev. A 60 3136 [Cited within:1]
22 Xi J H, Wu L J, He X H and Li B W 1992 Phys. Rev. A 46 5806 DOI:10.1103/PhysRevA.46.5806 [Cited within:1]
23 Shi T Y, Qiao H X and Li B W 2000 J. Phys. B: At. Mol. Opt. Phys. 33 L349 DOI:10.1088/0953-4075/33/9/106 [Cited within:1]
24 Guan X X, Li B W and Taylor K T 2004 J. Phys. B: At. Mol. Opt. Phys. 37 1985 DOI:10.1088/0953-4075/37/9/C01 [Cited within:3]
25 López Vieyra J C, Hess P and Turbiner A V 1997 Phys. Rev. A 56 4496 DOI:10.1103/PhysRevA.56.4496 [Cited within:1]
26 Larsen D M 1982 Phys. Rev. A 25 1295 DOI:10.1103/PhysRevA.25.1295 [Cited within:1]
27 Lai D, Salpeter E and Shapiro S L 1992 Phys. Rev. A 45 4832 DOI:10.1103/PhysRevA.45.4832 [Cited within:1]
28 Vincke M and Baye D 1985 J. Phys. B: At. Mol. Opt. Phys. 18 167 DOI:10.1088/0022-3700/18/2/005 [Cited within:1]
29 Ozaki J and Hayashi Y 1989 J. Phys. Soc. Jpn. 58 3564 DOI:10.1143/JPSJ.58.3564 [Cited within:1]
30 Hu S L and Shi T Y 2013 Chin. Phys. B 22 093101 DOI:10.1088/1674−1056/22/9/093101 [Cited within:5]
31 Baye D, Vincke M and Hesse M 2008 J. Phys. B: At. Mol. Opt. Phys. 41 055005 DOI:10.1088/0953-4075/41/5/055005 [Cited within:1]