Sie sind auf Seite 1von 11

Spectrochimica Acta Part B 64 (2009) 863–873

Contents lists available at ScienceDirect

Spectrochimica Acta Part B


j o u r n a l h o m e p a g e : w w w. e l s ev i e r. c o m / l o c a t e / s a b

A few level approach for the electronic partition function of atomic systems
Gianpiero Colonna a,⁎, Mario Capitelli b
a
CNR IMIP Bari Via Amendola 122/d, 70126, Bari, Italy
b
Department of Chemistry, University of Bari, Via Orabona 4, 70126, Bari, Italy

a r t i c l e i n f o a b s t r a c t

Article history: A simplified model to calculate partition functions and thermodynamic properties of atomic species is
Received 18 March 2009 presented. This model consists in grouping the atomic states in few virtual levels. Their statistical weights
Accepted 1 July 2009 and energies are calculated summing or mediating over the states belonging to each group. The partition
Available online 16 July 2009
function is calculated considering the virtual levels to follow the Boltzmann distribution. A theoretical
foundation of the model has been described for a general case and verified for hydrogen, oxygen and nitrogen
PACS:
atoms. Two- and three-level models are adequate to keep the error within a few percent. A second order
51.30. + i
52.38.Mf
correction term can also be added to further reduce the error without changing the advantages of the model.
© 2009 Elsevier B.V. All rights reserved.
Keywords:
Atomic partition function
Statistical thermodynamic
Superconfiguration

1. Introduction polating and interpolating observed levels by using Ritz-Rydberg formu-


lae [13,14]. In any case, a very large number of electronic levels arise so that
Partition functions and thermodynamic properties of atomic computation time can exponentially increase specially if a thermody-
systems are a research field in continuous evolution due their namic model has to be self-consistent with flow (e.g. Navier-Stokes)
importance in astrophysical applications [1–4] and in plasma spectro- equations [15,16]; in this kind of models, being the cutoff dependent on
scopy [5–7], with particular attention to LIBS [8–10]. In particular pressure, temperature and composition [17,18], the single species
spectroscopy of LTE (Local Thermodynamic Equilibrium) plasmas thermodynamic properties of the mixture must be recalculated any
makes extensive use of electronic partition functions and thermo- time one of these quantities changes. A further complication arises when
dynamic properties of the relevant components. As an example the trying to insert in the partition function levels near to the continuum
absolute intensity I of a line involving u (upper) and l (lower) levels of under extremely low pressure and high temperature conditions [19].
the a atomic species can be written as To better understand how complex this problem is let us consider
two specific examples i.e. atomic hydrogen and nitrogen systems. In the
 ε 
1 N first example we consider a low pressure plasma for which a suitable
Iul = Aul hνul a gu exp − u ð1Þ
4π Qa T cutoff criterion allows insertion of electronic levels up to the principal
quantum number n = 50. Keeping in mind that the degeneracy of a
where Aul is the Einstein coefficients, hvul the photon energy, Na the atom given level is 2n2 we can understand the very large number of levels to
density, Qa the internal partition function, εu and gu are respectively the be inserted in the partition function. In the case of atomic hydrogen the
energy and the statistical weight of the u state. The use of absolute small dependence of electronic energy on the angular and magnetic
intensities to determine the atom density needs accurate values of the quantum numbers prevents the splitting of energy levels so that we can
partition function, a problem underestimated due to the difficulties in its sum for the case under study only 49 terms plus the ground state.
evaluation. Moreover to reproduce the absolute intensity of a spectral line Let us now discuss the case of atomic nitrogen N(4S). The principal
N series of electronic levels in this system arises from the combination of
one has to self-consistently calculate the ratio a [11]. These quantities can
Qa
be calculated when a complete set of energy levels subjected to a the core 3P interacting with the excited electron with n N 2. In this case
given cut-off criterion is available [12]. Complete set of levels means a the number of excited states becomes 2gc n2 where gc = 9 is the
database including both observed and unobserved levels. The last can be statistical weight of the 3P core of N(4S) system [20]. This means that, as a
calculated by using quantum mechanical approaches as well as by extra- consequence of the dependence of the energy on the angular and
magnetic quantum numbers, already for n = 3 we have 162 non-
⁎ Corresponding author.
degenerate states and the number of different levels rapidly grows as the
E-mail addresses: gianpiero.colonna@ba.imip.cnr.it (G. Colonna), principal quantum number increases. In addition, the atomic nitrogen
mario.capitelli@ba.imip.cnr.it (M. Capitelli). system presents the so called low lying excited states resulting from the

0584-8547/$ – see front matter © 2009 Elsevier B.V. All rights reserved.
doi:10.1016/j.sab.2009.07.002
864 G. Colonna, M. Capitelli / Spectrochimica Acta Part B 64 (2009) 863–873

redistributions of electrons in the 2p electronic states. These redistribu- 2. Statistical thermodynamics


tions must fulfill Pauli's principle getting the well known 2D and 2P states
with statistical weights and energy respectively equal to 10 and 6 and The partition function is the starting point to determine all the
2.38 eV and 3.58 eV. These states, in turn, can contribute to the creation thermodynamic properties of a species [12]. For a species s in gas or
of other series in the “zoology” of electronic levels by the interaction of plasma phase, the internal partition function, when the temperature T
the respective cores with the excited electron [20]. Some of these states is expressed in energy unit, is given by
are in the continuum of the first ionization potential; these levels,
Ns εsl
commonly called self–ionizing, are generally excluded (without a −
Qs = ∑ gsl e T ð2Þ
serious discussion) in the calculation of the partition function. The l=0
history is not ended yet because other series can arise from the
interaction of ‘exotic’ cores and the excited electron as well as by the where Ns in the number of levels, gsl and εsl respectively statistical
contemporaneous excitation of more than one electron. weight and energy of the l-th level. The internal partition function is
Another complication is the variation of the level energies due to related to the internal Helmhotz free energy by the equation
Debye screening of charged species [21,22] in plasma environment,
int
resulting in the loss of degeneracy also for the hydrogen atom. In high As = T ln Qs ð3Þ
pressure regimes, this contribution can be relevant as shown in Ref.
[23] for hydrogen plasma. while internal energy (Us) and specific heat (Cv) can be calculated
These considerations are indicative of the difficulty of calculating the from its derivatives with respect to the gas temperature [12]. It can be
electronic partition functions by a level-by-level approach. Different easily demonstrated that the internal energy is the sum of the
approximations, depending on the availabilty and accuracy of electronic translational and internal contributions
states, are currently used [14,24]. These approximations however cannot T int
be inserted in 2D and 3D fluidynamic models, because of the million Us = Us + Us ð4Þ
points of the computational grid used to integrate the relevant differential
where the translational contribution is independent of the species
equations. The problem of rapid evaluation of the atomic partition
functions, to keep the overall computational time to reasonable values, 3
T
has been investigated considering hydrogen-like excited energy levels to Us = T: ð5Þ
2
obtain an approximate analytical expression for the partition function
[25], resulting in analytical expression of the themodynamic properties as The internal contribution is given by
a function of the number of levels considered. Another paper [26]
ε
discusses a reduced model to calculate the plasma composition for a int 2 d lnQs 1 dQs 1 Ns − sl
Us = T =−   = ∑ g ε e T = ≪ε≫s ð6Þ
mixture of different atomic species (neutrals and ions) solving a single dT Qs d 1 Qs l = 0 sl sl
transcendental equation obtained by the system of Saha's equations and T
adding a nonlinear equation to determine, in the same time, the Debye
considering ≪ · ≫ s the statistical thermodynamic mean over the
length and the atomic partition functions. These approaches use some
Boltzmann distribution. The constant volume specific heat is given by
physical properties to determine a simplified model, but cannot be ex-
tended to general case of a multicomponent non-ideal plasma equilibrium
[23,27–30], especially if the partition function must be evaluated directly dUs 3 int
Cvs = = + Cvs ð7Þ
by summing the Boltzmann distribution over the whole level system. dT 2
The aim of this paper is to show the possibility of calculating the
where
partition function of atomic species (neutral and ionized) by reducing
the electronic ladder to a few level problem. The idea is to group all the
int dUsint 1 2 2
levels available in the databases to few lumped levels with suitable Cvs = = 2 ð≪ε ≫s −≪ε≫s Þ ð8Þ
dT T
average energy and global statistical weight, being these quantities
dependent only on the number of levels belonging to the group. This being
approach, when considering only the principal series coming from the
interaction of the more stable core and of excited electron, ends in a two- ε
2 1 Ns 2 −
sl
1 d2 Qs
level (ground and lumped) partition function and to a three-level ≪ε ≫s = ∑ gsl εsl e T =   : ð9Þ
Qs l = 0 Qs d 1 2
system (ground, lumped low lying and lumped high lying levels) for T
systems presenting low lying excited states. Particular attention has
been given to the mathematical framework of our simplified model, It must be pointed out that, being the temperature in energy unit, the
emphasizing the errors introduced by the reported procedure. A general specific heat is a pure number. From the calculated quantities it is
theory of multilevel approach has been widely discussed, demonstrating possible to determine all the other thermodynamic state functions
its validity if the energy levels are not much dispersed around the mean such as enthalpy (H), Gibbs free energy (G) and entropy (S)
value. The mathematical foundation of the theory is based on the Taylor
series expansion: the first order term is null, making the multi-group 8
>
> Hs = Us + T
approach accurate at the first order. The second order correction can be >
<
Gs = As + T
also calculated, increasing the accuracy of the method without changing : ð10Þ
>
>
the main properties of the theory. >
: Ss = 1 ðUs −As Þ
The structure of the paper is as follows. Section 2 reports the T
summary of the equations describing the partition function and
thermodynamic properties of single species. Section 3 discusses the 3. Atomic partition function
mathematical basis for reducing the electronic partition function to a
few level system, emphasizing in particular the two-level (ground and The calculation of the internal partition function of atomic species
lumped) model. Relevant errors normalized to the energy dispersion are presents some problems: being the number of levels infinite, the
discussed. Section 4 discusses real cases including atomic hydrogen, partition function always diverges [31]. However, there are infinite
nitrogen and oxygen. Section 5 reports conclusions and perspectives. levels only in an infinite spacecontaining a single atom, while the
G. Colonna, M. Capitelli / Spectrochimica Acta Part B 64 (2009) 863–873 865

interaction with other particles limits their number. This problem is 1 N


ε= ∑ g ε = 〈ε〉 ð19Þ
solved truncating the partition function to a given energy limit G l=1 l l

εsl b εM εsl
s
− where 〈·〉 is the classical weighted statistical mean. Under these
Qs = ∑ gsl e T ð11Þ
l=0 assumption the mean energy (see Eqs. (4)–(6)) is given by

where ε  
⋆ 3 1 − ε 3 1 −z
U = T + ⋆ Gεe T = + ⋆ Gze ð20Þ
2 Q z 2 Q
M
εs = Is −Δs ð12Þ
and the specific heat (see Eqs. (7) and (8)) by
Is being the ionization potential and Δs the energy cutoff.
Some databases reporting partition functions and other internal 2 ε ε 2 ε 2
⋆ 3 Gε − ⋆ − 3 Gg ε − 3 Gg0 z −z
thermodynamic properties for different cutoffs are available [14,24]. It Cv = + 2 ⋆2 e T Q −Ge T = + 2 0 ⋆2 e T = + ⋆2
e
2 T Q 2 T Q 2 Q
must be pointed out that in real systems the cutoff is not constant but
it depends on some quantities such as particle density (Fermi ð21Þ
criterion) or electron density (Griem criterion). As a consequence,
the cutoff can be determined once the plasma composition is known. The demonstration of Eqs. (17)–(21) starts rewriting Eq. (17) by
This approach increases the computational time necessary to separating the contribution of the ground states from excited levels
ε

determine the atomic properties, especially when kinetic or thermo- and by multiplying and dividing the last term by the factor e T

dynamic calculations are coupled with complex fluid dynamic codes


[15]. Another possible approach consists in storing the thermodynamic ε N ε −ε
− − l
quantities as a function of both temperature and cutoff, requiring a large Q = g0 + e T ∑ gl e T : ð22Þ
l=1
amount of memory for data storage [14].
In this paper we will show that it is possible to describe the
Each exponential term inside the summation can be expanded in
contribution of an internal level subset with a single level
Taylor series
fin
Lsk εsl εsk
− − ε −ε
∑ gsl e T ≈Gsk e T ð13Þ − l εl −ε
e T = 1− + τl ð23Þ
l = Lin
sk T

approximating the internal partition function as being τl the truncation error, and calculating the sum over the levels

Nsgr εsk

ε
εl −ε −
ε
G −
ε

ε
⋆ −
Qs ≈Qs = ∑ Gsk e T ð14Þ Q = g0 + Ge T 〈1− + τl 〉 = g0 + Ge T − e T ð〈ε〉−εÞ + Ge T 〈τ〉:
T T
k=0
ð24Þ
where the number of groups (Ngr s ) is much smaller than the number of
levels. As we will show later, the group statistical weight (Gsk) and Being the ε = 〈ε〉 by definition (see Eq. (19)), the third term in Eq. (24)
energy (Esk), given by is null, obtaining

ε ε ε
Lfin − − ⋆ −
Gsk = ∑ gsl
sk
ð15Þ Q = g0 + Ge T + Ge T 〈τ〉 = Q + Ge T 〈τ〉 ð25Þ
l = Lin
sk

and giving the error


fin
Lsk
1 ε
εsk = ∑ g ε ð16Þ ⋆ −
Gsk l = Lin sl sl δQ = Q −Q = Ge T 〈τ〉: ð26Þ
sk

are good approximations under the condition that the mean energy of This result shows that Eq. (17) is accurate at the first order. Some
the group is much higher than the level energy variance. The symbols considerations can explain why this approximation works. Supposing
fin
Lin
sk and Lsk are the first and last level belonging to the k-th group. This that the level dispersion around the mean energy is small with respect
method is very useful because the quantities E and G do not depend to the mean energy. The Taylor expansion can be used only if the
on the temperature and only the influence of the cutoff must be taken temperature is not too low. In particular this approximation can be
into account. safely used if the temperature is of the same order as the mean energy
Let's consider the simple case of two groups of levels, the first one or higher. On the other hand, for temperature much lower than
includes only the ground state, with zero energy, and the second one all the mean energy, the exponential is very small, and therefore also
the excited levels. To simplify the notation the group and species indexes in this case the expression is accurate, even if the Taylor series
are omitted for the rest of this section. We want to demonstrate that truncated at the first term cannot be used. The value of the error can
be limited by the following relations. For the properties of the ex-
εl ε
N
− ⋆ − −z ponential function
Q = ∑ gl e T ≈Q = g0 + Ge T = g0 + Ge ð17Þ
l=0
δQ N 0 ð27Þ
ε
where z = and ε and G are calculated following the Eqs. (15)
T
and (16) i.e.
and, considering the levels ordered by increasing energies
N
εN ε1
G = ∑ gl ð18Þ − ⋆ −
l=1 g0 + Ge T ≤Q ≤Q≤g0 + Ge T ð28Þ
866 G. Colonna, M. Capitelli / Spectrochimica Acta Part B 64 (2009) 863–873

and subtracting to all terms Q⁎ we have


 εN ε  ε1 ε
− − − −
G e T −e T ≤ 0 ≤ δ Q ≤ G e T −e T : ð29Þ

A more rigorous evaluation of the error can be done extending the


Taylor series to one more term

1 2
τl ≈ ðεl −εÞ ð30Þ
2T 2
ε ε
G −T 2 G −T 2 G 2 −z 2 ⋆
δQ ≈ e 〈ðεl −εÞ 〉 = e σε = z e δε = δQ ð31Þ
2T 2 2T 2 2

where σε2 is the variance and

σε
δε = ð32Þ
ε

is the relative level energy dispersion. The temperature of maximum


absolute error is given by the solution of the equation
 
dðδQ ⋆ Þ Gz −z 2 ⋆ 2−z
= e δε ð2−zÞ = δQ = 0: ð33Þ
dz 2 z Fig. 1. Value of the relative error of the partition function (see Eq. (35)) reduced by the
factor δ2ε as a function of z for different values of α. The value of α is varied by a factor 10.
This equation has three roots
8
>
> z=∞ ⇒ T=0 ⇒ δQ ⋆ = 0
>
< is 1% at the maximum value of α results in an error in the partition
z=0 ⇒ T =∞ ⇒ δQ ⋆ = 0 : ð34Þ function below 1%.
>
>
:z=2 ⇒ T = ε ⇒
> δQ ⋆ = 2Gδ2ε e−2 To study the propagation of error on the internal energy (see
2 Eq. (4))

This means that to the very low and very high temperature, as already
   
discussed, the error tends to be null, and the maximum error is δU ⋆ dQ dQ ⋆ 1 dQ dQ ⋆ dδQ dδQ ⋆
observed for temperature values of the same order of the mean =− − ⋆ ≈− ⋆ − =− ⋆ ≈ ⋆ ð37Þ
ε Qdz Q dz Q dz dz Q dz Q dz
energy. The maximum error of the partition function is of the order of ⋆ 
δQ z−2
the square of the relative level energy dispersion and the total = ⋆
Q z
statistical weight.
The relative error can be approximated as

δQ ⋆ αz2 e−z δ2ε αz2 e−z 2 δQ ⋆


= ≈ −z δε = ð35Þ
Q −z −z
2 + 2αe + αz e δε 2 + 2αe
2 2 Q⋆

G
which is almost proportional to δ2ε , where α = . The plot of the
g0
relative error of partition function, normalized by the square of the
dispersion has been reported in Fig. 1. We can observe that for high
temperature, the relative error is independent of α showing a power
dependence on z. Close to the maximum it depends weakly on α,
decreasing rapidly.
The maximum of the relative error is calculated finding the roots of
the derivative

   ⋆
d δQ ⋆ 1 dδQ ⋆ ⋆ dQ 1 ⋆ dδQ

⋆ dQ
= 2 Q −δQ = 2 Q −δQ
dz Q Q dz dz Q dz dz
 ⋆

1 −z δQ ⋆ −z ð36Þ
= 2 ðg0 + Ge Þ ð2−zÞ + δQ Ge
Q z
⋆ ⋆
δQ −z g0 δQ −z
= ½g ð2−zÞ + 2Ge = ½ð2−zÞ + 2αe  = 0:
zQ 2 0 zQ 2

 
ε
The root of Eq. (36) zM = ⋆ can be found numerically and it
T
depends only on α. In Fig. 2 we have reported the root of Eq. (36),
which gives the temperature of maximum error, and the correspond-
ing value of the relative error, reduced by the factor δ2ε as a function of Fig. 2. Value of zM = Tε⋆ (root of Eq. (36)) and the corresponding maximum value of the
the ratio α. It can be observed that these quantities depends on log α partition function relative error (reduced by the factor δ2ε ) as a function of the ratio
and that their values are not very large. In fact if the energy dispersion α = gG0 .
G. Colonna, M. Capitelli / Spectrochimica Acta Part B 64 (2009) 863–873 867

where roots are those in Eq. (34). As a consequence the maximum


absolute error in the partition function results in a small error in the
internal energy. The small error is also obtained for small and high z,
as for δQ⁎. The relative error is given by

δU ⋆ 2δQ ⋆ ðz−2Þ αz2 ðz−2Þe−z 2


⋆ = −z = δ ð38Þ
U ⋆
3Q + 2Gze 3 + 2αze−z ε

and has been reported in Fig. 3 as a function of z for different values of


α. The maximum of the relative error of internal energy is obtained by
solving the equation

d δU ⋆ −z 2 2
= 0⇒αe ð4z + 5z−12Þ−3ðz −5z + 4Þ = 0 ð39Þ
dz U⋆

and the values of the maximum error are reported in Fig. 4. It must be
noted that the maximum of the internal energy relative error is of the
same order as the maximum of the relative error of the partition
function (see Fig. 2).
For the specific heat, due to the additive property, it is

2 ⋆ ⋆
⋆ int⋆ z dδU δQ ⋆
δCv = δCv =− = ⋆2 ½2ð1−zÞQ + g0 zðz−2Þ ð40Þ Fig. 4. Value of zM =
ε
(root of Eq. (39)) and the corresponding internal energy relative
ε dz Q T⋆
G
error reduced by the factor δ2ε as a function of the ratio α = .
g0

and the relative error is given by (see Eq. (21))

δCv⋆ ⋆
⋆ 2ð1−zÞQ + g0 zðz−2Þ in Fig. 6. Compared with the relative error of the partition function
= 2δQ : ð41Þ
Cv⋆ 3Q + 2g0 Gz2 e−z
⋆2 (see Fig. 2) and of the internal energy (see Fig. 4) the one of the
specific heat is 4 times larger.
We have shown that the first order approximation results in a
This expression, divided by the level energy dispersion, has been
simple reduced level representation. It must be stressed that
plotted as a function of z in Fig. 5. It can be observed the function trend
increasing the order of approximation does not increase the
is more complex than the previous quantities, as well as the analytical
complexity of the simplified scheme, and, furthermore, uses
expression of its derivative. Moreover two different maxima are
coefficients as σε that do not depend on the temperature, keeping
observed, the second is larger than the first one at low α but this
the same advantages as the first order approximation, but reducing
relation is inverted as α increases. Both maxima have been calculated
evaluating the derivative numerically. The results have been reported

Fig. 5. Value of the relative error of the constant volume specific heat (see Eq. (41))
Fig. 3. Value of the relative error of the internal energy U (see Eq. (38)) reduced by the reduced by the factor δ2ε as a function of z for different values of α. The value of α is
factor δ2ε as a function of z for different values of α. The value of α is varied by a factor 10. varied by a factor 10.
868 G. Colonna, M. Capitelli / Spectrochimica Acta Part B 64 (2009) 863–873

Fig. 6. Position zM of the two maxima (see Fig. 5) (lower frame) and the corresponding
relative error reduced by the factor δ2ε (upper frame) for constant volume specific heat
G
as a function of the ratio α = .
g0

the error. The second order approximation can be simply obtained by


considering

8 ⋆⋆ ⋆ ⋆
< Q = Q + δQ
>
⋆⋆ ⋆ ⋆
U = U + δU : ð42Þ Fig. 7. Total statistical weight (G), mean energy (E [eV]) and square of the relative level
>
: ⋆⋆ energy dispersion (δ2ε ) (see Eqs. (45) and (47)) of excited levels as a function of the
⋆ ⋆
Cv = Cv + δCv maximum value of the principal quantum number (N) for hydrogen atoms.

From a computational point of view, it is possible to calculate the


second order correction by the difference of the first and second order 4.1. Hydrogen atom
approximations (see Eqs. (6)–(9)), calculated considering the first
and second derivative of δQ (see Eq. (33)), given by As first case study we consider the hydrogen atom. For this atom,
8 energies and statistical weights are given by simple formulas
>
> dδQ dδQ 2−z
>
>   = ε = εδQ  
>
> 1 dz z 1
<d T εn = IH 1− 2
: ð43Þ n ð44Þ
>
>
2
d δQ
2
2 d δQ 2−4z + z
2
>
>   ε ε
2
δQ
>
>
= = gn = 2n2
:d 1 2 dz2 z2
T
where n is the principal quantum number (see Table 1). For this
system we have
4. Real cases
N NðN + 1Þð2N + 1Þ−6
In the previous section the accuracy of the proposed method has G = ∑ 2n2 =
n=2 3
been discussed only on the bases of the statistical parameters of the   ð45Þ
distribution of level energies around the mean value. To validate the ε 1 N
2 1 N−1
= ∑ 2n 1− 2 = 1−
theory real examples must be considered. IH G n=2 n G

Table 1
Statistical weight, energy (eV) and electronic configuration of the first levels of H, N and O atoms.

i Hydrogen Nitrogen Oxygen


config. g E [eV] config. g E [eV] config. g E [eV]
0 n=1 2 0.0 2s22p3(4S0) 4 0.0 2s22p4(3P) 5 0.0
1 n=2 8 10.2040 2s22p3(2D0) 10 2.38396 2s22p4(3P) 3 0.0196515
2 n=3 18 12.0936 2s22p3(4P0) 6 3.57561 2s22p4(3P) 1 0.0280824
3 n=4 32 12.7549 2s22p3(3P) 3s(4P) 12 10.3322 2s22p4(1D) 5 1.96734
4 n=5 50 13.0611 2s22p3(3P) 3s(2P) 6 10.6865 2s22p4(1S) 1 4.18973
5 n=6 72 13.2273 2s22p4(4P) 12 10.9270 2s22p3(4S0) 4s(5S0) 5 9.14605
6 n=7 98 13.3276 2s22p2(3P) 3p(2S0) 2 11.6026 2s22p3(4S0) 3s(3S0) 3 9.52133
7 n=8 128 13.3927 2s22p2(3P) 3p(4D0) 20 11.7583 2s22p3(4S0) 3p(5P) 15 10.7402
8 n=9 162 13.4373 2s22p2(3P) 3p(4P0) 12 11.8416 2s22p3(4S0) 3p(3P) 9 10.9888
9 n = 10 200 13.4692 2s22p2(3P) 3p(4S0) 4 11.9955 2s22p3(4S0) 4s(5S0) 5 11.8376
10 n = 11 242 13.4928 2s22p2(3P) 3p(2D0) 10 12.0058 2s22p3(4S0) 4s(3S0) 3 11.9304
I 13.6053 14.5489 13.6180

Ionization potential (in eV) is also reported.


G. Colonna, M. Capitelli / Spectrochimica Acta Part B 64 (2009) 863–873 869

To validate the previous theoretical approach, exact and approx-


imate error values have been calculated considering 1st and 2nd order
approximations. For the H atom partition function this analysis is
shown in Fig. 8, where the approximate relative error of the partition
function (see Eq. (35)) has been compared with the exact error
calculated with respect to the first and second order approximation,
considering a different number of levels. It can be easily demonstrated
that the second order error corresponds also to the difference
between exact and approximate first order error. This result
demonstrates that the general mathematical analysis described in
the previous chapter is accurate, slightly underestimating the first
order relative error. The convergence of the Taylor series is also
shown, being the error of the second order approximation much
smaller than the first order one. The first and second order relative
errors are, in all the cases, smaller than 2% and 0.3% respectively, also
for a small number of levels (N = 5). The error decreases as the
number of levels increases, because the variance of the level
distribution decreases as N increases (see Fig. 7). The same behaviors
can be observed in the H atom energy (see Fig. 9) and specific heat
(see Fig. 10), this last having the largest relative error, less than3% for
the first order and 1% for the second order for N = 5, rapidly
decreasing as N increases (see also Fig. 11).

4.2. Nitrogen and oxygen atoms

The energy levels and the relevant statistical weights of atomic


nitrogen and oxygen are much more complicated than the correspond-
ing atomic hydrogen system. We are using for these systems complete

Fig. 8. Relative error of the H partition function vs. temperature for different values of
the maximum value of the principal quantum number. Legend: ⁎ - calculated with
Eq. (35); 1 - exact error with the first order approximation; 2 - exact error with the
second order approximation.

and the mean square energy is


   
〈ε2 〉 1 N 2 1 2 1 N 2 2 1
2
= ∑ 2n 1− 2 = ∑ 2n 1− 2 + 4
IH G n=2 n G n=2 n n
ð46Þ
ε 2 N 1 ε N−2
=2 −1 + ∑ ≈2 −1 +
IH G n = 2 n2 IH GN

where the approximation, (more accurate as N increases) is obtained


substituting the last summation with the integral, giving for the level
dispersion
" #
2 σε2 〈ε2 〉−ε2 G N−2 ðN−1Þ2
δε = 2 = ≈ − : ð47Þ
ε ε2 ðG−N + 1Þ2 N G

In Fig. 7, some statistical quantities of the hydrogen atoms have been


reported. For N = 3, G = 26 and δ2ε ≈ 0.0081, for N = 4, G = 58 and
δ2ε ≈ 0.0066, and substituting in Eq. (35) the maximum relative error
δQ
is respectively 0.015 and 0.019. For a large number of levels, which
Q⋆
is the case of our interest, it is G∝N3 and, neglecting the smallest
terms, Eq. (46) can be approximated as
" #
2 G 3N2 1 3
δε ≈ 1− ≈ ≈ 3: ð48Þ
G2 2N3 G 2N

Fig. 9. Relative error of the H internal energy vs. temperature for different values of the
Comparing this result with that reported in Fig. 2, we can assert that maximum value of the principal quantum number. Legend: ⁎ - calculated with Eq. (35);
the relative error of the partition function, decreasing with the 1 - exact error with the first order approximation; 2 - exact error with the second order
number of levels if N is large enough, is limited to 2%. approximation.
870 G. Colonna, M. Capitelli / Spectrochimica Acta Part B 64 (2009) 863–873

Table 2
Number of levels and group data as a function of the cutoff for H (Ng1st = 1), N (Ng1st = 3)
g
and O (N1st = 5) atoms.

Δs Hydrogen Nitrogen Oxygen


cm− 1 meV N G ε [eV] N G ε [eV] N G ε [eV]
250 30.9961 20 5738 13.5152 413 51594 14.4602 94 22,940 13.4656
500 61.9921 14 2028 13.4308 256 18362 14.3766 83 22,660 13.4643
750 92.9882 12 1298 13.3747 190 9528 14.2878 79 22,564 13.4639
1000 123.984 10 768 13.2864 158 6954 14.2305 68 7844 13.3727

databases obtained in these last years by completing existing observed


levels reported in Moore's tables [20] (see also databases [24,32,33])
with extrapolation and interpolation equations based on Ritz and Ritz-
Rydberg procedures [13,14,34].
As a result a very large number of levels are presently available in
these databases, that, in the case of nitrogen, originates mainly from
the interaction of the core of nitrogen atom (i.e. the state 3P) and the
excited electron jumping on all levels with principal quantum number
n N 2. The first eleven levels of this database have been reported in
Table 1. We can easily recognize the ground state (4S), the two low
lying excited states(2D, 2P) originating from a rearrangement of p3
levels satisfying Pauli's principle. Then starts the very large number of
levels from the interaction of the more stable core with the excited
electron. Note also we are reporting in the database other ‘unusual’
levels such as 2s2p4 (i = 5 in Table 1) as well as other series coming
from other core configurations (not reported in Table 1). Our database
distinguishes the different configurations up to n = 20, using an
hydrogenic structure for levels with n N 20. In particular, as anticipated,
the total statistical weight for all excited states originating from the
interaction of the 3P core with ns, np, nd… excited electron is 2gcn2. It
is impossible to report all the levels and corresponding degeneracies
Fig. 10. Relative error of the H constant volume specific heat vs. temperature for different
inserted in our database. We report indeed in Table 2 the characteristic
values of the maximum value of the principal quantum number. Legend: ⁎ - calculated with of an hypothetic nitrogen level starting from row i = 3 in Table 1,
Eq. (35); 1 - exact error with the first order approximation; 2 - exact error with the second characterized by lumped G and E values, these values depending on
order approximation. the lowering of the ionization potential used to truncate the partition

Fig. 11. Maximum relative error of the H thermodynamic quantities as a function of the Fig. 12. Absolute value of the first order relative error of nitrogen thermodynamic
number of excited levels. Legend: ⁎ - calculated with Eq. (35); 1 - exact error with the functions versus temperature, calculated considering Δs = 500 cm− 1. The different
g
first order approximation; 2 - exact error with the second order approximation. curves refers to the values of δ1st.
G. Colonna, M. Capitelli / Spectrochimica Acta Part B 64 (2009) 863–873 871

Table 3
Maximum of 1st and 2nd order percentage error, as a function of the initial group level, for N and O thermodynamic quantities.
g
N1st Nitrogen Oxygen
1st order 2nd order 1st order 2nd order
Q U [eV] Cv Q U [eV] Cv Q U [eV] Cv Q U [eV] Cv
1 46 29 70 41 28 70 52 32 62 50 33 73
3 2.0 1.3 2.8 0.52 0.39 0.91 18 10 19 15 9.7 20
5 1.3 0.85 1.8 0.28 0.21 0.49 1.7 1.3 2.7 0.41 0.37 0.86
7 0.99 0.64 1.4 0.18 0.14 0.31 1.1 0.80 1.7 0.18 0.16 0.38
9 0.64 0.42 0.89 0.094 0.072 0.16 0.87 0.65 1.4 0.13 0.12 0.28

The maximum has been calculated with respect to temperature and number of levels.

g
function. As an example the lumped level which includes all nitrogen c) N1st = 10
levels from i = 3 to N = 413, the maximum level of our database
2:38396 3:57561 11:9955 ε10
compatible with a lowering of the ionization potential of 250 cm− 1, − − − −
Q = 4 + 10e T + 6e T + … + 4e T + G10 e T
has a total statistical weight G = 51594 and average energy
E = 14.4602 eV. Note that the G values dramatically decrease with ð51Þ
increasing Δs. In the same table we have also reported the corre- where Gi and εi are calculated starting the sum in Eqs. (18) and (19)
sponding values for atomic hydrogen; in this case N corresponds to from l = i
the principal quantum number of the last considered level.
Fig. 12 reports the first order relative error on the partition N
function, internal energy and specific heat of nitrogen system with a Gi = ∑ gl
l=i
parametric cutoff of 500 cm− 1, which approximately corresponds to : ð52Þ
an LTE (Local Thermodynamic Equilibrium) atmospheric nitrogen 1 N
εi = ∑ gε
plasma. The three curves, reported in the figure, correspond to the Gi l = i l l
level from which arises the lumped level. So
g
g we can expect that the calculation with N1st = 1 will give unsatisfactory
a) N1st = 1 means that we are calculating the partition function (and 2
results due to the insertion of the P states in the lumped level, while
related thermodynamic properties) of atomic nitrogen as g
N1st = 3 case should give satisfactory results as in the case of atomic
g
ε1 hydrogen. We also expect an increase of accuracy with N1st = 10. This is

Q = 4 + G1 e T ð49Þ indeed the case as can be understood by inspection of Table 3. We can
g
see that the N1st = 1 approximation gives not acceptable errors for all
g
g
b) N1st = 3 quantities in all the temperature range, while N1st = 3 approximation
g
gives errors below 1%. The case with N1st = 10 refines the approximation
2:38396 3:57561 ε3 g
− − − reducing the error by a factor 2 as compared with the N1st = 3 case.
Q = 4 + 10e T + 6e T + G3 e T ð50Þ These errors are practically independent of the lowering of the
ionization potential used to truncate the partition function as can be

Fig. 13. Exact values of nitrogen thermodynamic functions (Q, ε [eV], C = v) versus Fig. 14. Exact values of oxygen thermodynamic functions (Q, ε [eV], Cv) versus
temperature, calculated considering Δs = 500 cm− 1. On the right axis, 1st and 2nd order temperature, calculated considering Δs = 500 cm− 1. On the right axis, 1st and 2nd order
percent error have been reported. percent error have been reported.
872 G. Colonna, M. Capitelli / Spectrochimica Acta Part B 64 (2009) 863–873

Table 4 (ground state + lumped level describing excited states) can be used
Maximum of the relative error calculated with respect to temperature and number of with a fair amount of confidence. The maximum of first order error for
levels in the three-group model, for N and O thermodynamic quantities.
atomic hydrogen decrease by about 1% to about 0.1% for partition
Nitrogen Oxygen functions which increase the number of excited states from the
⁎ 1 2 ⁎ 1 2 principal quantum number n = 3 to n = 40. Similar deviations are
Q 3.1 3.0 0.49 2.1 2.2 0.68 observed for the internal energy and the specific heat. Second order
U [eV] 2.4 2.2 0.49 1.5 1.0 0.71 errors are always much lower than the corresponding first order
Cv 5.0 4.6 1.10 3.3 2.1 1.40 errors. The situation changes for atomic systems which present low
⁎: Approximate error (see Eqs. (31)–(41)); 1: first order error; 2: second order error. lying excited states either as a rearrangement of valence electrons (N,
O) or originating from L-S coupling in the ground state (O). These
appreciated in Table 3, where we have reported the maximum first levels should be taken into account during the reduction procedure.
and second order approximation errors found for the different Δ The final step consists in reducing these systems to a three level
values (from 250 to 1000 cm− 1). model, one characterizing the ground state, one the low lying excited
The situation for oxygen is different due to the level structure of states and the last one the lumped level accounting for the very large
this species, as can be observed from Table 1: in fact the ground state number of levels characterizing the given system. The advantage of
of atomic oxygen is split into three levels due to the L–S coupling, such a procedure is the possibility to introduce a self consistent
which have close energies. Then we have two low lying excited states thermodynamic model for extreme conditions characterized by large
(1D and 1S in Table 1) originating from the rearrangement of p4 levels. variations of electron densities i.e. with different cutoff values. In such
After that, the very large number of high lying levels originates from cases use of Eqs. (53) and (54), taken as an example, in combination
the interaction of the more stable oxygen core 4S with the excited with the parameters reported in Table 2 can strongly reduce the
electron jumping on ns,np,nd,nf… levels. A complete database for computational times allowing also a self consistent thermodynamic
atomic oxygen including observed and unobserved levels has been model in the fluidynamics equations. Future work in this direction
compiled by our group (see Ref. [14]) and directly used in different should try to reduce the existing databases to three level systems as
papers [15,29,34]. This database has been reduced in the form of well as to complete existing databases on observed levels (e.g. Moore
Table 2 for different cutoff values, to trasform the database in a single [20] , NIST [32], IAEA [33]) to transform them in few level systems.
lumped level with given G and ε for i N 5. This new database can be used with a fair amount of confidence for a
g
The results show that the selection of N1st should take into account self-consistent description of thermodynamic and spectroscopical
as separated levels those near to the ground state and the two low properties of LTE plasmas avoiding the complications of calculating
lying excited states. Inspection of Table 3 explains this point. We can in the electronic partition functions through a level to level approach.
g
fact appreciate that the choice of N1st = 1 as expected, gives strong
g
errors. The errors persist up to N1st = 5 when maximum first and
second errors become really acceptable in all the temperature and Acknowledgements
cutoff ranges. These results demonstrate that partition function and
thermodynamic properties of atomic species can be reduced to a few This work has been partially supported by MIUR under contract
level system grouping all the high lying excited states in a single PRIN 2007H9S8SW_003 project and Progetto Strategico Regione
lumped level. Puglia PS-136 “Sviluppo di un rivelatore a film di diamante per
A further reduction can be operated by grouping also the low lying radiazione ultravioletta".
excited states thus reducing nitrogen and oxygen to a three level
system. We can in fact write the three level partition function as References

2:83083 ε3 [1] W. Dappen, A. Nayfonov, The sun as an equation-of-state laboratory, Astrophys. J.


− −
Q N = 4 + 16e T + G3 e T ð53Þ Suppl. Ser. 127 (2000) 287–292.
[2] T. Nishikawa, Simple expression of occupation probability in the “chemical
picture”, Astrophys. J. 532 (1, Part 1) (2000) 670–672.
0:00967077 2:33774 ε5 [3] R. Trampedach, W. Dappen, V.A. Baturin, A synoptic comparison of the mihalas–
− − −
Q O = 9e T + 6e T + G5 e T ð54Þ hummer–dappen and opal equations of state, Astrophys. J. 646 (1, Part 1) (2006)
560–578.
[4] W. Dappen, The equation of state for the solar interior, J. Phys., A, Math. Gen. 39
respectively for nitrogen and oxygen, where G and E values are those (2006) 4441–4446.
reported in Table 2 and T is expressed in eV. First and second order [5] M. Capitelli, F. Capitelli, A. Eletskii, Non-equilibrium and equilibrium problems in
laser-induced plasmas, Spectrochim. Acta Part B 55 (2000) 559–574.
errors resulting from this further approximation have been reported [6] J. Bauchea, C. Bauche-Arnoulta, O. Peyrusseb, Effective temperatures in hot dense
in Figs. 13 and 14 for nitrogen and oxygen systems for Δs = 500 cm− 1 plasmas, J. Quant. Spectros. Radiat. Transfer 99 (2006) 55–66.
together with the absolute values of the relevant quantities calculated [7] S. Hansena, K. Fourniera, C. Bauche-Arnoultb, J. Baucheb, O. Peyrussec, A
comparison of detailed level and superconfiguration models of neon, J. Quant.
by the complete set of energy levels. Table 4, on the other hand, Spectrosc. Radiat. Transfer 99 (2006) 272–282.
reports the maximum first and second error for all the reported cutoff [8] P. Yaroshchyk, D. Body, R.J.S. Morrison, B.L. Chadwick, A semi-quantitative
values, the first order error (column ⁎ in Table 4) being also estimated standard-less analysis method for laser-induced breakdown spectroscopy,
Spectrochim. Acta Part B 61 (2006) 200–209.
by Eqs. (31)–(41). All the results indicate that the use of Eqs. (53) and [9] G.C.-Y. Chan, G.M. Hieftje, Investigation of plasma-related matrix effects in
(54) in calculating partition function for atomic nitrogen and oxygen inductively coupled plasma-atomic emission spectrometry caused by matrices
does not affect the accuracy of the proposed method to reduce the with low second ionization potentials-identification of the secondary factor,
Spectrochim. Acta Part B 61 (2006) 642–659.
very large number of levels to a few level system.
[10] P.S. Dalyander, I.B. Gornushkin, D.W. Hahn, Numerical simulation of laser-induced
breakdown spectroscopy: modeling of aerosol analysis with finite diffusion and
5. Conclusions vaporization effects, Spectrochim. Acta Part B 63 (2008) 293–304.
[11] M. Capitelli, G. Ferraro, Cut-off criteria of electronic partition functions: effects on
spectroscopi quantities, Spectrochim. Acta Part B 31 (1976) 323–326.
We have reported in the previous pages an approximate method to [12] D. Landau, E. Lifshitz, Statistical Physics, Pergamon Press, Oxford, 1986.
calculate the partition function of atomic systems by using a few level [13] M. Capitelli, E. Molinari, Problems of determination of high-temperature thermo-
model to describe the very large number of excited levels entering in dynamic properties of rare gases with application to mixtures, J. Plasma Phys. 4
(1970) 335–355.
the partition function of high temperature plasmas. In particular we [14] M. Capitelli, G. Colonna, D. Giordano, L. Marraffa, A. Casavola, P. Minelli, L.D.
can distinguish hydrogen-like systems where a two level model Pietanza, F. Taccogna, Tables of internal partition functions and thermodynamic
G. Colonna, M. Capitelli / Spectrochimica Acta Part B 64 (2009) 863–873 873

properties of high–temperature mars–atmosphere species from 50 k to 50000 k, [25] O. Cardona, E. Simonneau, L. Crivellari, Analytic partition function for plasmas, Rev.
Esa str-246, isbn: 92–9092-349–0, European Space Agency (2005). Mex. Fís. 51 (2005) 476–481.
[15] A. Casavola, G. Colonna, M. Capitelli, Modeling laser-induced plasma expansion under [26] M.R. Zaghloul, Reduced formulation and efficient algorithm for the determination
equilibrium conditions, J. Thermophys. Heat Transfer 22 (2008) 407–413. of equilibrium composition and partition functions of ideal and nonideal complex
[16] M. Capitelli, A. Casavola, G. Colonna, A.D. Giacomoa, Laser-induced plasma expansion: plasma mixtures, Phys. Rev. E 69 (2004) 026702.
theoretical and experimental aspects, Spectrochim. Acta Part B 59 (2004) 271–289. [27] G. Colonna, A. D'Angola, A hierarchical approach for fast and accurate equilibrium
[17] H.R. Griem, High-density corrections in plasma spectroscopy, Phys. Rev. 128 (1962) calculation, Comput. Phys. Commun. 163 (2004) 177–190.
997–1003. [28] G. Colonna, Improvements of hierarchical algorithm for equilibrium calculation,
[18] H.R. Griem, Principles of Plasma Spectroscopy, Cambridge University Press, Cam- Comput. Phys. Commun. 177 (2007) 493–499.
bridge, 1997. [29] A. D'Angola, G. Colonna, C. Gorse, M. Capitelli, Thermodynamic and transport
[19] A. Bar-Shalom, J. Oreg, W.H. Goldstein, D. Shvarts, A. Zigler, Super-transition- properties in equilibrium air plasmas in a wide pressure and temperature range,
arrays: a model for the spectral analysis of hot, dense plasma, Phys. Rev. A 40 Eur. Phys. J. D 46 (2008) 129–150, doi:10.1140/epjd/e2007-00305-4.
(1989) 3183–3193. [30] J. Aubreton, M.F. Elchinger, V.R.P. Fauchais, P. André, Thermodynamic and
[20] C.E. Moore, Selected tables of atomic spectra, National Bureau of Standards, Jun. transport properties of a ternary ar–h2–he mixture out of equilibrium up to
1949 Nbs-467. 30 000k at atmospheric pressure, J. Phys. D Appl. Phys. 37 (2004) 2232–2246.
[21] K.M. Roussel, R.F. O'Connell, Variational solution of schrödinger equation for the [31] M. Capitelli, E. Ficocelli, The contribution of electronic excitation to the total
static screened coulomb potential, Phys. Rev. A 9 (1974) 52–56. specific heats of high-temperature gases: a misinterpreted absence, J. Plasma Phys.
[22] Y.P. Varshni, Scaling relation for the energy levels of a hydrogen atom at high 5 (1971) 115–121.
pressures, Z. Naturforsch. 57a (2002) 915–918. [32] http://www.nist.gov/srd/index.htm[online] (2009).
[23] M. Capitelli, D. Giordano, G. Colonna, The role of debye-hückel electronic energy [33] http://www-amdis.iaea.org/aladdin/[online] (2009).
levels on the thermodynamic properties of hydrogen plasmas including isentropic [34] M. Capitelli, G. Colonna, D. Giordano, L. Marraffa, A. Casavola, P. Minelli, D. Pagano,
coefficients, Phys. Plasmas 15 (2008) 082115. L. Pietanza, F. Taccogna, High-temperature thermodynamic properties of mars-
[24] M.W. Chase Jr., NIST-JANAF Thermochemical Tables, American Institue of Physics, atmosphere components, J. Spacecr. Rockets 42 (2005) 980–989.
Woodbury, New York, 1998.

Das könnte Ihnen auch gefallen