research article properties of stark resonant states in ...4. stark resonances and their properties...

12
Research Article Properties of Stark Resonant States in Exactly Solvable Systems Jeffrey M. Brown and Miroslav Kolesik College of Optical Sciences, University of Arizona, Tucson, AZ 85721, USA Correspondence should be addressed to Miroslav Kolesik; [email protected] Received 15 September 2015; Revised 23 November 2015; Accepted 29 November 2015 Academic Editor: Emmanuel Lorin Copyright Β© 2015 J. M. Brown and M. Kolesik. is is an open access article distributed under the Creative Commons Attribution License, which permits unrestricted use, distribution, and reproduction in any medium, provided the original work is properly cited. Properties of Stark resonant states are studied in two exactly solvable systems. ese resonances are shown to form a biorthogonal system with respect to a pairing defined by a contour integral that selects states with outgoing wave boundary conditions. Analytic expressions are derived for the pseudonorm, dipole moment, and coupling matrix elements which relate systems with different strengths of the external field. All results are based on explicit calculations made possible by a newly designed integration method for combinations of Airy functions representing resonant eigenstates. Generalizations for one-dimensional systems with short- range potentials are presented, and relations are identified which are likely to hold in systems with three spatial dimensions. 1. Introduction Resonance states have been used to solve a wide range of problems in the fields of nuclear physics [1, 2], quantum chemistry [3], nonlinear optics [4–6], and semiconductor physics [7, 8]. Despite their widely recognized utility, rel- atively little is known about their general properties since they do not live in the familiar Hilbert space associated with Hermitian quantum mechanics [9]. e properties and issues that are less well-understood than in Hermitian quantum mechanics include inner products, normalization and com- pleteness [10–17], complex expectation values [18, 19], and their physical interpretation [20]. Despite the mathematical difficulties related to their applications, resonance states do contain valuable physical information and it is important to investigate systems that could provide some guidance. Here we add to the present understanding of resonance systems by analytically calculating a number of useful quan- tities for two exactly solvable quantum systems: the 1D Dirac- delta potential and 1D square-well models in the presence of a homogeneous field. Despite the latter model being a textbook example, and the former being studied and used in applications for decades (e.g., [21–24]), their resonances have so far been studied mainly with numerical tools [8, 25]. e quantities of interest, explicitly evaluated for the first time in this work, are the normalization factors, eigenvalue equations, dipole matrix elements, and off-diagonal transi- tion elements that characterize the dependence of resonant basis states on the external field. We also generalize our results to more complex models with piecewise constant potentials, and for general one- dimensional systems with finite-range potentials. We identify relations between the generalized dipole moments and the gradient of the atomic potentials, which resemble similar properties in systems with self-adjoint Hamiltonians. Last but not least, all of our new results are based on direct evaluation of integral expressions, for which we have developed a new integration technique that is applicable to functions representing Stark resonances in one dimension with a piecewise constant potential. Beyond developing a deeper understanding of exactly solvable systems, the additional motivation for this work is in the use of resonance states as a basis for time-dependent SchrΒ¨ odinger evolution, with applications in modeling elec- tron ionization and nonlinear polarization due to a time varying optical pulse field [5]. Detailed study of exactly solvable systems with Stark resonant states brings multiple benefits. First, having explicit expressions for complex-valued observables and the ability to study their field and time dependence gives intuition of how one maps these complex values and open-system dynamics back to the real expec- tation values and the norm-preserving evolution found in Hindawi Publishing Corporation Advances in Mathematical Physics Volume 2015, Article ID 125832, 11 pages http://dx.doi.org/10.1155/2015/125832

Upload: others

Post on 26-Jan-2021

1 views

Category:

Documents


0 download

TRANSCRIPT

  • Research ArticleProperties of Stark Resonant States in Exactly Solvable Systems

    Jeffrey M. Brown and Miroslav Kolesik

    College of Optical Sciences, University of Arizona, Tucson, AZ 85721, USA

    Correspondence should be addressed to Miroslav Kolesik; [email protected]

    Received 15 September 2015; Revised 23 November 2015; Accepted 29 November 2015

    Academic Editor: Emmanuel Lorin

    Copyright Β© 2015 J. M. Brown and M. Kolesik. This is an open access article distributed under the Creative Commons AttributionLicense, which permits unrestricted use, distribution, and reproduction in any medium, provided the original work is properlycited.

    Properties of Stark resonant states are studied in two exactly solvable systems. These resonances are shown to form a biorthogonalsystem with respect to a pairing defined by a contour integral that selects states with outgoing wave boundary conditions. Analyticexpressions are derived for the pseudonorm, dipole moment, and coupling matrix elements which relate systems with differentstrengths of the external field. All results are based on explicit calculations made possible by a newly designed integration methodfor combinations of Airy functions representing resonant eigenstates. Generalizations for one-dimensional systems with short-range potentials are presented, and relations are identified which are likely to hold in systems with three spatial dimensions.

    1. Introduction

    Resonance states have been used to solve a wide range ofproblems in the fields of nuclear physics [1, 2], quantumchemistry [3], nonlinear optics [4–6], and semiconductorphysics [7, 8]. Despite their widely recognized utility, rel-atively little is known about their general properties sincethey do not live in the familiar Hilbert space associated withHermitian quantummechanics [9].The properties and issuesthat are less well-understood than in Hermitian quantummechanics include inner products, normalization and com-pleteness [10–17], complex expectation values [18, 19], andtheir physical interpretation [20]. Despite the mathematicaldifficulties related to their applications, resonance states docontain valuable physical information and it is important toinvestigate systems that could provide some guidance.

    Here we add to the present understanding of resonancesystems by analytically calculating a number of useful quan-tities for two exactly solvable quantum systems: the 1DDirac-delta potential and 1D square-well models in the presenceof a homogeneous field. Despite the latter model being atextbook example, and the former being studied and usedin applications for decades (e.g., [21–24]), their resonanceshave so far been studied mainly with numerical tools [8, 25].The quantities of interest, explicitly evaluated for the firsttime in this work, are the normalization factors, eigenvalue

    equations, dipole matrix elements, and off-diagonal transi-tion elements that characterize the dependence of resonantbasis states on the external field.

    We also generalize our results to more complex modelswith piecewise constant potentials, and for general one-dimensional systems with finite-range potentials.We identifyrelations between the generalized dipole moments and thegradient of the atomic potentials, which resemble similarproperties in systems with self-adjoint Hamiltonians.

    Last but not least, all of our new results are based ondirect evaluation of integral expressions, for which we havedeveloped a new integration technique that is applicable tofunctions representing Stark resonances in one dimensionwith a piecewise constant potential.

    Beyond developing a deeper understanding of exactlysolvable systems, the additional motivation for this work isin the use of resonance states as a basis for time-dependentSchrödinger evolution, with applications in modeling elec-tron ionization and nonlinear polarization due to a timevarying optical pulse field [5]. Detailed study of exactlysolvable systems with Stark resonant states brings multiplebenefits. First, having explicit expressions for complex-valuedobservables and the ability to study their field and timedependence gives intuition of how one maps these complexvalues and open-system dynamics back to the real expec-tation values and the norm-preserving evolution found in

    Hindawi Publishing CorporationAdvances in Mathematical PhysicsVolume 2015, Article ID 125832, 11 pageshttp://dx.doi.org/10.1155/2015/125832

  • 2 Advances in Mathematical Physics

    Hermitian quantum mechanics of a closed system. Such aconnection is crucial for applications in nonlinear optics(e.g., [5, 26, 27]). Secondly, the ability to compare differentresonance systemsmay indicate which properties or relationsare universally valid or common to all resonance systems.For example, we have witnessed in numerical simulations afield-dependent relation connecting the expectation valuesof the gradient of the atomic potential to the resonant statepseudonorm for one- and three-dimensional systems.

    2. Non-Hermitian Hamiltonians

    In this section, we give some background of the class ofHamiltonians that we want to investigate. We begin with a1D Hamiltonian that is parameterized by the strength of theexternal field 𝐹:

    𝐻(𝐹) = βˆ’1

    2βˆ‡2+ 𝑉 (π‘₯) βˆ’ π‘₯𝐹, (1)

    where the function 𝑉(π‘₯) represents the atomic potential. Tostudy the Stark resonances, one usually assumes outgoingwave boundary conditions at π‘₯ β†’ ∞ and seeks solutionsof π»πœ“

    π‘˜= πΈπ‘˜πœ“π‘˜where 𝐸

    π‘˜is the eigenvalue. With outgoing

    boundary conditions, the system is open as the particlecan escape toward π‘₯ β†’ ∞ and the operator 𝐻 is non-Hermitian. Therefore, energy 𝐸

    π‘˜, along with many other

    observables, is complex-valued.Without Hermiticity, we losemany of the guarantees of Hermitian quantum mechanics,such as conservation of the number of particles, real-valuedobservables, and square integrable wave functions. There isnot yet a full consensus on how to handle and interpretmany of these quantities, including normalization and innerproducts.

    Due to the non-square-integrable character of wavefunctions πœ“

    π‘˜, the standard inner product and normalization

    prescriptions do not apply, since the integrals normally usedto calculate them are divergent. Some regularization methodmust be used, and a number of approaches can be found in theliterature [10, 13, 28]. However, it is important to appreciatethat there may not be as much choice as it may seem inhow the Stark resonant states should be normalized. Forexample, if a resonant state expansion of Green’s operatorexists, the eigenstates appear in it with a definite β€œnorm” [29].In what follows we utilize biorthogonality of the Stark reso-nant system and obtain the eigenstates with such preferrednormalization factors.

    We consider theHamiltonian𝐻 to act on functions livingon a complex contour C, where the contour follows the realaxis in the vicinity of the atom and then deviates from the realaxis far from the origin. To select the space of outgoing wavefunctions, the contour departs into the upper complex planeas π‘₯ β†’ ∞. The shape of this contour is inconsequential,except its property that it approaches infinity in the sectorof the complex plain in which all outgoing waves, and inparticular the resonance states, decay exponentially. Onepossible example utilizing a piecewise linear path is shownin Figure 1. At the far end of the contour, outgoing wavefunctions that behave as βˆΌπ‘’π‘–π‘˜π‘₯ β†’ π‘’βˆ’π‘˜Im{𝑧} (with positive π‘˜)

    a

    I{z}

    R{z}Θ

    Figure 1: An example contour in the complex plane that servesas a β€œcomplexified” spatial axis in a model of an open quantumsystem in which a non-Hermitian Hamiltonian (1) acts in the spaceof functions defined along the contour. Both smooth and piecewiselinear contours are admissible for our purposes. In this example,the domain of the Hamiltonian would be specified by requiring that𝑓(π‘Žβˆ’) = 𝑓(π‘Ž

    +) and that an analogue of Cauchy-Riemann condition,

    𝑓(π‘Žβˆ’) = π‘’βˆ’π‘–Ξ˜

    𝑓(π‘Ž+), is satisfied for derivatives along the contour for

    all 𝑓 ∈ 𝐷(𝐻). We also assume that the potential𝑉(π‘₯) has a compactsupportwith a β€œradius” smaller than π‘Ž, so that nonanalytic potentialscan be considered.

    decay exponentially. Thus, the introduction of the contour isin the spirit of the external complex scaling.

    The differential expression𝐻 acting in the space of func-tions defined along the contour results in a non-Hermitianoperator that represents an open system. A 𝑐-product definedas a contour integral

    βŸ¨πœ™ | πœ“βŸ©def= ∫

    C

    πœ™ (𝑧) πœ“ (𝑧) d𝑧 (2)

    is the tool that replaces the standard scalar product inworking with non-self-adjoint operators. A formal argumentcan bemade that if two resonant states belong to two differenteigenvalues, then they are orthogonal with respect to theabove 𝑐-product [30], and the latter can serve as a definitionfor a pseudonorm in resonant states.

    In this work we aim to avoid any reliance on formal oper-ator properties. Instead, we show by explicit calculation of theunderlying contour integrals that the following orthogonalityrelation holds for outgoing Stark resonances:

    βŸ¨πœ“π‘›| πœ“π‘˜βŸ©

    def= ∫

    C

    πœ“π‘›(𝑧, 𝐹) πœ“

    π‘˜(𝑧, 𝐹) d𝑧 = 𝑁2

    𝑛(𝐹) π›Ώπ‘›π‘˜. (3)

    In particular, we verify the orthogonality of different func-tions (𝑛 ΜΈ= π‘˜), and we evaluate the normalization factor𝑁

    𝑛(𝐹)

    explicitly for two exactly solvable models.

    3. Motivation

    We have recently presented a proof of principle for appli-cation of metastable electronic states to calculate nonlinearresponse in a time-dependent field of an optical pulse [5].We have also demonstrated that the resulting description isextremely efficient in that a single Stark resonance is sufficientto obtain quite accurate nonlinear polarization for realisticmodels of atoms [31]. Here we recap some of the ideas behindtheMetastable Electronic State Approach (MESA) in order toidentify the relevant properties of resonant states.

  • Advances in Mathematical Physics 3

    If we represent a particular quantum stateΞ¨ as a sum overresonance states, plus a β€œbackground,”

    Ξ¨ (π‘₯, 𝑑) = βˆ‘

    𝑛

    𝑐𝑛(𝑑) πœ“π‘›(π‘₯, 𝐹 (𝑑)) + πœ“

    𝐡(π‘₯, 𝑑) , (4)

    and ask what is its evolution due to the time-dependent field𝐹(𝑑), then we can find a system of equations describing theevolution of coefficients 𝑐

    𝑛with the help of orthogonality

    relation (3):

    𝑐

    𝑛(𝑑) = βˆ’π‘–π‘

    𝑛(𝑑) 𝐸𝑛(𝐹 (𝑑)) + βˆ‘

    π‘˜

    π‘π‘˜(𝑑) 𝐹(𝑑) βŸ¨πœ•πΉπœ“π‘›| πœ“π‘˜βŸ© , (5)

    where𝐹(𝑑) is the time derivative of the electric field intensity.Here we have assumed that the expansion statesπœ“

    𝑛are slaved

    to the time-dependent field and are normalized to unity,βŸ¨πœ“π‘š(𝐹(𝑑)) | πœ“

    𝑛(𝐹(𝑑))⟩ = 𝛿

    π‘šπ‘›, at all times. Here we neglect

    the coupling to the β€œbackground” πœ“π΅which originates from

    the continuum contribution contained in various resonantstate expansions. MESA works with a physically motivatedassumption (see [5] for details) that, for systems initiallyin the ground state, the temporal decay of resonant statesrepresents ionization and that the flow of probability out ofthe space spanned by {πœ“

    π‘˜} manifests as the increase of the

    norm β€–πœ“π΅β€–.

    For the purposes of this work, (5) identifies the quan-tities that we aim to calculate. First, we need the complex-valued energies 𝐸

    𝑛. Second, we require normalization and

    orthogonality relation (3) to be satisfied by all resonancestates. To evaluate the induced polarization, we also need thegeneralized dipole moments

    π‘‘π‘›π‘˜

    = βŸ¨πœ“π‘›| π‘₯ | πœ“

    π‘˜βŸ© = ∫

    C

    πœ“π‘›(𝑧, 𝐹) π‘§πœ“

    π‘˜(𝑧, 𝐹) d𝑧, (6)

    and lastly we must calculate the coupling terms βŸ¨πœ•πΉπœ“π‘›| πœ“π‘˜βŸ©,

    which describe the change of the resonant state basis as itevolves slaved to the external field.

    The coupling terms can be related to the dipole momentmatrix elements with the help of the following argumentutilizing the parametric dependence of the Hamiltonian on𝐹 [32, 33]:

    πœ•πΉβŸ¨πœ“π‘›| 𝐻 (𝐹) | πœ“

    π‘˜βŸ© = (πœ•

    𝐹𝐸) π›Ώπ‘›π‘˜

    = βˆ’ (πΈπ‘›βˆ’ πΈπ‘˜) βŸ¨πœ•πΉπœ“π‘›| πœ“π‘˜βŸ©

    βˆ’ βŸ¨πœ“π‘›| π‘₯ | πœ“

    π‘˜βŸ© .

    (7)

    Moreover, for the normalized resonances the couplingterms are antisymmetric in indices 𝑛, π‘˜, since πœ•

    πΉπ›Ώπ‘›π‘˜

    = πœ•πΉβŸ¨πœ“π‘›|

    πœ“π‘˜βŸ© = βŸ¨πœ•

    πΉπœ“π‘›| πœ“π‘˜βŸ©+βŸ¨πœ“π‘›| πœ•πΉπœ“π‘˜βŸ© = 0.Thismeans that we have

    βŸ¨πœ•πΉπœ“π‘›| πœ“π‘˜βŸ© = βˆ’βŸ¨πœ“

    𝑛| πœ•πΉπœ“π‘˜βŸ© and for 𝑛 = π‘˜ the self-coupling

    vanishes βŸ¨πœ•πΉπœ“π‘˜| πœ“π‘˜βŸ© = 0 as a consequence of the 𝑐-product

    symmetry.Thus, the evolution system (5) can be alternativelywritten with the substitution

    βŸ¨πœ•πΉπœ“π‘›| πœ“π‘˜βŸ© =

    {{

    {{

    {

    0 𝑛 = π‘˜

    βˆ’βŸ¨πœ“π‘›| π‘₯ | πœ“

    π‘˜βŸ©

    (πΈπ‘›βˆ’ πΈπ‘˜)

    𝑛 ΜΈ= π‘˜.

    (8)

    In the appendix, we present a new integration technique thatallows calculation of the coupling terms directly, withoutreliance on the formal derivation underlying relation (8).

    4. Stark Resonances and Their Properties

    In this section we outline the properties of resonant wavefunctions for two different systems, and we find associatedresonant eigenvalue equations. We also calculate explicitnormalization factors for all Stark resonances using theorthogonality relation (3). Having established these tools, wecontinue to calculate dipole moment matrix elements andcoupling factors.

    In this work we assume that the potential 𝑉(π‘₯) has acompact support contained in (βˆ’π‘‘, +𝑑). Thus, the asymptoticform of both the conventional and resonant wave functionscan be obtained as a combination of Airy functions, 𝑐

    1𝐴𝑖 +

    𝑐2𝐡𝑖. The requirement that the solutions are regular for π‘₯ β†’

    βˆ’βˆž dictates 𝑐2= 0. For π‘₯ > 0 one can use combinations

    𝐢𝑖±(π‘₯) = 𝐡𝑖(π‘₯) Β± 𝑖𝐴𝑖(π‘₯) which behave as outgoing (+)

    and incoming (βˆ’) waves at π‘₯ β†’ ∞. Representation of theeigenstates of the original𝐻 (i.e., the one that acts on the realaxis) which is particularly suitable for our purposes can bewritten as

    πœ“πΈ(π‘₯) =

    βˆ’2𝐴𝑖 [𝛼 (π‘₯ + 𝐸/𝐹)]

    π‘ˆβˆšπ·+ (𝐸)π·βˆ’(𝐸)

    π‘₯ < βˆ’π‘‘,

    πœ“πΈ(π‘₯) =

    𝑖

    π‘ˆ

    βˆšπ·βˆ’(𝐸)

    𝐷+ (𝐸)𝐢𝑖+[𝛼 (π‘₯ +

    𝐸

    𝐹)]

    βˆ’π‘–

    π‘ˆ

    √𝐷+(𝐸)

    π·βˆ’ (𝐸)πΆπ‘–βˆ’[𝛼 (π‘₯ +

    𝐸

    𝐹)]

    π‘₯ > +𝑑,

    (9)

    where 𝛼 = βˆ’(2𝐹)1/3 and 𝐷±(𝐸) are sought expressionsrepresenting the eigenvalue equations for outgoing (+) andincoming (βˆ’) wave functions. The fact that the originaloperator (i.e., the one acting on the real axis) is self-adjointguarantees that the above states can be normalized to a deltafunction in energy:

    ∫

    ∞

    βˆ’βˆž

    πœ“βˆ—

    𝐸(π‘₯) πœ“πΈ (π‘₯) dπ‘₯ = 𝛿 (𝐸 βˆ’ 𝐸) . (10)

    It is sufficient to examine the asymptotic behavior of thesestates to verify that this normalization is obtained with π‘ˆ =22/3

    𝐹1/6. While a particular normalization is not crucial for

    us, the above form of eigenstates allows us to infer the formof the resonant functions sought below.

    To obtain the remaining portion(s) of energy eigenstates,one has to β€œfill in” the wave function in the central regionof βˆ’π‘‘ to 𝑑 and in doing so satisfy whatever conditions agiven potential imposes on them. In both cases treated in thiswork, this means to find functions that are continuous and torequire continuity of derivatives in the square-well case and aβ€œcusp condition” (13) in the Dirac-delta case. This procedurereveals the concrete form of expressions 𝐷±(𝐸) for a given𝑉(π‘₯).

  • 4 Advances in Mathematical Physics

    The asymptotic form of the energy eigenstates as shownin (9) indicates the form of the resonance wave functions. Forexample, if we find a complex-energy root of 𝐷+(𝐸) = 0 theincoming part of the wave function πΆπ‘–βˆ’ will be eliminated.At the same time a pole will appear in the projection ontothe outgoing wave function. This tells us that the resonancebehaves as 𝐴𝑖 and 𝐢𝑖+ for large negative and positive π‘₯,respectively.

    We present our final results, including those on resonantstate normalization, in the form that is independent of howone chooses to parameterize the eigenfunctions. Wheneverwe show intermediate results, it is for the wave functionswritten as follows. For the Dirac-delta model, we take theunnormalized ansatz for the outgoing resonance in the form

    𝑔 (π‘₯, 𝐹) =

    {

    {

    {

    𝐢𝑖 (𝛼𝛽)𝐴𝑖 [𝛼 (π‘₯ + 𝛽)] π‘₯ < 0

    𝐴𝑖 (𝛼𝛽)𝐢𝑖 [𝛼 (π‘₯ + 𝛽)] π‘₯ > 0.

    (11)

    For the system with a square-well potential of width 2𝑑 anddepth 𝑉

    0we take

    𝑔 (π‘₯, 𝐹)

    =

    {{{{

    {{{{

    {

    πœ…0𝐴𝑖 [𝛼 (π‘₯ + 𝛽)] π‘₯ < βˆ’π‘‘

    πœ…1𝐴𝑖 [𝛼 (π‘₯ + 𝛽

    )] + πœ…

    2𝐡𝑖 [𝛼 (π‘₯ + 𝛽

    )] βˆ’π‘‘ < π‘₯ < 𝑑

    πœ…3𝐢𝑖 [𝛼 (π‘₯ + 𝛽)] π‘₯ > +𝑑

    (12)

    with coefficients πœ…π‘–to be fixed to ensure continuity of wave

    function value and derivative across well boundaries. Forthese functions to become resonant states the energies in 𝛽 =𝐸/𝐹 and 𝛽 = (𝐸 βˆ’ 𝑉

    0)/𝐹must be solutions to the eigenvalue

    equation(s) we present next.

    4.1. Eigenvalue Equations. For the Dirac-delta potential𝑉(π‘₯) = βˆ’π΅π›Ώ(π‘₯), where 𝐡 is the depth of the potential, and theeigenstate representations (9) are valid with 𝑑 = 0. The deltafunction potential imposes a boundary condition on thewavefunction’s value and derivative:

    π‘‘πœ“ (0+)

    𝑑π‘₯βˆ’π‘‘πœ“(0βˆ’)

    𝑑π‘₯= βˆ’2π΅πœ“ (0) . (13)

    This β€œcusp condition,” when applied to the above eigenstateparameterization, leads directly to thewell-known [4] expres-sion for the eigenvalue equation for resonant energies:

    𝐷±(𝐸) ≑ 1 βˆ’

    2πœ‹π΅

    (2𝐹)1/3

    𝐴𝑖(βˆ’2𝐸

    (2𝐹)2/3

    )𝐢𝑖±(

    βˆ’2𝐸

    (2𝐹)2/3

    ) . (14)

    Complex-valued solutions to 𝐷+(𝐸) = 0 determine theresonant energies of the outgoing Stark functions.

    Longer calculations are required to obtain the analogousequation [25, 34] for the square-well potential. One needs toconnect the outer regions with a linear combination of Airyfunctions (12) and eliminate the unknown coefficients. Theresult reads

    𝐷±(𝐸) ≑ (𝐴

    0𝐴

    1βˆ’ 𝐴

    0𝐴1) (𝐡2𝐢

    3βˆ’ 𝐡

    2𝐢3)

    βˆ’ (𝐴0𝐡

    1βˆ’ 𝐴

    0𝐡1) (𝐴2𝐢

    3βˆ’ 𝐴

    2𝐢3) ,

    (15)

    βˆ’0.5 0.50.0Ereal

    βˆ’0.6

    βˆ’0.4

    βˆ’0.2

    0.0

    0.2

    Eim

    ag

    (a)

    0.0 0.5βˆ’0.5Ereal

    βˆ’0.6

    βˆ’0.4

    βˆ’0.2

    0.0

    0.2

    Eim

    ag

    (b)

    Figure 2: Outgoing resonance eigenvalue equation landscapes forDirac-delta ((a) 𝐡 = 1) and square-well potential ((b) 𝑑 = 4 and𝑉0= βˆ’0.5) systems in the external field 𝐹 = 0.03. To visualize the

    locations of the energy eigenvalues, we evaluate (14) and (15) over arange of 𝐸 in the complex plane and convert |𝐷| to a height map viathe formula (1 βˆ’ (1+ |𝐷|0.3)βˆ’1 + πœ–)βˆ’1, so that its roots are representedby poles that are easy to locate (πœ– β‰ˆ 0.1).

    where we utilized shorthand notations to compress the other-wise long expression.𝐴,𝐡, and𝐢 stand for the correspondingAiry functions, and primes denote derivatives. For a well withdepth 𝑉

    0and half-width 𝑑, the subscripts indicate on which

    sides of the well walls the arguments of the functions areevaluated, with 0, 1, 2, 3 representing 𝛼(π‘₯ + 𝛽) at π‘₯ = βˆ’π‘‘ βˆ’ πœ–,βˆ’π‘‘+πœ–,+π‘‘βˆ’πœ–, andπ‘₯ = +𝑑+πœ–, respectively.The value of𝛽 is alsodependent on where the functions are evaluated. For regionsoutside of the well 𝛽 = 𝐸/𝐹 and for regions inside the well𝛽 = (𝐸 βˆ’ 𝑉

    0)/𝐹. Thus, the subscripts 0 and 1 represent Airy

    functions evaluated just outside and inside the left boundaryof the well, while subscripts 2 and 3 represent arguments atthe right boundary of the well.

    It is helpful to visualize the resonance energy β€œlandscapes”illustrated in Figure 2. Although the eigenvalue equation

  • Advances in Mathematical Physics 5

    for the square-well (15) is more complicated, it is apparentthat the two systems share some common properties. Themain differing feature is the possibility of multiple boundstates (highlighted in red) in the square-well potential, whilethe Dirac-delta system only supports a single bound state.However, there exist two other infinite families of resonances.The β€œright” family has eigenvalues located along the real axisand corresponds to longer living states, while the β€œleft” family,with energies along the ray arg(𝑧) = βˆ’2πœ‹/3, are fast decaying,short-lived states. This resonance structure is most likelya generic feature at least in case of short-ranged attractivepotentials.

    4.2. Stark State Normalization. To establish formulas fornormalization and to verify the orthogonality relation (3) forboth systems, we make use of the formula (VS 3.50) in Valleand Soares [35] giving a primitive (antiderivative) functionfor a square of arbitrary combination of Airy functions.

    To calculate the contour integral(s), the correspondingprimitive functions are evaluated at points of discontinuitiesof the potential and at both ends of the contour, and this iswhere the choice of the contour is important. The resonantwave functions decay exponentially for π‘₯ β†’ βˆ’βˆž, and thecorresponding boundary terms vanish. On the other side ofthe contour, the asymptotic behavior of the primitive functionis dominated by

    𝐢𝑖+[𝛼 (𝑧 + 𝛽)]

    β‰ˆ

    π‘’βˆ’π‘–πœ‹/4exp [+𝑖 (2/3) [βˆ’π›Ό (𝑧 + 𝛽)]3/2]

    βˆšπœ‹ [βˆ’π›Ό (𝑧 + 𝛽)]1/4

    ,

    (16)

    where 𝛼 = βˆ’(2𝐹)1/3, 𝛽 = 𝐸𝑛/𝐹, and 𝐸

    𝑛is a root of

    the eigenvalue equation. It is straightforward to verify thatasymptotically along the contour, where 𝑧 ∼ πœŒπ‘’π‘–Ξ˜, this func-tion decays for arbitrary fixed 𝛽 as 𝜌 β†’ ∞. Thus, theboundary terms brought by this end of the contour alsovanish. Moreover, since the integrands are in all cases entirefunctions (containing no singular points), the precise shapeof the integration path does not affect the outcome. As aresult, in any piecewise constant atomic potential 𝑉(π‘₯) itis only the special points of 𝑉(π‘₯) that give rise to nonzerocontribution(s).

    Thus, direct integration along the contour, followed bysimplifications making use of the eigenvalue equation andthe Wronskian for Airy functions, yields the following nor-malization factor for the Stark resonance in the Dirac-deltamodel:

    𝑁2=𝐴𝑖 (𝛼𝛽)𝐢𝑖

    (𝛼𝛽) + 𝐴𝑖

    (𝛼𝛽)𝐢𝑖 (𝛼𝛽)

    π›Όπœ‹

    =1

    π›Όπœ‹[πœ“(π‘₯ = 0

    +) + πœ“(π‘₯ = 0

    βˆ’)] .

    (17)

    To calculate the normalization factor for the square-wellsystem, a similar but more complicated procedure to evaluate

    (3) using formula (VS 3.50) results in a surprisingly simpleexpression for the normalization factor:

    𝑁2

    =𝑉0

    𝐹[(𝐴2𝐢

    3βˆ’ 𝐴

    2𝐢3)2

    𝐴2

    0βˆ’ (𝐴

    0𝐴1βˆ’ 𝐴0𝐴

    1)2

    𝐢2

    3]

    =𝑉0

    𝐹[πœ“ (π‘₯ = βˆ’π‘‘)

    2βˆ’ πœ“ (π‘₯ = 𝑑)

    2] .

    (18)

    To the authors’ knowledge, this is the first time these resultshave been presented in the explicit form.

    To verify the mutual orthogonality with respect to (3) fordifferent resonant states, we use formula (VS 3.53), togetherwith the fact that the complex energies satisfy the eigenvalueequation(s).

    To conclude this subsection, we note that our direct verifi-cation of the orthogonality relation (3) and the explicit calcu-lation of the corresponding normalization factors means thatthere exists an infinite dimensional space of functions thatcan be expressed as superpositions of Stark resonances. Inthe spirit of [29], one should ask if this gives us the preferredexpansion. Indeed, one can alternatively use the self-adjointHamiltonian eigenstates (9) and find the projector onto agiven resonant state as the residue of πœ“

    𝐸(π‘₯)πœ“πΈ(𝑦) at the pole

    in the complex plane that corresponds to its energy 𝐸 = 𝐸𝑛.

    While it is beyond the scope of the present paper, we notethat the two approaches in fact lead to the same expansioncoefficients.

    4.3. Dipole Matrix Elements. Next we calculate the gen-eralized dipole matrix elements, both diagonal and off-diagonal. Expressedwith the help of unnormalized resonanceeigenfunctions 𝑔[𝛼(π‘₯ + 𝛽

    π‘š)], the contour integrals we need

    to evaluate read

    βŸ¨πœ“π‘š| π‘₯ | πœ“

    π‘›βŸ©

    =1

    𝑁2𝑛

    ∫C

    𝑔 [𝛼 (π‘₯ + π›½π‘š)] 𝑧𝑔 [𝛼 (π‘₯ + 𝛽

    𝑛)] d𝑧.

    (19)

    Note that these quantities differ from their Hermitian coun-terparts as there is no complex conjugation in the integrand,and the result is complex-valued [19].Herewe assume that thecontourC is chosen such that it only starts to deviate from thereal axis forπ‘₯ > 𝑑, that is, outside of the potential support. Forboth the delta potential model and the square-well potentialsystem, we integrate over each distinct interval of constantpotential𝑉making use of the formula (VS 3.51). As the shapeof the contour ensures vanishing contributions from its ends,one only needs to evaluate the primitive functions at π‘₯ = 0for Dirac-delta and π‘₯ = ±𝑑 for the square-well system.

  • 6 Advances in Mathematical Physics

    For the diagonalmatrix element in the Dirac-delta poten-tial model, we obtain the following expression in terms ofAiry functions:

    βŸ¨πœ“π‘›| π‘₯ | πœ“

    π‘›βŸ©π·=

    1

    𝑁2𝑛

    ⋅𝐴𝑖 (𝛼𝛽

    𝑛) 𝐢𝑖 (𝛼𝛽

    𝑛)

    3𝛼2[𝐴𝑖(𝛼𝛽𝑛) 𝐢𝑖 (𝛼𝛽

    𝑛)

    βˆ’ 𝐴𝑖 (𝛼𝛽𝑛) 𝐢𝑖(𝛼𝛽𝑛)] +

    2𝐸

    3𝐹𝑁2𝑛

    β‹…1

    𝛼[(𝐴𝑖(𝛼𝛽𝑛) 𝐢𝑖 (𝛼𝛽

    𝑛))2

    βˆ’ (𝐴𝑖 (𝛼𝛽𝑛) 𝐢𝑖(𝛼𝛽𝑛))2

    ]⏟⏟⏟⏟⏟⏟⏟⏟⏟⏟⏟⏟⏟⏟⏟⏟⏟⏟⏟⏟⏟⏟⏟⏟⏟⏟⏟⏟⏟⏟⏟⏟⏟⏟⏟⏟⏟⏟⏟⏟⏟⏟⏟⏟⏟⏟⏟⏟⏟⏟⏟⏟⏟⏟⏟⏟⏟⏟⏟⏟⏟⏟⏟⏟⏟⏟⏟⏟⏟⏟⏟⏟⏟⏟⏟⏟⏟⏟⏟⏟⏟⏟⏟⏟⏟⏟⏟⏟⏟⏟⏟⏟⏟⏟⏟⏟⏟

    βˆ’π‘2

    𝑛

    .

    (20)

    These results can be further simplified using the formulafor the normalization factor𝑁2 together with the eigenvalueequation (14) and are found to be related to the change in theeigenvalue with respect to the field 𝐹 as follows:

    βŸ¨πœ“π‘›| π‘₯ | πœ“

    π‘›βŸ©π·=

    1

    6πœ‹2𝛼𝐡𝑁2𝑛

    βˆ’2𝐸

    3𝐹= βˆ’πœ•πΉπΈπ‘›, (21)

    where the second equality can be verified by differentiatingthe eigenvalue equation (14) with respect to 𝐹.

    For the off-diagonal dipole matrix elements, we use (VS3.54). Utilizing 𝐢

    𝑛,π‘˜= 𝐢𝑖(𝛼𝛽

    𝑛,π‘˜) and 𝐴

    𝑛,π‘˜= 𝐴𝑖(𝛼𝛽

    𝑛,π‘˜) to

    shorten the notation, we find that

    βŸ¨πœ“π‘›| π‘₯ | πœ“

    π‘˜βŸ©π·=

    1

    π‘π‘›π‘π‘˜

    2

    𝛼5 (π›½π‘›βˆ’ π›½π‘˜)3[πΆπ‘›πΆπ‘˜π΄

    π‘›π΄π‘˜

    βˆ’ πΆπ‘›πΆπ‘˜π΄π‘›π΄

    π‘˜βˆ’ π΄π‘›π΄π‘˜πΆ

    π‘›πΆπ‘˜+ π΄π‘›π΄π‘˜πΆπ‘›πΆ

    π‘˜]

    +1

    π‘π‘›π‘π‘˜

    2

    𝛼4 (π›½π‘›βˆ’ π›½π‘˜)2[πΆπ‘›πΆπ‘˜π΄

    𝑛𝐴

    π‘˜

    βˆ’ π΄π‘›π΄π‘˜πΆ

    𝑛𝐢

    π‘˜] .

    (22)

    Grouping terms in order to identify Wronskians allows us tosimplify the expression down to

    βŸ¨πœ“π‘›| π‘₯ | πœ“

    π‘˜βŸ©π·=

    𝐹

    2 (πΈπ‘›βˆ’ πΈπ‘˜)2[𝑁𝑛

    π‘π‘˜

    +π‘π‘˜

    𝑁𝑛

    ] , (23)

    which in turn can be written solely in terms of the wavefunction properties at π‘₯ = 0,

    βŸ¨πœ“π‘›| π‘₯ | πœ“

    π‘˜βŸ©π·

    =βˆ’πΉ

    𝛼 (πΈπ‘›βˆ’ πΈπ‘˜)2[πœ“

    𝑛(0βˆ’) πœ“

    π‘˜(0βˆ’) βˆ’ πœ“

    𝑛(0+) πœ“

    π‘˜(0+)] ,

    (24)

    with πœ“ standing for the derivative of the normalized Starkwave function.

    Let us proceed with the dipole moment calculations forthe square-well potential. The matrix elements can be foundby again making use of (VS 3.51) and (VS 3.54).Thanks to thecontinuity properties of the wave function, many terms that

    arise in the course of this calculation cancel, and the resultingdiagonal terms are

    βŸ¨πœ“π‘›| π‘₯ | πœ“

    π‘›βŸ©π‘†

    =𝑉0

    3𝐹[(𝑑 βˆ’ 4𝛽)πœ“ (βˆ’π‘‘)

    2+ (𝑑 + 4𝛽)πœ“ (𝑑)

    2]

    +2𝑉2

    0

    3𝐹2[πœ“ (βˆ’π‘‘)

    2βˆ’ πœ“ (𝑑)

    2]

    +2𝑉0

    3𝐹𝛼[πœ“(βˆ’π‘‘)2βˆ’ πœ“(𝑑)2] ,

    (25)

    which is simplified further using the normalization factor(again as with the delta model) to

    βŸ¨πœ“π‘›| π‘₯ | πœ“

    π‘›βŸ©π‘†=2𝑉0βˆ’ 4𝐸𝑛

    3𝐹

    +𝑉0𝑑

    3𝐹[πœ“ (βˆ’π‘‘)

    2+ πœ“ (𝑑)

    2]

    +2𝑉0

    3𝛼𝐹[πœ“(βˆ’π‘‘)2βˆ’ πœ“(𝑑)2] .

    (26)

    It can be checked numerically that the diagonal elements arealso equal to βˆ’πœ•

    𝐹𝐸𝑛, just as in the delta potential model.

    Now we calculate the off-diagonal elements. When using(VS 3.54), taking into account that continuity of the wavefunction allows us to ignore terms that do not involve π‘₯ andthe sum 𝛽

    𝑛+ π›½π‘˜, as a result every other term cancels, and the

    resulting equation is surprisingly simple:

    βŸ¨πœ“π‘›| π‘₯ | πœ“

    π‘˜βŸ©π‘†

    =𝑉0

    (πΈπ‘›βˆ’ πΈπ‘˜)2[πœ“π‘›(βˆ’π‘‘) πœ“

    π‘˜(βˆ’π‘‘) βˆ’ πœ“

    𝑛(𝑑) πœ“π‘˜(𝑑)] ,

    (27)

    and one should note the similarity with its counterpartformula for the Dirac-delta model. We thus arrive at theconclusion that all dipole matrix elements can be expressedin simple formulas which only depend on the values of thewave functions (and their derivatives) at the special pointsgiven that characterize the potential. It will be interesting tosee if these results can be generalized for arbitrary systemswith piecewise constant potentials.

    4.4. Coupling Matrix Elements. We now turn our attentionto the terms βŸ¨πœ•

    πΉπœ“π‘›| πœ“π‘˜βŸ© identified in (5). These quantities

    mediate the connections between Stark resonances of a givensystem at different values of the field 𝐹 and we call themaccordingly coupling terms. They are needed to describe theevolution of the system in a time-dependent 𝐹(𝑑). Instead oftrusting the formally derived relation (8), we compute thesequantities through direct integration. Thus, our result can bealso interpreted as a direct verification of (8) for two modelsystems. Moreover, the procedure and the particular repre-sentation of the results lead us to a generalization for arbitraryone-dimensional systems with short-range potentials.

  • Advances in Mathematical Physics 7

    To calculate the coupling terms, we first differentiate thewave function with respect to the field 𝐹 and then integrate

    ∫

    ∞

    βˆ’βˆž

    (πœ•πΉπœ“π‘›) πœ“π‘˜dπ‘₯ = 1

    π‘π‘›π‘π‘˜

    ∫

    ∞

    βˆ’βˆž

    (πœ•πΉπ‘”π‘›) π‘”π‘˜dπ‘₯

    βˆ’(πœ•πΉπ‘π‘›)

    𝑁𝑛

    ∫

    ∞

    βˆ’βˆž

    πœ“π‘›πœ“π‘˜dπ‘₯

    ⏟⏟⏟⏟⏟⏟⏟⏟⏟⏟⏟⏟⏟⏟⏟⏟⏟⏟⏟⏟⏟⏟⏟

    π›Ώπ‘›π‘˜

    .

    (28)

    It is clear that the diagonal (𝑛 = π‘˜) coupling terms betweennormalized states vanish, and we only need to calculate off-diagonal elements and therefore can disregard the secondterm, leaving only the integral βŸ¨πœ•

    𝐹𝑔𝑛| π‘”π‘˜βŸ©. For the model

    with Dirac-delta potential, the integral is

    βŸ¨πœ•πΉπ‘”π‘›| π‘”π‘˜βŸ©π·

    = Δ𝐢

    π‘›πΆπ‘˜βˆ«

    0

    βˆ’βˆž

    𝐴𝑖 [𝛼 (π‘₯ + 𝛽𝑛)] 𝐴𝑖 [𝛼 (π‘₯ + 𝛽

    π‘˜)] dπ‘₯

    + Δ𝐴

    π‘›π΄π‘˜βˆ«

    ∞

    0

    𝐢𝑖 [𝛼 (π‘₯ + 𝛽𝑛)] 𝐢𝑖 [𝛼 (π‘₯ + 𝛽

    π‘˜)] dπ‘₯

    + Ξ”πΆπ‘›πΆπ‘˜βˆ«

    0

    βˆ’βˆž

    𝐴𝑖[𝛼 (π‘₯ + 𝛽

    𝑛)] 𝐴𝑖 [𝛼 (π‘₯ + 𝛽

    π‘˜)] dπ‘₯

    + Ξ”π΄π‘›π΄π‘˜βˆ«

    ∞

    0

    𝐢𝑖[𝛼 (π‘₯ + 𝛽

    𝑛)] 𝐢𝑖 [𝛼 (π‘₯ + 𝛽

    π‘˜)] dπ‘₯

    βˆ’2

    3𝛼2

    β‹… πΆπ‘›πΆπ‘˜βˆ«

    0

    βˆ’βˆž

    π‘₯𝐴𝑖[𝛼 (π‘₯ + 𝛽

    𝑛)] 𝐴𝑖 [𝛼 (π‘₯ + 𝛽

    π‘˜)] dπ‘₯

    βˆ’2

    3𝛼2

    β‹… π΄π‘›π΄π‘˜βˆ«

    ∞

    0

    π‘₯𝐢𝑖[𝛼 (π‘₯ + 𝛽

    𝑛)] 𝐢𝑖 [𝛼 (π‘₯ + 𝛽

    π‘˜)] dπ‘₯,

    (29)

    whereΞ” = 1/(3πœ‹2𝛼3𝐡𝑁2), andπ΄π‘˜,π΅π‘˜,πΆπ‘˜stand for the corre-

    sponding Airy functions or their combinations evaluated atπ›Όπ›½π‘˜(i.e., at π‘₯ = 0).For the model with the square-well potential we have a

    combination of integrals of similar types; namely,

    βŸ¨πœ•πΉπ‘”π‘›| π‘”π‘˜βŸ©π‘†= ∫

    βˆ’π‘‘

    βˆ’βˆž

    πœ…

    0𝛾0𝐴𝑖 [𝛼 (π‘₯ + 𝛽

    𝑛)]

    β‹… 𝐴𝑖 [𝛼 (π‘₯ + π›½π‘˜)] dπ‘₯

    + ∫

    𝑑

    βˆ’π‘‘

    (πœ…

    1𝐴𝑖 [𝛼 (π‘₯ + 𝛽

    𝑛)] + πœ…

    2𝐡𝑖 [𝛼 (π‘₯ + 𝛽

    𝑛)])

    β‹… (𝛾1𝐴𝑖 [𝛼 (π‘₯ + 𝛽

    π‘˜)] + 𝛾2𝐡𝑖 [𝛼 (π‘₯ + 𝛽

    π‘˜)]) dπ‘₯

    + ∫

    ∞

    𝑑

    πœ…

    3𝛾3𝐢𝑖 [𝛼 (π‘₯ + 𝛽

    𝑛)] 𝐢𝑖 [𝛼 (π‘₯ + 𝛽

    π‘˜)] dπ‘₯

    +𝛼

    3𝐹∫

    ∞

    βˆ’βˆž

    (3𝐸

    𝑛+ π‘₯ βˆ’ 2𝛽

    𝑛) 𝑔[𝛼 (π‘₯ + 𝛽

    𝑛)]

    β‹… 𝑔 [𝛼 (π‘₯ + π›½π‘˜)] dπ‘₯ +

    2𝛼𝑉0

    3𝐹2∫

    𝑑

    βˆ’π‘‘

    𝑔[𝛼 (π‘₯ + 𝛽

    𝑛)]

    β‹… 𝑔 [𝛼 (π‘₯ + π›½π‘˜)] dπ‘₯,

    (30)

    where 𝛾 and πœ… are the coefficients that guarantee continuity of𝑔𝑛and 𝑔

    π‘˜, respectively, at the well boundaries. These factors

    are expressed in terms of Airy functions and depend on𝐹 through 𝛼, 𝛽

    𝑛, and 𝛽

    π‘˜. Their primed notations represent

    derivatives with respect to 𝐹.The above integrals for both models contain terms of

    two kinds. The first group can be evaluated making useof known, previously published Airy integrals. These havethe form βˆ«π‘

    π‘Žπ΄π΅ dπ‘₯, where 𝐴, 𝐡 are a linear combination

    of Airy functions with arguments 𝛼(π‘₯ + 𝛽), where 𝛽 isdifferent in 𝐴 and 𝐡. Thus, the first two lines in (29) and firstthree lines in (30) can be dealt with, although each integralrequires lengthy computations, especially in the square-wellcase. Then there are integrals of the types βˆ«π‘

    π‘Žπ΄π΅ dπ‘₯ and

    βˆ«π‘

    π‘Žπ‘₯𝐴𝐡 dπ‘₯. No known formulas are available for these, and

    we have developed a new integration technique that weoutline in the appendix. As a result, analytic results can beobtained for both systems.

    For the Dirac-delta model, the individual integrals on theRHS of (29) simplify pairwise. The first pair is calculatedusing (VS 3.53), while the last two pairs use the identitiesderived in the appendix. The intermediate expression reads

    βŸ¨πœ•πΉπ‘”π‘›| π‘”π‘˜βŸ©π·

    =Ξ”

    𝛼2 (π›½π‘›βˆ’ π›½π‘˜)[𝐴

    π‘›π΄π‘˜πΆπ‘›πΆ

    π‘˜βˆ’ 𝐴𝑛𝐴

    π‘˜πΆ

    π‘›πΆπ‘˜]

    βˆ’Ξ”

    𝛼2 (π›½π‘›βˆ’ π›½π‘˜)[𝐴

    𝑛𝐴

    π‘˜πΆπ‘›πΆπ‘˜βˆ’ π΄π‘›π΄π‘˜πΆ

    𝑛𝐢

    π‘˜]

    βˆ’2

    3πœ‹π›Ό5

    1

    (π›½π‘›βˆ’ π›½π‘˜)2[π›½π‘›πΆπ‘›π΄π‘›βˆ’ π›½π‘˜πΆπ‘˜π΄π‘˜]

    βˆ’4

    𝛼7

    1

    (π›½π‘›βˆ’ π›½π‘˜)3[𝐴

    𝑛𝐢𝑛𝐴

    π‘˜πΆπ‘˜βˆ’ 𝐴𝑛𝐢

    π‘›π΄π‘˜πΆ

    π‘˜] .

    (31)

    The first two lines can be simplified using the Wronskianand normalization factors, while the third can be rewrittenwith the help of the eigenvalue equation. Combining theseshows that the first three lines sum up to zero, leaving only

  • 8 Advances in Mathematical Physics

    the last term which we write in terms of derivatives of thewave functions at the origin:

    βŸ¨πœ•πΉπœ“π‘›| πœ“π‘˜βŸ©π·

    =𝐹

    𝛼 (πΈπ‘›βˆ’ πΈπ‘˜)3[πœ“

    𝑛(0βˆ’) πœ“

    π‘˜(0βˆ’) βˆ’ πœ“

    𝑛(0+) πœ“

    π‘˜(0+)] .

    (32)

    Comparing this expression to (24), it is clear that the rela-tion between coupling term and off-diagonal dipole matrixelements (8) is valid.

    Coupling terms (30) for the square-well system are alsocalculated in analytic form but due to excessive numberof terms comprising the result they are not listed here. Inprinciple, a similar procedure to simplify the square-wellcoupling term (30) should work. However, the resultingexpression is extremely large and we could not find apractical way to compress it to a manageable length. Themain difficulty in simplifying the square-well result is thatthe eigenvalue equation is muchmore complicated (cf. (14) to(15)). Nevertheless, having explicit formulas in terms of Airyfunctions, we verified numerically that the integrated resultdoes relate to the dipole matrix element as suggested by (8):

    βŸ¨πœ•πΉπœ“π‘›| πœ“π‘˜βŸ©π‘†

    =βˆ’π‘‰0

    (πΈπ‘›βˆ’ πΈπ‘˜)3[πœ“π‘›(βˆ’π‘‘) πœ“

    π‘˜(βˆ’π‘‘) βˆ’ πœ“

    𝑛(𝑑) πœ“π‘˜(𝑑)] .

    (33)

    To conclude this section, we have shown that the couplingterms can be directly calculated using a new Airy integraltechnique detailed in the appendix. While our explicit calcu-lations do not justify the formal steps taken to obtain (8), theydo corroborate that the relation between the coupling termsand the dipole moment holds. Pragmatically, one shouldchoose to use the dipolematrix elements, since they are easierto calculate numerically.

    5. Generalization for Arbitrary PiecewiseConstant Potentials

    Comparing results (18) and (27) we see that they havea similar form, with the dipole moment matrix elementsexpressed in terms of wave function values evaluated atdiscontinuities of the potential 𝑉. While the same result canbe easily written explicitly with Airy functions, this particularform indicates that the expressions are in fact sums overatomic potential discontinuities, with weights correspondingto the potential-value jumps. This suggests the followinggeneralization of the dipole matrix element formulas for asystem with arbitrary piecewise constant potential:

    𝑁2= βˆ‘

    𝑖

    Δ𝑉𝑖

    πΉπœ“ (π‘₯𝑖)2

    , (34)

    βŸ¨πœ“π‘›| π‘₯ | πœ“

    π‘˜βŸ©π‘†=

    1

    (πΈπ‘›βˆ’ πΈπ‘˜)2βˆ‘

    𝑖

    Ξ”π‘‰π‘–πœ“π‘›(π‘₯𝑖) πœ“π‘˜(π‘₯𝑖) , (35)

    where the sums run over all potential discontinuities. It is infact not too difficult to realize that the procedures utilized

    above can be modified for a more general case of piecewiseconstant potential. From there, one can take a continuumlimit, approximating an arbitrary potential as a limit ofpiecewise constant functions, and arrive at

    𝑁2=

    1

    πΉβˆ«π‘‘π‘‰

    𝑑π‘₯πœ“ (π‘₯𝑖)2 dπ‘₯, (36)

    βŸ¨πœ“π‘›| π‘₯ | πœ“

    π‘˜βŸ©π‘†=

    1

    (πΈπ‘›βˆ’ πΈπ‘˜)2βˆ«π‘‘π‘‰

    𝑑π‘₯πœ“π‘›(π‘₯) πœ“π‘˜(π‘₯) dπ‘₯. (37)

    We have assumed that the potential is short-ranged and theintegration in these formulas is along the real axis (i.e., thecontourC is not necessary for convergence).

    This is an intriguing result, because an identical formulacan be derived for the discrete-energy eigenstates of a self-adjoint Hamiltonian by evaluating its double commutatorwith the position operator. However, here we have Stark res-onances represented by complex-valued functions living onthe contour C. So it seems that as long as the normalizationand β€œscalar product” are defined with the help of pairing (3),the Stark resonance states satisfy relations analogous to thoseobeyed by their self-adjoint counterparts.

    We have used numerical simulations (not shown here)to verify that the relation between the Stark resonancepseudonorm and the generalized expectation value of theβ€œatomic” potential gradient could also be valid for three-dimensional systems. It is tempting to speculate that the off-diagonal dipole element relation (37) could be generally validfor Stark resonant states in higher dimensions.

    6. Conclusion

    We have derived analytic expressions for a number of quanti-ties that characterize the Stark resonance states in two exactlysolvable systems. The first model studied in this work isthe one-dimensional particle in a Dirac-delta potential withadditional homogeneous field, and the second has the square-well potential.

    We have studied these systems as open, non-Hermitianmodels, and identified a natural choice for the pairingconnecting the states in the domain of the Hamiltonian withthe states in the domain of its adjoint operator. With respectto this pairing, Stark resonances form an orthogonal system,and many of their properties can be evaluated analytically.

    Despite the fact that both models have been studied foryears, explicit expressions for their (pseudo-) norms, dipolemoment expectation values, and their relations connectingthe resonant state wave functions at different field valuesare new. Our results thus further the understanding of themathematical properties that underline the Stark effect.

    Moreover, we have shown that certain results naturallyextend to a wide class of one-dimensional models and wehave also identified relations that appear to be candidatesfor properties generally applicable to three-dimensional Starksystems. In particular, we have found that the generalizeddipole moment matrix elements between the nonphysicalresonant states can be related to the expectation values ofthe atomic potential gradient in a way that is completely

  • Advances in Mathematical Physics 9

    analogous to relations that hold for real-valued discrete-energy eigenfunctions in Hermitian systems. We speculatethat these relations apply generally in three dimensions andcould be used in numerical calculations to assess the fidelityof the resonant eigenfunctions.

    An important by-product of this study is a new integra-tion technique applicable to combinations of Airy functionsthat represent Stark resonances in one-dimensional modelswith piecewise constant potentials.

    Results presented in this work have also an immediatepractical impact on modeling of light-matter interactions instrong time-dependent optical fields in the framework of theMetastable Electronic State Approach.

    Appendices

    TheAiry integral technique outlined in this appendix is usedto evaluate integrals that contain linear combinations of Airyfunctions𝐴, 𝐡 and their derivatives𝐴, 𝐡, where the two setsof functions have shifted arguments. We use the shorthandnotation of 𝐴 = 𝐴[𝛼(π‘₯ + 𝛽

    𝑛)] and 𝐡 = 𝐡[𝛼(π‘₯ + 𝛽

    π‘˜)].

    The two unknown integrals that we are required to solveare

    ∫

    ∞

    βˆ’βˆž

    𝐴𝐡 dπ‘₯,

    ∫

    ∞

    βˆ’βˆž

    π‘₯𝐴𝐡 dπ‘₯,

    (βˆ—)

    where the integration is understood over π‘₯ along the contourC. Simply applying integration by parts leads nowhere sincethe solutions always involve integrals that are also not known.The technique described here circumvents that problem. Foran integral whose integrand has the form π‘₯𝑛𝐴𝐡, we performthe following steps:

    (1) Multiply integrand π‘₯𝑛𝐴𝐡 by π‘₯.

    (2) Differentiate with respect to π‘₯ and integrate.

    (3) Repeat (1) and (2) with the β€œsymmetric” integrandπ‘₯𝑛𝐴𝐡.

    (4) Subtract the two equations.

    And we find that we can write the unknown integral in termsof integrals that do not contain derivatives of 𝐴 or 𝐡 andare therefore found in published literature [35]. In the nextsectionswe demonstrate the procedure on the integrands𝐴𝐡and π‘₯𝐴𝐡.

    The procedure relies on the fact that wave functions andtheir derivatives are continuous at jumps of 𝑉(π‘₯). Moreover,𝐴 and 𝐡 must be resonant eigenstates that belong to theorthonormal system with respect to the pairing defined bythe contour integral along C (3). As such, the formulaswe derive below do not apply to arbitrary combinations ofAiry functions. On the other hand, the method does applyto a general case of Stark resonances in piecewise constantpotential and in this sense the result shown below is general.

    A. Integrating 𝐴𝐡

    To solve the integral of𝐴𝐡, wemultiply byπ‘₯ and differentiate

    πœ•π‘₯(π‘₯𝐴𝐡) = 𝐴

    𝐡 + 𝛼π‘₯𝐴

    𝐡 + 𝛼π‘₯𝐴

    𝐡

    = 𝐴𝐡 + 𝛼2π‘₯ (π‘₯ + 𝛽

    𝑛) 𝐴𝐡 + 𝛼π‘₯𝐴

    𝐡.

    (A.1)

    Next, we integrate both sides and find that there are knownand unknown Airy integrals:

    π‘₯𝐴𝐡

    +∞

    βˆ’βˆžβŸβŸβŸβŸβŸβŸβŸβŸβŸβŸβŸβŸβŸβŸβŸβŸβŸ

    =0

    = ∫

    ∞

    βˆ’βˆž

    𝐴𝐡 dπ‘₯

    ⏟⏟⏟⏟⏟⏟⏟⏟⏟⏟⏟⏟⏟⏟⏟⏟⏟⏟⏟⏟⏟

    want

    + 𝛼2∫

    +∞

    βˆ’βˆž

    π‘₯2𝐴𝐡 dπ‘₯ + 𝛼2𝛽

    π‘›βˆ«

    ∞

    βˆ’βˆž

    π‘₯𝐴𝐡 dπ‘₯⏟⏟⏟⏟⏟⏟⏟⏟⏟⏟⏟⏟⏟⏟⏟⏟⏟⏟⏟⏟⏟⏟⏟⏟⏟⏟⏟⏟⏟⏟⏟⏟⏟⏟⏟⏟⏟⏟⏟⏟⏟⏟⏟⏟⏟⏟⏟⏟⏟⏟⏟⏟⏟⏟⏟⏟⏟⏟⏟⏟⏟⏟⏟⏟⏟⏟⏟⏟⏟⏟⏟

    known

    + π›Όβˆ«

    ∞

    βˆ’βˆž

    π‘₯𝐴𝐡 dπ‘₯

    ⏟⏟⏟⏟⏟⏟⏟⏟⏟⏟⏟⏟⏟⏟⏟⏟⏟⏟⏟⏟⏟⏟⏟⏟⏟

    unknown

    .

    (A.2)

    To remove the unknown integrals that contain both deriva-tive functions 𝐴 and 𝐡, we differentiate the β€œsymmetric”integrand

    πœ•π‘₯(π‘₯𝐴𝐡

    ) = 𝐴𝐡

    + 𝛼π‘₯𝐴

    𝐡+ 𝛼π‘₯𝐴𝐡

    = 𝐴𝐡+ 𝛼π‘₯𝐴

    𝐡+ 𝛼2π‘₯ (π‘₯ + 𝛽

    π‘˜) 𝐴𝐡

    (A.3)

    and then integrate to find the complementary equation to(A.2):

    π‘₯𝐴𝐡

    +∞

    βˆ’βˆžβŸβŸβŸβŸβŸβŸβŸβŸβŸβŸβŸβŸβŸβŸβŸβŸβŸ

    =0

    = ∫

    ∞

    βˆ’βˆž

    𝐴𝐡 dπ‘₯

    ⏟⏟⏟⏟⏟⏟⏟⏟⏟⏟⏟⏟⏟⏟⏟⏟⏟⏟⏟⏟⏟

    unknown

    + 𝛼2∫

    ∞

    βˆ’βˆž

    π‘₯2𝐴𝐡 dπ‘₯ + 𝛼2𝛽

    π‘˜βˆ«

    ∞

    βˆ’βˆž

    π‘₯𝐴𝐡 dπ‘₯⏟⏟⏟⏟⏟⏟⏟⏟⏟⏟⏟⏟⏟⏟⏟⏟⏟⏟⏟⏟⏟⏟⏟⏟⏟⏟⏟⏟⏟⏟⏟⏟⏟⏟⏟⏟⏟⏟⏟⏟⏟⏟⏟⏟⏟⏟⏟⏟⏟⏟⏟⏟⏟⏟⏟⏟⏟⏟⏟⏟⏟⏟⏟⏟⏟⏟⏟⏟⏟

    known

    + π›Όβˆ«

    ∞

    βˆ’βˆž

    π‘₯𝐴𝐡 dπ‘₯

    ⏟⏟⏟⏟⏟⏟⏟⏟⏟⏟⏟⏟⏟⏟⏟⏟⏟⏟⏟⏟⏟⏟⏟⏟⏟

    unknown

    .

    (A.4)

    We subtract (A.2) and (A.4)

    ∫

    ∞

    βˆ’βˆž

    𝐴𝐡 dπ‘₯ βˆ’ ∫

    ∞

    βˆ’βˆž

    𝐴𝐡 dπ‘₯

    = βˆ’π›Ό2(π›½π‘›βˆ’ π›½π‘˜) ∫

    ∞

    βˆ’βˆž

    π‘₯𝐴𝐡 dπ‘₯.(A.5)

    And then we relate the two terms on the LHS using thefact that 𝐴𝐡 vanishes at the ends of the integration contour:∫𝐴𝐡 dπ‘₯ + ∫𝐴𝐡 dπ‘₯ = 0. We arrive at the unknown integral,

    in terms of a known one (VS 3.54):

    ∫

    ∞

    βˆ’βˆž

    𝐴𝐡 dπ‘₯ = βˆ’π›Ό

    2

    2(π›½π‘›βˆ’ π›½π‘˜) ∫

    ∞

    βˆ’βˆž

    π‘₯𝐴𝐡 dπ‘₯⏟⏟⏟⏟⏟⏟⏟⏟⏟⏟⏟⏟⏟⏟⏟⏟⏟⏟⏟⏟⏟⏟⏟

    3.54

    . (A.6)

  • 10 Advances in Mathematical Physics

    Here we have used the integration along the entire contour,but it is possible to apply the same technique to integrals overa finite range (π‘Ž, 𝑏).The only changewould be the inclusion ofthe LHS terms of (A.2) and (A.4) evaluated at the points π‘Ž and𝑏. Such a finite-range integral over (βˆ’π‘‘, 𝑑) appears during thecalculation of the coupling term for the square-well potentialin (30).

    B. Integrating π‘₯𝐴𝐡

    To solve this integral, we use the same procedure: multiplyintegrand by π‘₯ and differentiate and subtract β€œsymmetric”integrand. Consider

    πœ•π‘₯(π‘₯2𝐴𝐡) = 2π‘₯𝐴

    𝐡 + 𝛼π‘₯

    2𝐴𝐡 + 𝛼π‘₯

    2𝐴𝐡

    = 2π‘₯𝐴𝐡 + 𝛼2π‘₯2(π‘₯ + 𝛽

    𝑛) 𝐴𝐡

    + 𝛼π‘₯2𝐴𝐡,

    (B.1)

    πœ•π‘₯(π‘₯2𝐴𝐡) = 2π‘₯𝐴𝐡

    + 𝛼π‘₯2𝐴𝐡+ 𝛼π‘₯2𝐴𝐡

    = 2π‘₯𝐴𝐡+ 𝛼2π‘₯2(π‘₯ + 𝛽

    π‘˜) 𝐴𝐡

    + 𝛼π‘₯2𝐴𝐡.

    (B.2)

    Subtract (B.1) and (B.2) and then integrate

    0 = 2[[

    [

    ∫

    ∞

    βˆ’βˆž

    π‘₯𝐴𝐡 dπ‘₯

    ⏟⏟⏟⏟⏟⏟⏟⏟⏟⏟⏟⏟⏟⏟⏟⏟⏟⏟⏟⏟⏟⏟⏟

    want

    βˆ’ ∫

    ∞

    βˆ’βˆž

    π‘₯𝐴𝐡 dπ‘₯

    ⏟⏟⏟⏟⏟⏟⏟⏟⏟⏟⏟⏟⏟⏟⏟⏟⏟⏟⏟⏟⏟⏟⏟

    unknown

    ]]

    ]

    + 𝛼2(π›½π‘›βˆ’ π›½π‘˜) ∫

    ∞

    βˆ’βˆž

    π‘₯2𝐴𝐡 dπ‘₯

    ⏟⏟⏟⏟⏟⏟⏟⏟⏟⏟⏟⏟⏟⏟⏟⏟⏟⏟⏟⏟⏟⏟⏟⏟⏟

    known

    .

    (B.3)

    Again the two terms above can be related using integrationper parts plus an orthogonality argument. As a result, we canwrite the sought integral in terms of (VS 3.55). Consider

    ∫

    ∞

    βˆ’βˆž

    π‘₯𝐴𝐡 dπ‘₯ = βˆ’

    𝛼2(π›½π‘›βˆ’ π›½π‘˜)

    4∫

    ∞

    βˆ’βˆž

    π‘₯2𝐴𝐡 dπ‘₯

    ⏟⏟⏟⏟⏟⏟⏟⏟⏟⏟⏟⏟⏟⏟⏟⏟⏟⏟⏟⏟⏟⏟⏟⏟⏟

    3.55

    . (B.4)

    Conflict of Interests

    The authors declare that there is no conflict of interestsregarding the publication of this paper.

    Acknowledgment

    Thisworkwas supported by theUnited States Air ForceOfficefor Scientific Research under Grants nos. FA9550-13-1-0228and FA9550-10-1-0561.

    References

    [1] G. Gamow, β€œZur Quantentheorie des Atomkernes,” Zeitschriftfür Physik, vol. 51, no. 3-4, pp. 204–212, 1928.

    [2] A. J. F. Siegert, β€œOn the derivation of the dispersion formula fornuclear reactions,” Physical Review, vol. 56, no. 8, pp. 750–752,1939.

    [3] W. P. Reinhardt, β€œComplex coordinates in the theory of atomicand molecular structure and dynamics,” Annual Review ofPhysical Chemistry, vol. 33, no. 1, pp. 223–255, 1982.

    [4] A. Teleki, E. M. Wright, and M. Kolesik, β€œMicroscopic modelfor the higher-order nonlinearity in optical filaments,” PhysicalReview A, vol. 82, no. 6, Article ID 065801, 2010.

    [5] M. Kolesik, J. M. Brown, A. Teleki, P. Jakobsen, J. V. Moloney,and E. M. Wright, β€œMetastable electronic states and nonlinearresponse for high-intensity optical pulses,” Optica, vol. 1, no. 5,pp. 323–331, 2014.

    [6] L. Hamonou, T. Morishita, and O. I. Tolstikhin, β€œMolecularsiegert states in an electric field,” Physical Review A, vol. 86, no.1, Article ID 013412, 2012.

    [7] M. Glück, A. R. Kolovsky, and H. J. Korsch, β€œWannier-Starkresonances in optical and semiconductor superlattices,” PhysicsReports, vol. 366, no. 3, pp. 103–182, 2002.

    [8] W. Elberfeld and M. Kleber, β€œTunneling from an ultrathinquantum well in a strong electrostatic field: a comparison ofdifferent methods,” Zeitschrift für Physik B: Condensed Matter,vol. 73, no. 1, pp. 23–32, 1988.

    [9] C. M. Bender, β€œMaking sense of non-hermitian hamiltonians,”Reports on Progress in Physics, vol. 70, no. 6, pp. 947–1018, 2007.

    [10] N. Moiseyev, β€œQuantum theory of resonances: calculatingenergies, widths and cross-sections by complex scaling,” PhysicsReports, vol. 302, no. 56, pp. 211–293, 1998.

    [11] D. C. Brody, β€œBiorthogonal quantum mechanics,” Journal ofPhysics A: Mathematical and Theoretical, vol. 47, no. 3, ArticleID 035305, 2014.

    [12] A. Bohm, β€œResonances/decaying states and the mathematics ofquantum physics,” Reports on Mathematical Physics, vol. 67, no.3, pp. 279–303, 2011.

    [13] T. Berggren, β€œOn the use of resonant states in eigenfunctionexpansions of scattering and reaction amplitudes,” NuclearPhysics A, vol. 109, no. 2, pp. 265–287, 1968.

    [14] R. de la Madrid, G. Garcia-Calderon, and J. G. Muga, β€œReso-nance expansions in quantummechanics,”Czechoslovak Journalof Physics, vol. 55, no. 9, pp. 1141–1150, 2005.

    [15] J. Julve and F. J. de Urŕıes, β€œInner products of resonancesolutions in 1D quantum barriers,” Journal of Physics A: Mathe-matical and Theoretical, vol. 43, no. 17, Article ID 175301, 2010.

    [16] W. J. Romo, β€œInner product for resonant states and shell-modelapplications,”Nuclear Physics A, vol. 116, no. 3, pp. 617–636, 1968.

    [17] P. Lind, β€œCompleteness relations and resonant state expansions,”Physical Review C, vol. 47, no. 5, pp. 1903–1920, 1993.

    [18] O. Civitarese and M. Gadella, β€œPhysical and mathematicalaspects of Gamow states,” Physics Reports, vol. 396, no. 2, pp.41–113, 2004.

    [19] T. Berggren, β€œExpectation value of an operator in a resonantstate,” Physics Letters B, vol. 373, no. 1–3, pp. 1–4, 1996.

    [20] N. Hatano, T. Kawamoto, and J. Feinberg, β€œProbabilistic inter-pretation of resonant states,” Pramana, vol. 73, no. 3, pp. 553–564, 2009.

    [21] S. Geltman, β€œIonisation dynamics of a model atom in anelectrostatic field,” Journal of Physics B: Atomic and MolecularPhysics, vol. 11, no. 19, pp. 3323–3337, 1978.

    [22] G. P. Arrighini and M. Gavarini, β€œIonization of a model atomby strong and superstrong electric fields,” Lettere al NuovoCimento, vol. 33, no. 12, pp. 353–358, 1982.

  • Advances in Mathematical Physics 11

    [23] G. V. Dunne and C. S. Gauthier, β€œSimple soluble molecularionization model,” Physical Review A, vol. 69, no. 5, Article ID053409, 2004.

    [24] J. M. Brown, A. Lotti, A. Teleki, and M. Kolesik, β€œExactlysolvable model for nonlinear light-matter interaction in anarbitrary time-dependent field,” Physical Review Aβ€”Atomic,Molecular, and Optical Physics, vol. 84, no. 6, Article ID 063424,2011.

    [25] A. Emmanouilidou and N. Moiseyev, β€œStark and field-bornresonances of an open square well in a static external electricfield,” The Journal of Chemical Physics, vol. 122, no. 19, ArticleID 194101, 2005.

    [26] A. H. Larsen, U. De Giovannini, D. L. Whitenack, A. Wasser-man, and A. Rubio, β€œStark ionization of atoms and moleculeswithin density functional resonance theory,” The Journal ofPhysical Chemistry Letters, vol. 4, no. 16, pp. 2734–2738, 2013.

    [27] Y. Okajima, O. I. Tolstikhin, and T.Morishita, β€œAdiabatic theoryof high-order harmonic generation: one-dimensional zero-range-potential model,” Physical Review A, vol. 85, no. 6, ArticleID 063406, 2012.

    [28] Ya. B. Zel’dovich, β€œOn the theory of unstable states,” Journal ofExperimental andTheoretical Physics, vol. 39, pp. 776–780, 1960.

    [29] R. M. More and E. Gerjuoy, β€œProperties of resonance wavefunctions,” Physical Review A, vol. 7, pp. 1288–1303, 1973.

    [30] N. Moiseyev, Non-Hermitian Quantum Mechanics, CambridgeUniversity Press, Cambridge, UK, 2011.

    [31] A. Bahl, J.M. Brown, E.M.Wright, andM.Kolesik, β€œAssessmentof the metastable electronic state approach as a microscopicallyself-consistent description for the nonlinear response of atoms,”Optics Letters, vol. 40, no. 21, pp. 4987–4990, 2015.

    [32] D. Bohm, Quantum Theory, Dover Books on Physics Series,Dover Publications, 1951.

    [33] A. Mondragn and E. Hernndez, β€œAccidental degeneracy andberry phase of resonant states,” in Irreversibility and CausalitySemigroups andRiggedHilbert Spaces, A. Bohm,H.-D.Doebner,and P. Kielanowski, Eds., vol. 504 of Lecture Notes in Physics, pp.257–281, Springer, Berlin, Germany, 1998.

    [34] S. Panda and B. K. Panda, β€œAnalytic methods for field inducedtunneling in quantum wells with arbitrary potential profiles,”Pramana, vol. 56, no. 6, pp. 809–822, 2001.

    [35] O. Valle and M. Soares, Airy Functions and Applications toPhysics, Imperial College Press, London, UK, 2004.

  • Submit your manuscripts athttp://www.hindawi.com

    Hindawi Publishing Corporationhttp://www.hindawi.com Volume 2014

    MathematicsJournal of

    Hindawi Publishing Corporationhttp://www.hindawi.com Volume 2014

    Mathematical Problems in Engineering

    Hindawi Publishing Corporationhttp://www.hindawi.com

    Differential EquationsInternational Journal of

    Volume 2014

    Applied MathematicsJournal of

    Hindawi Publishing Corporationhttp://www.hindawi.com Volume 2014

    Probability and StatisticsHindawi Publishing Corporationhttp://www.hindawi.com Volume 2014

    Journal of

    Hindawi Publishing Corporationhttp://www.hindawi.com Volume 2014

    Mathematical PhysicsAdvances in

    Complex AnalysisJournal of

    Hindawi Publishing Corporationhttp://www.hindawi.com Volume 2014

    OptimizationJournal of

    Hindawi Publishing Corporationhttp://www.hindawi.com Volume 2014

    CombinatoricsHindawi Publishing Corporationhttp://www.hindawi.com Volume 2014

    International Journal of

    Hindawi Publishing Corporationhttp://www.hindawi.com Volume 2014

    Operations ResearchAdvances in

    Journal of

    Hindawi Publishing Corporationhttp://www.hindawi.com Volume 2014

    Function Spaces

    Abstract and Applied AnalysisHindawi Publishing Corporationhttp://www.hindawi.com Volume 2014

    International Journal of Mathematics and Mathematical Sciences

    Hindawi Publishing Corporationhttp://www.hindawi.com Volume 2014

    The Scientific World JournalHindawi Publishing Corporation http://www.hindawi.com Volume 2014

    Hindawi Publishing Corporationhttp://www.hindawi.com Volume 2014

    Algebra

    Discrete Dynamics in Nature and Society

    Hindawi Publishing Corporationhttp://www.hindawi.com Volume 2014

    Hindawi Publishing Corporationhttp://www.hindawi.com Volume 2014

    Decision SciencesAdvances in

    Discrete MathematicsJournal of

    Hindawi Publishing Corporationhttp://www.hindawi.com

    Volume 2014 Hindawi Publishing Corporationhttp://www.hindawi.com Volume 2014

    Stochastic AnalysisInternational Journal of