Polarizabilities as Probes for P, T, and PT Violation
Abstract
Searches for violations of the fundamental symmetries of parity and time reversal in atomic and molecular systems provide a powerful tool for precise measurements of the physics of and beyond the standard model. In this work, we investigate how these symmetry violations affect the response of atoms and molecules to applied electric and magnetic fields. We recover well-known observables like the -odd, -odd spin-electric field coupling that is used for searches of the electron electric dipole moment (eEDM) or the effect of -odd, -even optical rotation in atomic gases. Besides these, we obtain several other possible observables. This includes, in particular, effects that can only be seen when using oscillating or inhomogeneous fields.
1 Introduction
At the present day, precision measurements in atomic and molecular systems are among the most powerful tools to test the standard model of particle physics (SM) at low energy scales [1, 2]. This further enables searches for new particles or interactions that are proposed to answer some of the open questions in the foundations of physics. The interaction of atoms and molecules with these new particles would lead to small shifts in their spectra. The magnitude of almost all of these small corrections is far below the uncertainty at which absolute energy levels can presently be calculated. To avoid this problem, it makes sense to search for observables that are forbidden by the symmetries of atomic physics. If we detect non-zero values of those, their size would directly correspond to some new or exotic interaction. Two such symmetries are the ones of Parity and time reversal .
-violating processes were previously used to proof the validity of the Glashow-Weinberg-Salam theory at atomic energy scales [2]. Further, numerous searches for simultaneous violations of and were performed in the context of electric dipole moment (EDM) measurements [3, 4]. Only few direct searches for , without violation (-odd -even, also called ToPe) have been performed so far [5, 6].
Following previous works [7, 8, 9, 10, 11, 12], we derive possible observables for measurements of , , and -violations in the interaction of electromagnetic fields with atoms and molecules. We try to
be as general as possible by making minimal assumptions on the atoms/molecules under consideration, the spatial and temporal shape of the electromagnetic fields, and the explicit form of the - and - odd potentials.
We start by specifying the requirements of the system (sec. 2.1) and then introduce perturbation through electric and magnetic fields under a low-order multipole approximation (sec. 2.2). On this basis, we derive the time-dependent perturbative expressions for the induced electric and magnetic dipole moments and the electric quadrupole moment (sec. 2.3). We separate these terms into irreducible tensor components (sec. 3.3.2, and 3.3.4) and analyze their behavior under and transformations. This leads us to general expressions for the possible (generalized) atomic polarizabilities (sec. 3.4). In sec. 4, we match some of these polarizabilities to observables that have been considered in previous works. We further identify terms that have not been explored so far. This includes, in particular, interactions with inhomogeneous (sec. 4.2) and time-dependent sec. (4.3) fields. We show that the former leads to observables sensible to all three kinds of symmetry violations. In sec. 5 we argue that the latter introduces some difficulties in the context of -odd observables.
2 Derivation of the and odd polarizabilities
2.1 Requirements on the system
The atomic or molecular system under consideration has to fulfill a few requirements for the following treatment to be valid. Namely, the Hamiltonian of the unperturbed system needs to commute with Parity , the square of the total angular momentum and its projection on an arbitrary axis . Therefore we require to be invariant under rotations and spacial reflections. Generally, this will be true for any free atom or molecule due to the rotational symmetry of their wavefunctions. Interestingly enough, this holds independent of how precise our model for the system is. It even stays valid for a full relativistic QED+QCD description of all its inner processes. Only the inclusion of the weak force or hypothetical new physics interactions breaks these symmetries. Therefore, we exclude these from the definition of and instead add them later as small perturbations.
From all this, we know that the eigenstates to can (without approximation) be written in a basis . Here, is the quantum number of the total angular momentum, is the quantum number related to its projection on a chosen quantization axis, and includes all the other quantum numbers of the system.
This comes with a small caveat. To use all this, we need to be able to prepare a system in these ”true” eigenstates. This will, for example, not be possible for chiral molecules whose states of different chirality are separated by such a large energy barrier that they will realistically never be found in a Parity eigenstate (Hund’s Paradox [13]). Apart from chiral molecules [14, 15],
spontaneous breakings of and -symmetries can also arise in solid state systems depending on their crystal symmetries [16, 17]. These do not originate from the fundamental physics processes we want to consider here. We, therefore, require that all such spontaneous symmetry breaking are absent. This is generally fulfilled in free atoms and (small) molecules.
Throughout this paper, we are using natural units: . In the following, we refer to the situation where an operator is both odd under and as -violation. We use the notation for the total angular momentum, including possible contributions of the nuclear angular momentum, even though usually the momentum in such a case would be denoted by .
2.2 Interaction with electromagnetic fields
We want to study how the system behaves under the influence of applied electromagnetic fields. To do this, we are using a perturbative treatment in which we add the semiclassical interaction with an electromagnetic field in the 2nd-order multipole approximation:
(1) |
where , and are the electric dipole, the magnetic dipole, and the electric quadrupole (Hermitian) operators, respectively. We recall that in a simple single electron system, these moments are defined as follows [18]:
Here, is the absolute value of the electron charge, is the position- and the momentum-vector of the electron, is the Bohr magneton, , the electron -factor, and is the vector of the three Pauli matrices. In the definition of , we included the first-order relativistic correction [19]. In the remainder of this work, we do not refer to the specific form of these operators and only use their behaviors under parity and time reversal transformations. That way, the treatment stays valid for multi-electron atoms and molecules in which , , and are obtained by summing over all bodies in the system.
The electric and magnetic fields and are defined through the electromagnetic scalar and vector potentials:
For a simple separation between time-independent and time-dependent terms, we use the gauge freedom to set .
We define in a rather general way, as a superposition of a static component and plane waves with arbitrary frequencies and polarizations:
is the (real) static component of the vector potential (). is the wavevector, and is the angular frequency of the wave. The real part is defined as , with being the complex conjugate to . Throughout this paper, both upper and lower indices are used for ease of notation without the positioning carrying a specific meaning. From these choices, the electric field is given by:
(2) |
with the first term describing the static and the second the time-dependent field component. Accordingly, for the magnetic field :
(3) |
The Jacobian of is given by:
(4) |
We solve the system in the inertial center of mass frame of the atom (molecule). For a moving system, it is necessary to transform the field values and frequencies accordingly.
In the multipole approximation, only the field values at the center of mass contribute [18]:
For the time-dependent field components, this implies . For simplicity of notation, we will not explicitly indicate this limit in the following.
2.3 Perturbative treatment
As usual, the Energy of the system is given by:
with being the eigenvectors of .
In the following, we want to treat as a perturbation to . We will perform this perturbation explicitly for the electric dipole operator . The perturbations for and follow analogously. The general case, valid for both static and varying fields, can be built from the solutions to the time-independent and dependent problems. Thus, we do the perturbation in two steps. First, we perform a time-independent perturbation calculation for the static components of the field and, afterward, a time-dependent calculation for the time-varying components. There are multiple, slightly different, formulations for the Dirac-type time-dependent perturbation theory. One has to be careful in choosing the right one to avoid the appearance of secular terms and obtain a correct normalization and convergence in a static limit [20, 21, 22, 23].
A free atom (or molecule) is fully degenerate in the quantum number . We assume that there are no further (accidental) degeneracies present. For these assumptions, the states will be ”good states”[24] in respect to all the operators involved, which allows us to perform a treatment analogous to a nondegenerate one. The maximal field strengths to which such a perturbative treatment will stay valid are highly dependent on the system and states under consideration.
2.3.1 Time-independent perturbation
In a static system, , and . The 1st order (Rayleigh-Schrödinger) perturbation theory for a potential leads to
(5) |
with . Thus, up to 1st order, the matrix element under the influence of the perturbation can be written as:
(6) | ||||
2.3.2 Time-dependent perturbation
For deriving the time-dependent contributions, we start by writing the non static part of Eq. (1) explicitly under the use of Eqs. (2), (3), and (4):
We can see that we have the special situation of a perturbation periodic in time (see, for example [20, 25]). In this case, the induced dipole moment is given by:
(7) | ||||
This expression is only true far from resonance . The situation close to resonance is discussed in sec. 5.
By expanding the fractions and identifying with the time derivative of , we can rewrite this expression as:
(8) |
In the static case, , vanishes, and the eq: returns to the same form as Eq. (6). Therefore, we can combine both. After reinserting the definition of , we arrive at a general expression for the electric dipole moment element:
(9) | ||||
Here and in the following, we imply, according to the Einstein convention, the summation over (including the static case , ). Above, we further introduced the (generalized) polarizabilities:
(10) | ||||
with , , , and following analogously.
Likewise, we can perform the same derivation for the magnetic dipole moment, which leads to:
(11) | ||||
Finally, for the electric quadrupole moment, we get:
(12) | ||||
For simplicity and consistency, we will refer to all terms as (generalized) polarizabilities, even if, for specific cases, other names such as permeability or quadrupolarizability exist. We further even refer to the field-independent static moments as polarizabilities.
3 Analyzing the polarizabilities in terms of parity and time reversal
In usual atomic physics derivations for electric and magnetic dipole moments, most of the terms in Eqs. (9), (11), and (12) are absent. As we will see in a moment, this is due to the symmetries of . These terms can, however, become observable when is additionally perturbed by - and -odd potentials :
(13) | ||||
is an internal -odd, -even, a -odd, -odd, and a -even, -odd interaction. The global minus sign in does not carry a physical meaning and is just chosen in analogy to Eq (1).
In the following, we will take to be the eigenstates of this full Hamiltonian .
To make the treatment more general, we do not specify the structure of these operators but instead only use their transformation behaviors under and . For explicit examples of , , and operators that arise in electron-nucleon interactions, see, for instance, Ref. [26, 27]. We restrict ourselves to symmetry-violating interactions that take place inside the free atom or molecule and do not directly couple to the externally applied fields. For many situations, this assumption makes sense as the internal electric and magnetic fields will exceed the externally applied ones in strength.
In the following, we want to investigate in detail which terms in Eqs. (9), (11), and (12) correspond to the breakings of which particular symmetries. The most important properties of the and operators are indicated in table 1.
- | + | + | + | - | - | + | |
+ | - | + | - | + | - | - |
3.1 and odd polarizabilities
The effects of , , and on , , and can again be derived using perturbation theory. We discuss the details of this procedure in appendix B. From it, we get a new, more general, expression for the induced electric dipole moment in terms of -even and -odd polarizabilities:
(14) | ||||
with . The first term in each row is, as before, defined according to eq: (10). As indicated by their lower left indices, the remaining polarizabilities are proportional to one of the three potentials , , and . We define as (see sec. 3.3):
(15) |
with and following accordingly.
The remaining terms are defined analogously to:
(16) |
Likewise, expressions for and in terms of and odd polarizabilities can be derived. We will now analyze these polarizabilities regarding their transformation behavior under and . From this, we will see that around three-quarters of all possible terms in Eq. (14) necessarily vanish in all systems that obey our initial assumptions made in sec. 2.1.
3.2 and properties of polarizabilities in zero field: 0th order in
We start by discussing the vector polarizabilities , , , and . Because these arise in the 0th order perturbation in , they are independent of the externally applied field. From Eq. (14), we can see that their existence results in a permanent electric dipole moment (EDM).
3.2.1 P-violation
It can be easily shown that due to the odd parity of . Using the properties given in Table 1 we get:
(17) |
Thus .
To avoid this, we need to mix states of different parity. Effectively, this is what happens when we introduce the perturbation trough or . In this case:
Therefore is not forbidden by parity arguments.
3.2.2 T-violation
It is further possible to show that also vanishes due to reasons of -symmetry. This fact was first demonstrated by Lee and Yang [28]. Through the Wigner Eckart theorem, we know that the vector operator has to be oriented along the total angular momentum :
(18) |
with a real -independent constant [18]. We can now introduce the time reversal operator to the LHS of Eq. (18). Using properties given in Table 1 we get:
Where the complex conjugation originates from the antilinear properties of (see appendix A). Analogously, we can show for the RHS of Eq. (18):
where we used that is -odd. We can now rename and us the fact that both , and are Hermitian, to arrive at:
Then the comparison with Eq. (18) clearly implies .
We now show how this argument fails under the introduction of a -odd potential or :
Where we introduced the Hermitian vector operator . This operator is odd under time reversal (and even under Parity):
where we used that is -odd while and are -even.
Previously, for the Lee-Yang argument, the contradiction arose because and , which are proportional to each other, transform differently under . Now that both and are -odd, the contradiction vanishes.
The combined requirements on parity and time reversal leave as the only possible source for an EDM and tells us that such a property is directly proportional to the size of .
3.2.3 magnetic dipole moment
Contrary to the electric dipole moment, the 0st-order expectation value of the magnetic dipole moment is already classically allowed. We can also easily see this from the fact that and transform in the same way under and . From this and our derivation above, it directly follows that the value of the static magnetic dipole moment is insensitive to the effects of , , and . Therefore, is the only non-zero contribution to the static magnetic dipole moment. Through the Wigner-Eckart theorem, we can express it as
where is a real constant and equivalent to the factor that we introduced above. We further define the notation
for the expectation value of the total angular momentum in a given state.
3.2.4 Electric quadrupole moment
The electric quadrupole operator is both symmetric and traceless. The expectation value of such an object can only be proportional to the angular momentum quadrupole [29]:
Therefore, we can write:
3.3 and properties of polarizabilities in 1st order in
We now continue by analyzing the rank-2 tensor polarizabilities by using the techniques that we developed in the previous section.
3.3.1 Restrictions from parity
As previously in Eq. (17), we can analyze the that are defined in Eq. (16) by introducing into the expression. It can easily be seen that terms that contain an odd number of -odd operators will vanish ( for example contains the -even operators , and and the -odd operator ). Therefore, half of all possible polarizabilities vanish due to parity reasons.
3.3.2 Tensor decomposition of
Before we are able to analyze in terms of time reversal, we first need to understand the structure of these (Eucledian) rank-2 tensors. Such objects can be decomposed into a sum of three components that transform as scalar, antisymmetric vector, and traceless symmetric tensors respectively [29]:
with being the Kronecker delta. Angular momentum algebra tells us that the expectation value of a vector operator is necessarily oriented along , and the one of a traceless rank-2 tensor needs to be proportional to [29]. This allows us to express the tensor , with only three constants , and :
(19) |
with the Levi-Civita symbol. A non-zero value for the vector term requires the system to possess , while the tensor term will only exist for systems with . , and depend on the quantum numbers and , but not . We use the bar above them to indicate that they are complex valued. In the following, we will construct expressions for the tensor decomposition in terms of purely real constants.
3.3.3 Restrictions from time reversal
To further eliminate terms, we can again make use of their transformation under time reversal similar to [30] and to what we did in sec. 3.2.2.
We will perform this procedure on the example of (we ignore for now that this polarizability vanishes for Parity reasons). Again, we can decompose into its irreducible components:
(20) | ||||
We introduce into the LHS that we previously defined trough Eq. (10) to get:
(21) | ||||
The RHS of Eq. (20) transforms as:
(22) | ||||
The important observation here is that the vector part gets a minus sign in relation to the scalar and tensor parts.
We can now equate Eq. (21) with Eq. (22) and rename to obtain:
The comparison with Eq. (20) finally allows us to express in terms of real coefficients:
with . The decomposition of , that only differs by a factor of follows in the exact same way.
If we perform an analog treatment on , everything is basically the same except we do not get the global minus sign in Eq. (21) due to the different -transformation behavior of and . Therefore:
(23) |
In appendix C.1, we further show that the argument from above still works for and -odd polarizabilities that are defined according to Eq. (16).
We conclude that polarizabilities containing an even number of -odd operators possess purely real scalar and tensor components and a purely imaginary vector component. Polarizabilities with an odd number of -odd operators, on the other hand, have a real vector component but imaginary scalar and tensor parts. From the definitions in 3.1, we can see that the polarizabilities that couple to the fields need to be necessarily real while the polarizabilities that couple to the field derivatives need to be imaginary.
3.3.4 Dipole-quadrupole terms: rank-3 tensor decomposition
Before we can use the statements that we derived above to formulate general reduced expressions for the induced multipole moments, we still need to examine the rank-3 polarizability tensors. These arise in Eqs. (9), and (11) as cross polarizabilities between electric and magnetic dipoles and the electric quadrupole moment.
As a symmetric tensor, the quadrupole operator can only have 6 independent elements, leading us to an with elements. The decomposition under of a rank-3 tensor that is symmetric in its last two components was derived in Ref. [31]:
where the first two tensors are symmetric and the second two are of mixed symmetry. This allows us to write the rank-3 polarizabilities as:
where are real symmetric, and are real mixed-symmetry tensors. They depend on the indicated powers of . From the definitions given in [31], we can derive the following tensors:
(24) | ||||
So far, we have not used the fact that the quadrupole tensor is traceless. This reduces its number of independent elements by one, leading to . For the tensor , this means that the partial trace over the last two components vanishes. , , and fullfill this condition. For the symmetric vector component , we have however:
This implies . Therefore the contribution of vanishes and we arrive at the decomposition of the rank-3 tensor into only 3 irreducible structures:
Again, the decomposition of follows completely analogous.
3.3.5 Dipole-quadrupole terms: restrictions under parity and time reversal
The restrictions from parity still apply as before: only polarizabilities containing an even number of -odd operators can exist.
To analyze the behavior under time reversal, we can proceed analogously to sec. 3.3.3. There, we saw that if the polarizability contains an even number of -odd operators, then the -even scalar and tensor terms are real while the -odd vector component is purely imaginary. As can be seen by the indicated powers in , the rank-3 polarizability components proportional to , and are -odd, while the one proportional to is -even. Therefore, following the same line of argument as before, we get:
even number of -odd operators: | |||
odd number of -odd operators: |
with .
3.3.6 Rank-4 tensors
The only rank-4 tensors that appear in our derivation are: , , , and . Without decomposing the tensor, we can show that , , and vanish (see appendix C.2). Because we are mainly interested in symmetry-violating polarizabilities, we will not enter deeper into the structure of . The decomposition of a tensor with the same symmetries as can be found in [32]. This all applies to the primed tensors in the same way.
3.4 Full expressions of the induced moments
We can now use the results from sec. 3 to decompose all polarizabilities in Eq. (14) into their irreducible components and eliminate all terms that vanish due to reasons of parity and time reversal. From all this, we arrive at a general expression of the induced electric dipole moment:
(25) | ||||
We summarize the notations used in this expression: is the electric and the magnetic field, while and are their time derivatives.
is the (conserved) total angular momentum of the system, and . is the symmetric traceless matrix defined by . , , and are rank-3 tensors that have been defined in sec. 3.3.5. is a unit vector along the (Cartesian) coordinate .
The real coefficients , are the (generalized) polarizabilities that result from the tensor decomposition of the quantities defined in sec. 3.1.
As before, the lower right index of the defines which structure they belong to (=scalar trace, =antisymmetric vector, =symmetric traceless quadrupole tensor, = symmetric traceless octupole tensor, =mixed symmetry vector, =mixed symmetry tensor ).
The lower left index of the tells if the interaction requires perturbation by a , , or both odd potential.
Equivalently to the electric dipole, we can find the induced magnetic dipole moment:
(26) | ||||
For the electric quadrupole, we get:
(27) | ||||
We can see that each field configuration of , , , , and uniquely couples to a specific symmetry violation.
The absence of the odd polarizabilities , , and in the expressions above is explained in appendix C.2.
3.5 Induced , , and
Additionally to the induced moments that are defined above, also the expectation values of the symmetry violating potentials , , and themselves contribute to the full perturbative expressions of , and . Obviously, all these potentials vanish in 0st order in :
If we now consider how , , and are perturbed through , one could assume that the situation will be identical to the discussion above where we perturbed through , , and . This is, however, only the case for purely static fields, as we can see in the following example:
We consider the perturbation of through . Such a system can be solved completely analogously to our derivation in sec. 2.3, cf. Eq. (8), with the replacement :
(28) | ||||
From our discussion in sec. 3.2.2, it follows that must be fully imaginary.
While the real part (that is similar to the perturbation of through : ) vanishes, we still remain with a non-zero imaginary part:
From this, we can also conclude that even if the Lee-Yang argument (see sec.3.2.2) forbids an atom to possess a static electric dipole moment without -violation, this does not mean that there can not be a purely -odd linear response to an applied (time-dependent) electric field.
If we perform an analogous treatment to calculate how , , and are perturbed in 1st order through , we get only a small number of terms that are not forbidden by parity and time reversal arguments:
(29) | ||||
Regarding the (time-dependent) mixing of states or energy shifts, these constitute the lowest order (in ) corrections that violate , , and , respectively.
Deriving the 2nd order perturbations through is more complicated. This is because we have two time-dependent perturbers, which, through interferences, create a multitude of terms (see, for example, Refs. [20, 33, 23]). For this reason, we limit our discussion of the 2nd order corrections of , , and to the time-independent case. In appendix B.2, we demonstrate that the resulting terms coincide with those we derived in sec. 3.3.
4 Resulting observables
We now want to connect the terms we derived above to possible experimental observables. It will become apparent that the derivation presented in this paper recovers many - and/or -odd observables that have been considered in other works. Additionally, we obtain some -, -, and -odd observables that have, to our knowledge, not been discussed so far.
4.1 Static homogeneous fields
We start by reviewing the situation for static homogeneous fields (). Under these considerations, most terms in Eqs. (25), (26), and (27) vanish. We are left with:
(30) | ||||
with .
In Appendix D, we derive that the induced energy shift of the system can be expressed through the same polarizabilities:
(31) |
where the first line shows the terms that are even and the second line the terms that are odd under either or .
We start by discussing the former. The first two terms are well-known and usually have the largest influence on the system. They describe the linear Zeeman- and the quadratic Stark effect. The third term is the quadratic Zeeman effect [34]. It is best visible in diamagnetic systems where the linear Zeeman effect is absent. The fourth term is the Stark tensor-polarizability that can be relevant for systems with [35]. The final term is an equivalent Zeeman tensor-interaction. It has been first considered in Ref. [36] as a way to induce an electric quadrupole moment through an applied magnetic field.
Now, we are going to discuss the symmetry-violating interactions, which are the main interest of this study.
The term is well known in the context of searches for the electric dipole moment of the electron (eEDM). It allows to restrict - and -violating physics by measuring the absence of linear stark shifts in paramagnetic atoms and molecules. This interaction does not uniquely arise from the eEDM but can also be caused by a variety of other -violating interactions, like Schiff moments, the scalar-pseudoscalar electron-nucleon coupling, or the electron interaction with the nuclear magnetic quadrupole moment. For a review on the topic, see Ref. [3]. For purely diamagnetic systems, is zero. Such systems are, therefore, insensitive to the effects of .
The remaining three terms in Eq. (31) were first considered in Ref. [8]. The first of them, , describes a direct coupling of the electric and magnetic fields similar to the magnetoelectric effect that can arise in some solids [16]. In Ref. [12], its effects were investigated in more detail. Through its scalar nature, contributes to para- and diamagnetic systems alike.
The properties of the term were studied in Ref. [37], and [38]. Because of its antisymmetric structure, it is only nonvanishing when the angular momentum is misaligned with the direction of (and also the plane spanned by and ).
The final term again describes a -odd electric magnetic field coupling. Apart from the initial proposal of this expression in Ref. [8], we do not know of any other work discussing this coupling for static fields. This term requires a system with .
To measure these terms, one could directly measure the energy shift (31) depending on the strength and direction of the applied fields. Alternatively, one could use a large sample and measure the induced polarization , or magnetization , from Eq. (30). An induced polarization in a solid can be measured with a precise voltage measurement as it was done in [39] in the context of an eEDM search. Magnetizations of samples can, for example, be measured with a SQUID as it was done in another eEDM search in [40].
4.1.1 Alignment of in externally applied fields
The expectation value of the angular momentum is a vector whose orientation depends only on the (unperturbed) state of the system. But even though is not directly dependent on the applied and fields, its direction can still be influenced by them.
This is because the energy shift in (31) defines which -substate in the -manifold has the lowest energy. If the system possesses the ability to decay into its ground state, changes. For example, if we apply a strong magnetic field to a sample, the Zeeman coupling dominates all the other terms. It is minimized if becomes maximally aligned to the direction of . In this case, we can perform the replacement . Here, is the usual angular momentum quantum number.
Alternatively, if magnetic fields are absent, then the application of an electric field will orient directly through the EDM coupling . We know from sec. 3.2.3 that an oriented angular momentum directly leads to an oriented magnetic moment . It again could be picked up through magnetometry. This method of measuring the eEDM was first proposed in [41].
4.2 Static inhomogeneous fields
If we allow the electric field to vary in space (), then the electric quadrupole terms will also contribute to the energy of the system. One possibility to apply strong field gradients to atoms or molecules is to embed them into a crystalline matrix [42, 43, 44].
In inhomogeneous fields, the dipole moments are extended by:
The dots represent the terms from in Eq. (30). The inhomogeneous terms lead to an additional energy shift (see Appendix D):
(32) | ||||
The expressions in the first two lines are classically allowed. The first two of these terms describe the interactions of static and induced electric quadrupole moments with the inhomogeneous electric field analogous to the linear and quadratic Stark and Zeeman effects. The next two terms describe the induction of an electric quadrupole moment through an applied magnetic field. That such an interaction, linear in , exists, was only considered quite recently [45].
The third and fourth line in Eq. (32) describes , , and -odd effects. We proceed by analyzing these terms in the scenario where the angular momentum structures , , and become aligned through the externally applied fields.
4.2.1 Alignment of , , and in externally applied fields
The terms in the third line of Eq. (32) contain the following tensor, which is itself dependent on the quadrupole angular momentum structure :
Similar to the alignment of into the direction of an applied magnetic field discussed in sec. 4.1.1, as well can be aligned through (large) -even -even interactions and sufficient field strengths. In this case, , , and compete with each other to define the specific state that minimizes the energy and indirectly defines the orientation of . We can express this by performing the replacement
(33) |
Here , , and are real constants that depend on the sizes of the different classically allowed polarizabilities . The expression is then normalized through the Frobenius norm. represents the trace of the fraction. We need to subtract it, to ensure that the new object still remains traceless.
The first of the two -dependent terms in Eq. (32) describes an interaction that only violates , but not . That such a quantity could exist in an inhomogeneous electric field was first proposed in Ref. [38]. Now, we will derive what the resulting observable looks like. The quadrupole component of an electrostatic field is given by the potential: . It is oriented along some axis and has an amplitude of . without loss of generality, we can take in -direction: . It follows for the electric field Jacobian:
From Eqs. (24), (32), and (33), we can now derive the energy shift that is induced by the term:
From the antisymmetry of , it follows that , , and necessarily vanish. We can write the remaining term as:
(34) |
Expressions of a similar form have previously been derived in [46] and [47]. There, it was shown that the energy shift induced by the -odd nuclear anapole moment is at a range that could realistically be measured with the current technology.
Analogously to the derivation of , the same effect also leads directly to an induced electric dipole moment:
In contrast to the -odd electric dipole that results from, this -odd -even one is oriented orthogonal to the direction of the applied magnetic field and stays invariant under .
The second -dependent term in Eq. (32) behaves quite similarly. For it, we can derive:
(35) | ||||
Unlike all effects discussed so far, this one is -odd, but -even (ToPe). To our knowledge, all atomic ToPe observables that were previously considered in the literature depend on dynamic processes. This would make this the first static mechanism for detecting such a symmetry violation. A review on ToPe physics can be found in [5, 27]. It could be possible to measure in a similar scheme to the one used in [40].
The energy shift in Eq. (32) further contains two -odd terms. The first of them, couples the electric field and the Hessian to the octupole angular momentum structure . Similar to the quadrupole in Eq. (33), this octupole as well can be aligned through the application of external fields.
The other -odd term in Eq. (32) depends on the symmetric vector quantity:
We can derive a simplified form for similar to what we did for :
As discussed in sec. 4.1.1, can align to the direction of an applied magnetic field, leading to:
In contrast to the -odd -even, and the -even -odd terms in Eqs. (34), and (35), this expression is both odd under the inversion of and . If and are parallel, but orthorgonal to , then the contribution from this interaction becomes very similar to , and .
Again, we also get an associated induced electric dipole moment:
To our knowledge, the -odd terms described in this section have not been discussed before.
We can conclude that a static quadrupolar electric field, as it can be present in crystals, allows one to use a combination of electric and magnetic fields to search for all three types of symmetry violations. Alternatively, to align the different angular momentum structures through applied fields, one could, of course, also prepare the system in a different state by optical pumping.
4.3 Time-varying fields
If we allow for time-dependent - and -fields, all the terms in Eqs. (25), (26), and Eq. (27) contribute.
To measure these time-dependent observables, one could again directly pick up the induced magnetization or polarization .
For example, implies that applying an oscillating electric field to a polarized sample will induce an orthogonal magnetization. Its -odd part oscillates in phase with the applied field, but its -odd part is out of phase.
Another way to measure the time-dependent polarizabilities is to prepare the system in a well-defined state and observe the time evolution of its population when it is exposed to time-varying fields [48].
Finally, the effects of the and induced by electromagnetic waves propagating through a medium can be studied.
Let us consider the terms . As shown in Fig. 1, for a left circular polarized wave, is parallel to , while for a right circular polarized wave, they are antiparallel. From this, it follows that we can rewrite for the two polarizations as . For dilute gases, the refractive index relates to the polarizability through . This implies that the refractive index of left and right polarized light differ by an amount that is directly proportional to the size of the parity-violating potential . This difference leads to a small optical rotation of light propagating through an atomic or (nonchiral) molecular gas.
The effect of -odd optical rotation was successfully utilized to measure atomic parity violations in Bi, Pb, and Tl ([49] and references within).
This is just one example of , violation-induced birefringence and optical activity. In Refs. [9, 11] some others are discussed.
However, there is a caveat when dealing with time-dependent fields, which we discuss in the following section.
5 Effects from the natural linewidth
So far, we have neglected the effects from the natural linewidth of the excited states in our perturbative treatment. These are not naturally contained in a semiclassical model for the interaction with the electromagnetic field. However, it is possible to model them by adding a non-Hermitian term to the Hamiltonian [50].
Under these considerations, the time-dependent perturbation expression in Eq. (7) becomes [51]:
(36) | ||||
where is the natural linewidth of the state .
Non-Hermitian Hamiltonians describe dissipative systems. The dissipation leads to an increase in entropy, which in turn defines an arrow of time. This effectively breaks symmetry without the need for a fundamental symmetry violation. This makes the unambiguous interpretation of -odd observables in such a system difficult.
5.1 Effects on the induced electric dipole moment
We will now illustrate how this effect manifests in the induced moments that we derived in the paper.
After evaluating Eq. (36) analogously to what we did in sec. 2.3.2, we get:
This leads to the following expression for the induced dipole moment:
Here, we defined more general expressions for the polarizabilities that are valid for non-zero natural linewidths:
If we take the static limit , we get:
Therefore, in the case of static fields, everything considered in previous sections still holds. The only difference is that the size of all 1st order terms increases (slightly) through the difference in the denominator.
5.2 Implication for time-dependent fields
Going back to the general case , both the imaginary and real part of the polarizabilities couple to the fields and their derivatives. We saw in previous sections that the observables for -even and -odd terms are different in that if one of them contains a field, the other necessarily contains its derivative. Now, with the inclusion of the natural linewidth, this is not true anymore. Measuring a certain electromagnetic field response of a system no longer allows us to uniquely distinguish if it was caused by a -odd -even or by a -odd -odd effect because the same field configuration couples both to , and . In the known standard model interactions, the strength of -violating processes far exceeds that of - and -violating ones [26]. Therefore, it is natural to assume the hierarchy . -violating interactions can, therefore, pose a large background for experimental detections of -violating ones. To avoid this, we need the factors to become small. This happens when is far detuned from all the resonances contributing to the polarizability. As we have seen before, this can be achieved by choosing a low frequency. This has the potential disadvantage that also the field derivatives , and become small in this limit. In the other case of very high frequencies, the polarizabilities themselves decrease rapidly.
In Refs. [52, 53], it was shown that oscillating electric fields in neutral atoms and molecules are not well shielded at the nucleus. It was proposed to use this effect to measure nuclear EDMs. The here presented effect of - violation ”mimicking” as -violation through the effects of the natural linewidths could pose an important background for these kinds of measurements.
The situation is even worse for -even -odd observables. Through the same effect, these mix with classical -even, -even potentials, whose size will be many orders of magnitude above the expected size of -odd physics. It is, therefore, advantageous to use static polarizabilities such as to study ToPe effects.
6 Conclusion and outlook
In this work, we investigated how the violation of the symmetries of parity and time reversal leads to atomic and molecular responses to externally applied electric and magnetic fields that are classically forbidden. This led us to general expressions for the induced electric dipole (25), magnetic dipole (26) and electric quadrupole moment (27). We specifically showed how atoms or molecules can obtain electric dipole moments, not only through -violation (sec. 3.2.2), but also through without -violation in time-dependent (sec. 3.3.3, 3.5, 5), and even purely static fields 4.2. Additionally, we derived Eqs. (31), (32) for the energy shift in static fields and Eq. (29) for the effects in time-varying ones. We connected most of these terms to effects that have previously been considered or used in experimental searches for atomic - and - violations (sec. 4.1, 4.2, 4.3). Finally, we argued how the effects of natural linewidths could cause problems in searches of - and - odd effects (sec. 5).
To assess the feasibility of detecting the various terms presented in this work, it is, of course, necessary to evaluate their size as well as the signal-to-noise ratio that can be reached in a given experiment. However, making general statements about them is difficult as they depend heavily on the specific system under consideration. This is the level structure of the atom or molecule, which governs the size of , as well as the specific symmetry-violating effects that one wants to measure. Usually, one considers potentials , , and that involve electron-nucleon contact interactions and, therefore, scale steeply with nucleon numbers, and the overlap of the electrons with the nucleus [49, 27]. Therefore, searches involving very heavy atoms or molecules are usually the best choice [54]. However, there also exist models introducing long-range interactions , , and through new light Bosons that may be easier to restrict in lighter systems [55, 56, 57, 27].
To maximize the size of the polarizabilities, one also wants large , , and transition elements between close-lying states. Such configurations are usually easier to achieve in molecules than in atoms.
We hope that the treatment we presented in this paper can also provide a starting point for further investigations of other symmetry-violating effects that are not captured in our model due to its initial assumptions. For example, with some modifications, one can study or -forbidden dipole transitions. One can also extend the multipole expansion to include the magnetic quadrupole and toroidal moments. One could also include a different class of new physics interactions by introducing , odd potentials that are time-dependent or that couple directly to external fields. Finally, it would also be interesting to understand what happens if the field strength becomes too strong to justify the nondegenerate perturbative treatment that we used.
7 Acknowledgments
This research was financed in whole or in part by
Agence Nationale de la Recherche (ANR) under the project ANR-21-CE30 -0028-01.
A CC-BY public copyright license has been applied by the authors to the present document and will be applied to all subsequent versions up to the Author Accepted Manuscript arising from this submission, in accordance with the grant’s open access conditions.
We thank O. Dulieu for fruitful discussions.
Appendix A The time-reversal operator
As an antiunitary operator, the time reversal operator has some unintuitive properties with respect to the usual bra-ket notation (see, for example, Ref. [58]). For linear Operators , the bra-ket notation is associative in the sense that:
For an antiunitary operator, this is not true. There we have:
For ease of notation, we do not indicate the parentheses explicitly in the rest of the paper. Instead, we define:
For the action of the inverse of the antiunitary time reversal operator onto a follows:
Appendix B Derivation of the , violating polarizability expressions
We now derive the expressions for the polarizabilities that originate from the simultaneous perturbation through the electromagnetic potential and the symmetry-violating potential .
B.1 0st and 1st order in
Because of , we need to solve the system using time-dependent perturbation theory. We can, for example, define the perturbation potential and solve the 2nd-order correction according to Ref. [23]. We ignore here and throughout the paper terms that are quadratic in due to their small size. If we, for now, also ignore terms that are quadratic in , we obtain the following expression for the perturbed time-dependent wavefunction:
where we already removed a term that is proportional to .
From this, we can now calculate the 1st order (in ) perturbation of the , odd induced electric dipole moment:
(37) | ||||
Here, we ignored the term , which vanishes due to parity.
We can see that the solution to this problem not only gives us expressions for the , odd polarizabilities but also leads to a certain type of rest term:
Depending on the time-dependent perturbation treatment, these kinds of rest terms can be absent (see, for example, Ref. [33]). They are, however, necessary for a proper normalization and convergence in the limit of static fields [20, 23].
It is apparent that in the rest term, and are uncorrelated in the sense that their expectation values factorize. For example, in the expression above does not impose any restrictions on the -states that and connect to. For these reasons, it makes sense to view these as products of lower-order terms. As such, they only provide small corrections to already considered observables without introducing the new tensor structures that we saw in the . For these reasons, we ignore these -terms in our treatment.
B.2 2nd order in
In this order of perturbation, we will only consider the expectation values of the , odd Potentials , , and . These terms that are linear in but quadratic in will, therefore, be of the same size as the ones we discussed above. As mentioned before, a time-dependent treatment in this order is more involved and beyond the scope of this paper. We will only discuss the time-independent case. We get, for example, for the perturbation of through , and :
(38) | ||||
This is identical to (16) in the limit . Therefore it follows .
Appendix C Properties of the , violating polarizability expressions
In this appendix, we prove some of the properties of the , odd polarizabilities that we use throughout the paper.
C.1 Behavior of the -terms under time reversal
We now show, on the example of , that the argument presented in sec. 3.3.3 regarding transformation under time reversal of still applies for , odd polarizabilities like .
We start by decomposing into its irreducible components:
(39) | ||||
It is clear that the RHS will transform exactly the same as in Eq. (22). For on the LHS, we can introduce into Eq. (37). After using the transformation behaviors of the different operators given in table 1, we get:
(40) | ||||
Equating this with the transformed version of the RHS of Eq. (39) and renaming , we get:
This tells us that the symmetric part is purely imaginary while the antisymmetric vector component is real.
It is now also clear that when considering instead of , the RHS of Eq. (40) will change by a global minus sign due to the fact that is -odd. That way, we have:
(41) |
C.2 Insensitivity of , , and terms to
In Eq. (25), that describes the full induced electric dipole moment , the only polarizabilities that are quadratic in are classically allowed ones. It is clear from parity reasons that , and vanish. But from the arguments presented in sec. 3.3.3 one would still expect , and to have nonvanishing components. In this section, we show that the terms that are quadratic in a multipole moment possess an additional symmetry that prevents the existence of -odd terms. We demonstrate this on the example of :
From this definition, it follows that . Regarding the tensor decomposition, this relation implies:
(42) |
where we used that is antisymmetric. However, from its transformation under time reversal, we know that , which contains an odd number of -odd operators, must fulfill the relation:
(43) |
Together, these relations can only be true if all three components are zero.
We can make analogous arguments for the polarizabilities that are quadratic in .
In the case of , we are dealing with a rank-4 tensor. Because the quadrupole operator is symmetric, it obeys the relation . One can again see that this relation is in conflict with the relations that follow from time reversal.
In Ref. [12], macroscopic equations describing the possible , , and odd responses of systems to homogeneous time-dependent electromagnetic fields were formulated. From the results of this section, we can see that (under our initial assumptions) there are no microscopic processes that could generate the proposed ToPe responses.
Appendix D Energy shift from static electromagnetic fields
We now derive how the polarizations are related to the energy shift that is induced when static electromagnetic fields are applied to the system. As before, our system is described by . The perturbative energy corrections in such a situation are well known [25, 23] and given by:
As discussed in sec. 3.2, most terms in vanish:
In the next order, we get both the , odd polarizabilities that are linear in (see sec. 3.5) and the , even polarizabilities that are quadratic in (see sec. 2.3.2). For example for , and , we get:
The 3rd energy correction is a sum of two terms. The second one describes the shift from the uncorrelated terms that we discussed in appendix B.1. For the reasons stated before, we neglect these. The first term in gives us the , -odd polarizabilities quadratic in . From the example , and , we get:
From this, we can also see: . If we perform the treatment for the full and , we arrive at a general expression for the energy shift of an atom or molecule exposed to static electromagnetic fields:
References
- [1] M. Safronova, D. Budker, D. DeMille, D. F. J. Kimball, A. Derevianko, and C. W. Clark, “Search for new physics with atoms and molecules,” Reviews of Modern Physics, vol. 90, no. 2, p. 025008, 2018.
- [2] M. S. Safronova, “Searches for new physics,” in Springer Handbook of Atomic, Molecular, and Optical Physics, pp. 471–484, Springer, 2023.
- [3] T. Chupp, P. Fierlinger, M. Ramsey-Musolf, and J. Singh, “Electric dipole moments of atoms, molecules, nuclei, and particles,” Reviews of Modern Physics, vol. 91, no. 1, p. 015001, 2019.
- [4] T. S. Roussy, L. Caldwell, T. Wright, W. B. Cairncross, Y. Shagam, K. B. Ng, N. Schlossberger, S. Y. Park, A. Wang, J. Ye, et al., “An improved bound on the electron’s electric dipole moment,” Science, vol. 381, no. 6653, pp. 46–50, 2023.
- [5] D. Hopkinson and P. Baird, “An interferometric test of time reversal invariance in atoms,” Journal of Physics B: Atomic, Molecular and Optical Physics, vol. 35, no. 5, p. 1307, 2002.
- [6] H. Akdag, B. Kubis, and A. Wirzba, “C and cp violation in effective field theories,” Journal of high energy physics, vol. 2023, no. 6, pp. 1–76, 2023.
- [7] M. Bouchiat and C. Bouchiat, “Parity violation induced by weak neutral currents in atomic physics. part ii,” Journal de Physique, vol. 36, no. 6, pp. 493–509, 1975.
- [8] G. Feinberg, “Parity-violating electromagnetic interactions of nuclei,” Transactions of the New York Academy of Sciences, vol. 38, no. 1 Series II, pp. 26–43, 1977.
- [9] A. Moskalev, “Some macroscopic effects of p-and t-violation in atoms,” in Weak and Electromagnetic Interactions in Nuclei: Proceedings of the International Symposium, Heidelberg, July 1–5, 1986, pp. 638–639, Springer, 1986.
- [10] P. Sandars, “P and/or t violation,” Physica Scripta, vol. 1993, no. T46, p. 16, 1993.
- [11] V. Baryshevsky and D. Baryshevsky, “P-and t-violating spin rotation of an atom (molecule) passing through a laser wave,” Journal of Physics B: Atomic, Molecular and Optical Physics, vol. 27, no. 19, p. 4421, 1994.
- [12] A. Derevianko and M. Kozlov, “Cp-violating magnetic moments of atoms and molecules,” in Advances In Atomic, Molecular, and Optical Physics, vol. 58, pp. 77–112, Elsevier, 2010.
- [13] F. Hund, “Zur deutung der molekelspektren. iii. bemerkungen über das schwingungs-und rotationsspektrum bei molekeln mit mehr als zwei kernen,” Zeitschrift für Physik, vol. 43, pp. 805–826, 1927.
- [14] E. U. Condon, “Theories of optical rotatory power,” Reviews of modern physics, vol. 9, no. 4, p. 432, 1937.
- [15] L. D. Barron and A. D. Buckingham, “Time reversal and molecular properties,” Accounts of Chemical Research, vol. 34, no. 10, pp. 781–789, 2001.
- [16] M. I. Aroyo, International tables for crystallography. Wiley Online Library, 2013.
- [17] H. Schmid, “Some symmetry aspects of ferroics and single phase multiferroics,” Journal of Physics: Condensed Matter, vol. 20, no. 43, p. 434201, 2008.
- [18] D. A. Steck, “Quantum and atom optics,” 2007.
- [19] H. A. Bethe and E. E. Salpeter, Quantum mechanics of one-and two-electron atoms. Springer Science & Business Media, 2012.
- [20] P. Langhoff, S. Epstein, and M. Karplus, “Aspects of time-dependent perturbation theory,” Reviews of Modern Physics, vol. 44, no. 3, p. 602, 1972.
- [21] K. Bhattacharya, “A modification of the dirac time-dependent perturbation theory,” Proceedings of the Royal Society of London. A. Mathematical and Physical Sciences, vol. 394, no. 1807, pp. 345–361, 1984.
- [22] K. Bhattacharyya and D. Mukherjee, “Perturbative quantum dynamics: variants of the dirac method,” Journal of Physics A: Mathematical and General, vol. 19, no. 1, p. 67, 1986.
- [23] A. Mandal and K. L. Hunt, “Adiabatic and nonadiabatic contributions to the energy of a system subject to a time-dependent perturbation: Complete separation and physical interpretation,” The Journal of Chemical Physics, vol. 137, no. 16, 2012.
- [24] D. J. Griffiths and D. F. Schroeter, Introduction to quantum mechanics. Cambridge university press, 2018.
- [25] L. Landau and E. Lishitz, Quantum mechanics: non-relativistic theory, vol. 3. Elsevier, 2013.
- [26] I. B. Khriplovich, “Parity nonconservation in atomic phenomena,” 1991.
- [27] I. B. Khriplovich and S. K. Lamoreaux, CP violation without strangeness: electric dipole moments of particles, atoms, and molecules. Springer Science & Business Media, 2012.
- [28] T.-D. Lee and C. N. Yang, “Elementary particles and weak interactions,” tech. rep., Brookhaven National Lab.(BNL), Upton, NY (United States), 1957.
- [29] D. A. Varshalovich, A. N. Moskalev, and V. K. Khersonskii, Quantum theory of angular momentum. World Scientific, 1988.
- [30] O. Zhizhimov and I. Khriplovich, “P-odd van der waals forces,” Zh. Eksp. Teor. Fiz, vol. 82, pp. 1026–1031, 1982.
- [31] Y. Itin and S. Reches, “Decomposition of third-order constitutive tensors,” Mathematics and Mechanics of Solids, vol. 27, no. 2, pp. 222–249, 2022.
- [32] Y. Itin and F. W. Hehl, “The constitutive tensor of linear elasticity: its decompositions, cauchy relations, null lagrangians, and wave propagation,” Journal of Mathematical Physics, vol. 54, no. 4, 2013.
- [33] R. W. Boyd, A. L. Gaeta, and E. Giese, “Nonlinear optics,” in Springer Handbook of Atomic, Molecular, and Optical Physics, ch. 3, pp. 1097–1110, Springer, 2008.
- [34] L. Schiff and H. Snyder, “Theory of the quadratic zeeman effect,” Physical Review, vol. 55, no. 1, p. 59, 1939.
- [35] J. R. P. Angel and P. Sandars, “The hyperfine structure stark effect i. theory,” Proceedings of the Royal Society of London. Series A. Mathematical and Physical Sciences, vol. 305, no. 1480, pp. 125–138, 1968.
- [36] C. Coulson and M. Stephen, “The quadrupole polarizability of the hydrogen atom,” Proceedings of the Physical Society. Section A, vol. 69, no. 10, p. 777, 1956.
- [37] P. Frantsuzov, I. Khriplovich, and O. Zhizhimov, “Vector polarisability of atomic hydrogen,” Journal of Physics B: Atomic and Molecular Physics, vol. 20, no. 20, p. L655, 1987.
- [38] V. Flambaum, “Long-range parity-nonconserving interaction,” Physical Review A, vol. 45, no. 9, p. 6174, 1992.
- [39] Y. Kim, C.-Y. Liu, S. Lamoreaux, G. Visser, B. Kunkler, A. Matlashov, J. Long, and T. Reddy, “New experimental limit on the electric dipole moment of the electron in a paramagnetic insulator,” Physical Review D, vol. 91, no. 10, p. 102004, 2015.
- [40] S. Eckel, A. Sushkov, and S. Lamoreaux, “Limit on the electron electric dipole moment using paramagnetic ferroelectric eu 0.5 ba 0.5 tio 3,” Physical Review Letters, vol. 109, no. 19, p. 193003, 2012.
- [41] F. L. Shapiro, “Electric dipole moments of elementary particles.,” tech. rep., Joint Inst. for Nuclear Research, Dubna, USSR, 1968.
- [42] R. A. Satten, “Effects of atomic quadrupole moments upon the index of refraction,” The Journal of Chemical Physics, vol. 26, no. 4, pp. 766–772, 1957.
- [43] C. Pryor and F. Wilczek, ““artificial vacuum” for t-violation experiment,” Physics Letters B, vol. 194, no. 1, pp. 137–140, 1987.
- [44] S. Kanorsky, S. Lang, T. Eichler, K. Winkler, and A. Weis, “Quadrupolar deformations of atomic bubbles in solid 4 he,” Physical review letters, vol. 81, no. 2, p. 401, 1998.
- [45] R. Szmytkowski and P. Stefańska, “Magnetic-field-induced electric quadrupole moment in the ground state of the relativistic hydrogenlike atom: Application of the sturmian expansion of the generalized dirac-coulomb green function,” Physical Review A, vol. 85, no. 4, p. 042502, 2012.
- [46] M.-A. Bouchiat and C. Bouchiat, “An atomic linear stark shift violating p but not t arising from the electroweak nuclear anapole moment,” The European Physical Journal D-Atomic, Molecular, Optical and Plasma Physics, vol. 15, no. 1, pp. 5–18, 2001.
- [47] T. Mukhamedjanov, O. Sushkov, and J. Cadogan, “Manifestations of nuclear anapole moments in solid-state nmr,” Physical Review A, vol. 71, no. 1, p. 012107, 2005.
- [48] E. Altuntaş, J. Ammon, S. B. Cahn, and D. DeMille, “Demonstration of a sensitive method to measure nuclear-spin-dependent parity violation,” Physical Review Letters, vol. 120, no. 14, p. 142501, 2018.
- [49] M.-A. Bouchiat and C. Bouchiat, “Parity violation in atoms,” Reports on Progress in Physics, vol. 60, no. 11, p. 1351, 1997.
- [50] J. Sakurai, Advanced quantum mechanics. Pearson Education India, 1967.
- [51] R. Vexiau, D. Borsalino, M. Lepers, A. Orbán, M. Aymar, O. Dulieu, and N. Bouloufa-Maafa, “Dynamic dipole polarizabilities of heteronuclear alkali dimers: optical response, trapping and control of ultracold molecules,” International Reviews in Physical Chemistry, vol. 36, no. 4, pp. 709–750, 2017.
- [52] V. Dzuba, J. Berengut, J. Ginges, and V. Flambaum, “Screening of an oscillating external electric field in atoms,” Physical Review A, vol. 98, no. 4, p. 043411, 2018.
- [53] H. T. Tan, V. Flambaum, and I. Samsonov, “Screening and enhancement of an oscillating electric field in molecules,” Physical Review A, vol. 99, no. 1, p. 013430, 2019.
- [54] G. Arrowsmith-Kron, M. Athanasakis-Kaklamanakis, M. Au, J. Ballof, J. J. Dobaczewski, R. F. Garcia Ruiz, N. R. Hutzler, A. Jayich, W. Nazarewicz, and J. Singh, “Opportunities for fundamental physics research with radioactive molecules,” Reports on Progress in Physics, 2023.
- [55] S. G. Karshenboim, “Precision physics of simple atoms and constraints on a light boson with ultraweak coupling,” Physical review letters, vol. 104, no. 22, p. 220406, 2010.
- [56] M. P. Jones, R. M. Potvliege, and M. Spannowsky, “Probing new physics using rydberg states of atomic hydrogen,” Physical Review Research, vol. 2, no. 1, p. 013244, 2020.
- [57] Y. Stadnik, V. Dzuba, and V. Flambaum, “Improved limits on axionlike-particle-mediated p, t-violating interactions between electrons and nucleons from electric dipole moments of atoms and molecules,” Physical review letters, vol. 120, no. 1, p. 013202, 2018.
- [58] A. Messiah, Quantum mechanics. Courier Corporation, 2014.