Abstract
Schrödinger’s operator relations combined with Einstein’s special relativistic energy-momentum equation produce the linear Klein–Gordon partial differential equation. Here, we extend both the operator relations and the energy-momentum relation to determine new families of nonlinear partial differential relations. The Planck–de Broglie duality principle arises from Planck’s energy expression , de Broglie’s equation for momentum , and Einstein’s special relativity energy, where h is the Planck constant, and are the frequency and wavelength, respectively, of an associated wave having a wave speed . The author has extended these relations to a family that is characterised by a second fundamental constant and underpinned by Lorentz invariant power-law particle energy-momentum expressions. In this note, we apply generalized Schrödinger operator relations and the power-law relations to generate a new family of nonlinear partial differential equations that are characterised by the constant such that corresponds to the Klein–Gordon equation. The resulting partial differential equation is unusual in the sense that it admits a stretching symmetry giving rise to both similarity solutions and simple harmonic travelling waves. Three simple solutions of the partial differential equation are examined including a separable solution, a travelling wave solution, and a similarity solution. A special case of the similarity solution admits zeroth-order Bessel functions as solutions while generally, it reduces to solving a nonlinear first-order ordinary differential equation.
1. Introduction
The classical distinction between waves and particles breaks down at the quantum level, and all matter may exhibit both particle-like and wave-like properties so that matter may behave either as a localised particle or as a spreading wave, depending on the circumstance and how it is observed. Particles occupy a specific location in space and are characterised by their energy e and momentum p and are adequately described by Einstein’s special relativity. On the other hand, waves spread out over space, are characterised by frequency, wavelength, and amplitude, and are described by quantum wave theory for which the energy and momentum variables become operators involving probability density wave functions. Schrödinger’s quantum wave theory is not compatible with special relativity, and the determination of a formal connection between special relativity and quantum mechanics attracted many researchers, including Einstein, resulting in Dirac’s theory and the Klein–Gordon equation (see [,,,]) and culminating in quantum field theory, which provides a fully relativistic description of quantum phenomena (see, for example, [,,]).
The formal relations underpinning the particle-wave duality are the Planck–de Broglie relations for energy and momentum of an elementary particle, where m is the mass and u is the particle velocity. de Broglie [,,] supplemented Planck’s particle energy expression with the equation for particle momentum to establish the dual particle-wave nature of matter, where h is the Planck constant, and are the frequency and wavelength, respectively, of the associated wave having a wave speed (see [,,,] for further articles relating to de Broglie’s work). In [], these basic relations have been generalised to involve a second fundamental constant , and these new relations are shown to apply to the Lorentz invariant power-law energy-momentum relations proposed in [] (see also [] (p. 49)).
Einstein’s special relativistic mass-energy expression together with the Schrödinger operator relations yield the special relativistic Klein–Gordon partial differential Equation (21). In this note, we assume that the particle motion is described by the Lorentz invariant power-law energy-momenta relations derived in [], and for generalised Schrödinger operator relations, we derive a family of nonlinear partial differential Equation (60) arising from one possible operator interpretation of the corresponding algebraic identity. This particular interpretation arises from following a similar development to that for the Klein–Gordon equation. The resulting partial differential Equation (60) is characterised by a stretching symmetry which, in particular, means that there is a similarity solution for which the partial differential equation may be reduced to a second-order ordinary differential equation and which may subsequently be reduced to one of first order. There are other possible operator interpretations which are most likely distinct from one another, one of which gives rise to fractional partial differential equations which we do not consider here.
In the following section, we present preliminary material in four sub-sections for the conventional Schrödinger operator relations including special relativity and Lorentz transformations; the Lorentz invariant extended Newton’s second law; the operator relations based on two Lorentz invariants, and the Klein–Gordon partial differential equation. In the subsequent section, we summarise the Lorentz invariant power-law energy-momenta relations from [], which we suppose apply to an elementary particle. In the next section, we propose compatible operator relations which generalise Schrödinger’s operator relations, which we then use to generalise the Klein–Gordon equation, namely (59) or (60). In the next section, three simple solutions of the resulting nonlinear partial differential equation are examined, including a separable solution, a travelling wave solution, and a similarity solution. The final section of the paper contains a summary and some conclusions, and Appendix A contains some formal details involved in the reduction of the second-order nonlinear Equation (78) to one of first order, namely (93).
2. Schrödinger’s Operator Relations
In this section, we make some general comments relating to the Schrödinger operator relations. In the formulation of Schrödinger’s quantum wave theory, the variables become operators involving wave functions such that for a single spatial dimension x, the wave theory operator relations for momentum p and energy e become, respectively, and , namely
for some function . However, to the author’s knowledge, it appears that with the passing of time, the original motivation and formulation of these specific relations seems to have been lost. For some historical information relating to the formulation of wave mechanics, we refer the reader to [,]. Here, we simply note that with the usual definition of particle velocity , the relations (1) correspond to a certain orthogonality scalar product condition. Thus,
2.1. Special Relativity and Lorentz Transformations
With reference to Figure 1, capital letters refer to the fixed reference frame, and lower-case letters refer to the moving reference frame. If v denotes the constant relative frame velocity, then the Lorentz transformations are given by
with the identity transformation and given by . For the Lorentz invariant energy-momentum relations are given by
and together Equation (4) are referred to as the Lorentz invariant energy-momentum relations corresponding to the Lorentz transformation (3). From each of these relations, the two Lorentz invariants and are apparent and the following basic identities may be verified
and it is a simple matter to show that
from which we might deduce the combined Lorentz invariant
Figure 1.
Two inertial frames moving along the x-axis with relative velocity v.
From a special relativistic perspective, the conventional signatures of the operators (1) are entirely meaningful in the sense of being precisely what is required to produce the correct Lorentz invariances. Further, from Equation (3), we have the chain-rule differential formulae
so that if we adopt the corresponding relations and , then along with and , the transformation differential Formulae (9) are precisely equivalent to the Lorentz invariant energy-momentum relations (4). Furthermore, as shown below, the given signatures of the operators are also precisely those required to ensure the Lorentz invariances of the operators arising from the Lorentz invariants and (see also [] (p. 40) for further discussion on the Lorentz invariants and ).
2.2. Lorentz Invariant Extended Newton’s Second Law
The mechanical model proposed in [,,] formally includes both Newton’s second law and Schrödinger’s quantum wave theory and remains invariant under the combined transformations (3) and (4) (see [] (p. 185)). Thus, for a single spatial dimension x, the proposed equations become simply
where f and g denote certain applied external forces. It can be shown that these equations remain invariant under the Lorentz transformations (3) and the Lorentz invariant energy-momentum relations (4). In other words, the following relations hold:
and these details can be found in [] (p. 185). Importantly, we observe from (1) that the operator relations (10) are such that and , and on setting g as the source of an applied potential, together with the wave equation operator, provides the usual basis for the derivation of Schrödinger’s quantum wave equation [] (p. 282).
2.3. Invariants Invariants and
Following [] (p. 40), we adopt the convention that the operator corresponding to a given variable is subscripted, so that the operators for momentum p and energy e become, respectively,
and from and , we have and defined by
These are Lorentz invariant operators, since from (3) and (9), we may eventually deduce
and consequently, and are Lorentz invariant. The two operators can be shown to commute, namely , which means that the corresponding observables and are simultaneously measurable and that they share the same eigenfunctions (see for example [] (p. 101)). Also, the operator coincides with the Lorentz operator arising from the one-parameter group (3), which is given by
2.4. Klein–Gordon Partial Differential Equation
In this subsection, assuming Einstein’s energy-momentum relation , we show that the Lorentz invariant operators (13) give rise to the Klein–Gordon partial differential Equation (21) arising as the operator equivalent of the algebraic identity . The various properties of the two operators and are most apparent in terms of the characteristic coordinates and , and we comment that in terms of and , the constraint (2) becomes simply .
From the differential formulae
we may deduce
and from which there arises the simple formulae
Thus, the formal operator equation corresponding to the identity or gives rise to the Klein–Gordon Equation (21) ([] (p. 312) or [] (p. 289))
for some function and from which we have
Alternatively, in terms of the conventional wave equation operator, we have the more usual form of the Klein–Gordon equation [] (p. 313):
For the simple harmonic travelling wave, thus
where is a constant, and , the dispersion relation arises from the Klein–Gordon Equation (21) since substitution of (22) into (21) yields the dispersion relation , which, when using the Planck–de Broglie relations, and , where is the frequency and is the wavelength, so that and , is equivalent to the Einstein energy relation .
3. Elementary Particle Power-Law Energy-Momenta Relations
In this paper, for a single space dimension x, we suppose characterises the location x of an elementary particle at time t which is moving with velocity , with energy and momentum , and given by the Lorentz invariant power-law relations
where denotes an arbitrary constant and denotes the particle rest energy. These relations, first obtained in [] (see also [] (p. 49)), extend Einstein’s energy-momentum relations which are included by the case . We comment that the Lorentz invariance of these relations is revealed through the arbitrary constant , which we may give meaning both through the generalised Planck-de Broglie relations (24) for which , as well as through the generalised Schrödinger operator relations (38) for which . As described below (see Equation (35)), this is a less restrictive Lorentz invariance than conventional Lorentz invariance in the sense that it allows for the power-law factors in (23) for and includes conventional special relativity invariance through the special case .
In [], the conventional Planck–de Broglie relations and are generalised by the expressions
where denotes a new fundamental constant and the family of energy expressions (23) bears exactly the same relationship with this generalisation as does Einstein’s energy expression with the Planck–de Broglie formulae where . The dimensions of are those of h, and physically, the relations (24) reflect the parity of space and time in special relativity. Planck’s constant is a space characteristic so that could be referred as “Planck’s constant in time”.
At this point, there are no known experimental values for the constant . However, there is evidence (such as [,] and the numerous references therein) that the fine structure constant, which involves the Planck constant h and the velocity of light c, may not be an absolute constant and may be connected to the expansion of space. This gives rise to the possibility that Planck’s constant may not be strictly constant (see [,]). We further note that the expressions (24) are only novel when the wave is not travelling at the speed of light, since in that case, and (24) become and . This means that experimentally, in order to identify , data from either sub-luminal or superluminal waves must be examined.
In terms of the wave number and angular frequency , these new relations become and , from which we may deduce the relation
where , and .
The relations (23) arise from the following Lorentz invariant relations
and also satisfy
On using , this equation can be alternatively expressed as
which in turn becomes
The invariance of (26) under the Lorentz transformations (3) and (4) may be established either by using Einstein’s addition of velocities law (see [] or [] (p. 49)) or by using the variables and and the Lorentz transformations (4) and (7)2. In terms of these new variables, Equation (26)2 becomes simply
and from (7)2, it is clear that
and that (30) is unchanged by (31) and integrates immediately to give or , consistent with (29).
The important issue is that for , while the two relations (26) are fully Lorentz invariant, it nevertheless constitutes a less restricted Lorentz invariance in the sense that the resulting integral (either (28) or (29)) are not fully Lorentz invariant without the additional assumption that the constant rest energies and vary with the relative frame velocity v. Thus,
This assumption, together with (7)2, means that the actual invariant is
and therefore both the left and right-hand sides of (28) are Lorentz invariants. For conventional special relativity (), the rest energies are such that so that this is not an issue, and the equation is fully Lorentz invariant and provides the formal basis giving rise to the linear Klein–Gordon equation of quantum wave theory. However, for , the correct operator interpretation of either (28) or (29) hinges on the determination of the correct Lorentz invariant formulation.
For , the machinery of Schrödinger’s quantum wave theory is not as meaningful as for in the sense that for , there is no unique operator interpretation of either (28) or (29). Here, because of (33) and both the left and right-hand sides of (28) are Lorentz invariants, we adopt (28) in preference to (29). We exploit (28) and (33) in the form
and on making use of (38) and (41) below, we then follow a similar development to that for to deduce the highly nonlinear second order partial differential Equation (59) or (60). We do not attempt the alternative development based upon (29) because firstly, this version of (28) does not involve natural Lorentz invariants, and secondly, it necessarily gives rise to fractional nonlinear partial differential equations, which are notoriously difficult to solve.
With reference to the power-law expressions (23) and the assumption (32), we comment that from a special relativistic perspective, (32) is an unusual assumption. Physically, it singles out the rest energy as a characteristic that is relative velocity dependent and provides a link to the constant . We further comment that as discussed in [] and [] (p. 52) since , for , there is a directional dependence, which is also inherent in the combined Planck–de Broglie relations . Both [,] suggest that the universe may have a directional dependence so that the properties of the universe may not be isotropic and a preferred direction may exist. Further, in terms of the actual particle energy e, the assumption (32) implies that
which is certainly the case in conventional special relativity and follows immediately from (32) when using the Einstein addition of velocities law in the following form (see, for example, [] (p. 18)):
4. Generalised Schrödinger Operator Relations
In this section, based upon the Lorentz invariant power-law energy-momentum relations (23), we generalise the one-dimensional Schrödinger quantum wave theory operator relations and . We adopt the basic idea suggested in [] that both sub-luminal (particle) and superluminal (wave) characteristics might contribute to any fundamental physical relations. As previously noted, it is only for that the machinery of Schrödinger’s quantum wave theory is particularly meaningful, in the sense that for nonlinear partial differential equations are obtained from (28) rather than the special relativistic linear Klein–Gordon Equation (21).
4.1. Generalised Operator Relations
For a single spatial dimension, on noting the superluminal space-time transformation and , and using
the quantum wave theory operator relations, and , might be extended as follows:
for some function , where and denote arbitrary constants such that conventionally and the constant is pure imaginary, namely . From the relations in (38), we may deduce
where and .
Adopting the definition of particle velocity , we obtain
and the intermediate relation which in terms of and , becomes
consistent with (39). Further, on assuming the extended Newton’s second law (10) we obtain and , and using the relation implies that the assumption (38) is equivalent to the assumption that the applied forces f and g in (10) satisfy .
In terms of the requirement to determine the appropriate operator interpretation of the energy-momentum identities, the proposal (42) below does not make a great deal of sense since the major objective is the determination of a single partial differential equation for a single function . Despite this and as an aside, it may be worth noting that we may use the relations in (38) as an opportunity to introduce a second function such that
and the immediately above relations are not greatly changed. Thus, from
we have the intermediate relation , and in terms of and , this equation becomes
The extended Newton’s second law (10) yields and so that in the absence of further assumptions, there is no implied relation connecting the external applied forces f and g. There is, however, the definite implication that the function is closely associated with the sub-luminal or spatial world, while the function is associated with the superluminal or wave world, which we conventionally understand as the quantum world. In other words, is the wave function arising in Schrödinger’s quantum wave theory.
We further comment that the appropriate extension of the relations (38) to three dimensions is not altogether obvious but might be expected to have the following general structure:
for a certain scalar and vector and where ∇ denotes the usual del-operator, and noting that (38) is recovered from (45) in the case when the vector has the single component . In such a general model, any anisotropy of space might be incorporated with three distinct values corresponding to the three directions of space, namely .
4.2. Some Standard Operator Relations
Following [] (p. 288) and the convention introduced there that the operator corresponding to a given variable is subscripted, the above two operators arising from momentum p and energy e become
We observe that the conventional commutation relations remain the same with replaced by . Thus, for example, we have,
and the remainder follow similarly.
4.3. Invariant Operators and
Now, based upon the Lorentz transformations (3) and (4), the quantities and (see [] (p. 40) for details) are Lorentz invariant. Therefore, we might introduce corresponding Lorentz invariant operators and that are simplest in terms of the characteristic coordinates and . Using the differential formulae
we might deduce the Lorentz invariant operators
and from which we may obtain the commuting operators
The two cases and are necessarily separate, and for , there may be no unique operator interpretation of (27)–(29). Here, we follow the development for using (28) and (34) to obtain (59) and (60). The alternative development based upon (29) results in fractional partial differential equations which we do not attempt here.
4.4. Operator Interpretation for
In this subsection, we show that the Klein–Gordon partial differential Equation (21) formally arises from the generalised Schrödinger operator relations (38) when coupled with Einstein’s energy-momentum relation. If for the time being we formally ignore the condition , then for , Equation (29) becomes , which together with the relations and gives the identity or . This latter identity and the operator relations (50) formally give
for some function and from which we obtain
which is again the Klein–Gordon partial differential Equation (20) or (21), and it is clear that the constants and satisfy the condition and that is pure imaginary when is assumed to be zero.
4.5. Operator Interpretation for
For , in place of the algebraic identity , from (28) and (41), we have instead
so that in place of (51), we have
and on using (41), this equation becomes
On making a term-by-term identification of the generalised operator relations (38) with the generalised Planck–de Broglie formulae given in [],
then it is clear that . If we preserve the relationship , then combining this with gives
which means that we may represent the constants in (55) with a variety of forms. Here, it is simpler to assume only the relation , noting that the key issue for the representations of (55) is the magnitude of . Given the possibilities for and including pure imaginary, the sign of the constant in (59) and (60) may only be determined on a case-by-case basis.
Although not essential, here for definiteness, we assume that and Equation (55) becomes
which is a highly nonlinear partial differential equation that for formally includes the linear Klein–Gordon partial differential Equation (20) or (21). In terms of variables, this equation becomes
Equation (60) is a nonlinear wave equation which overall will exhibit similar physical characteristics to those of other well known wave equations such as the Klein–Gordon equation and the sine–Gordon equation, except that its behaviour is modulated by both the sign and magnitude of the constant K, which is defined by
and by both the sign and magnitude of the factor . The constant is a non-dimensional constant that reflects the relative space-time contributions, in the sense that for small , space or particle characteristics are the dominant contributor, while if is large then time or wave characteristics dominate, and clearly, the case , when there are perfectly balanced contributions, and the prospect of singular behaviour at both arises for the constant K. However, the constant K, which is a dimensional constant with the dimensions of , may assume a wide variety of characteristics depending upon the nature of the constant . For example, if we adopt the values in (58), then K is negative and given by
which has singular behaviour at both . We note that if , then there are a number of subtle changes throughout the subsequent analysis which need to be accommodated. However, the most significant change arises from a sign change in the constant , since we have instead .
5. Some Solutions for
Although (59) is nonlinear, the independent variables and only appear through the partial derivatives; therefore, there are many ad hoc solution techniques that are available (see [,,]). The majority of the techniques described in [] assume a certain functional form involving arbitrary functions and parameters, which are then determined by substitution into the given nonlinear partial differential equation, and the assumed form of the solution may or may not be compatible with the differential equation. Both [,] describe the use of one-parameter transformation groups for the determination of solutions of nonlinear partial differential equations, which, although complicated, is a more systematic solution procedure. In the following three sub-sections, we examine three such solutions. The first two involve ad hoc assumptions, while the third is based upon invariance under a one-parameter group of transformations.
5.1. Separable Solution for
If we assume a separable solution of the form , then on substituting the partial derivatives into (59), we find that and are necessarily determined from the differential relations
where and denote constants such that
On integration of (63), we obtain
where and denote arbitrary constants. Thus, altogether, we find that is given by
where the constants and satisfy the constraint (64). This solution may be confirmed by direct substitution of (66) into (59) and has the alternative form
where and denote further arbitrary constants, noting that depending upon the constants and , this solution may become unbounded. If either or is real, then (67) may become unbounded in either space or time, while if both or are pure imaginary, then (67) is essentially the travelling wave solution examined in the following sub-section.
5.2. Travelling Wave Solution for
Generally, for nonlinear partial differential equations, there are no super-postion principles available, so it is perhaps surprising that the travelling wave solution (22) also applies to (59), which is a consequence of the simple form of (22). The travelling wave solution (22) is a solution of the linear wave equation and generally it would be inapplicable to most nonlinear partial differential equations so that Equation (59) might be regarded as unusual in this respect. However, this does not mean that linear super-position applies to (59). The simple harmonic travelling wave , where is a constant, and , has wave velocity . On using , we obtain from the generalised expressions for energy and momentum
so that the particle velocity u as defined by becomes
and we observe that for , the de Broglie relation emerges, while for large values, the particle velocity coincides with the wave velocity; thus, .
In terms of and , (22) becomes
with partial derivatives
From these expressions and (59), we obtain the following dispersion relation:
It is easy to see why the simple harmonic travelling wave is also a solution, since if we look for general solutions of (60) of the form , where w is the constant wave velocity, we obtain
where K is the constant given by (61), primes here denote differentiation with respect to the argument , and apart from different constants, Equation (73) is essentially that arising from the Klein–Gordon equation. It is also interesting to observe that (72) may also be deduced from the relations and , which are used in the derivation of (25), since from (28), we have
and (72) now follows on noting the relations and .
Many fundamental physical problems are self-similar, and in the following sub-section, we examine a particular similarity solution of (59) arising from a stretching group of invariant transformations.
5.3. Self-Similar Solution for
If a partial differential equation and any associated boundary conditions remain invariant under a one-parameter group of transformations, then generally, there is a major simplification of the problem (see for example []). Here, without reference to possible boundary data, it is easy to see that (59) remains invariant under the following simple one-parameter group of stretching transformations (75):
where is the one-parameter and m denotes any arbitrary constant. Now, since is one invariant of (75), and since m is completely arbitrary, essentially, we may adopt as a second independent invariant so that the similarity solution of (59) takes the form
where denotes a function of to be determined. The functional form of the solution (76) arises by taking one invariant to be a function of the other. Of course, for nonlinear partial differential equations, the major significance of (76) is that if correct, then upon substitution into (59), the partial differential equation will reduce to an ordinary differential equation that is considerably simpler to solve. The arbitrary nature of the constant m relates to the invariance of (60) under arbitrary scaling, or in other words, the same Equation (60) is obtained whatever units are adopted.
On evaluating the partial derivatives for (76), we obtain
where primes denote derivatives with respect to . On substitution of these expressions into (59), we might eventually deduce the following nonlinear ordinary differential equation:
where is the constant defined by
As we have previously noted, the constant is a non-dimensional constant that reflects the relative space-time contributions, such that for small , space dominates, while for large , time is the major contributor, and the two cases arise when the contributions balance. In these cases, the constant becomes singular. The constant , which is a dimensional constant with dimensions of , may assume a variety of characteristics depending on the nature of the constant . For example, if we adopt the values in (58), then itself is pure imaginary and given by so that , which is the constant appearing in (78), is negative.
We observe that for the special case when , (78) becomes simply , which for has the modified Bessel function solutions , where here, and denote arbitrary constants, noting that since and its derivative have logarithmic singularities, the constant may be zero. We further note that for and with a slightly redefined constant , we would obtain the ordinary Bessel function solutions . We further note that for , (78) becomes , which for has the modified Bessel function solutions , where . We comment that Bessel functions appear throughout physics and applied mathematics as a known special function, and their appearance in the present context may or may not be physically significant in terms of establishing a correspondence between widely differing physical phenomena.
We mention briefly and in general terms only one possible class of fundamental problems for which the similarity solution (75) might apply. This is the class of moving boundary problems, which involve two regions with distinct physical characteristics that are separated by a boundary that is in motion. Such problems frequently occur, and more often than not, a constant similarity variable defines the moving boundary. The case is potentially important since this allows the moving boundary to move into a region of constant , say . These problems are difficult to solve analytically, and we mention below some of the issues that may be relevant. For heat-related moving boundary problems, which are referred to as Stefan moving boundary problems involving the determination of moving melting or freezing fronts, we refer the reader to [].
For a specific example, we know that defines the demarcation boundary between the particle and wave worlds, and we might pose the question as to whether or not other such particle-wave boundaries might exist. We have in mind the separate regions, say and , with the common boundary . We suppose that a power-law energy-momentum relation applies in the region , and that the similarity solution (76)
applies in the region . We further suppose that the similarity variable defines the moving boundary through the relation
where is a constant, perhaps yet to be determined, and we use the subscript zero to designate the value along the boundary. In general, this is a complicated implicit equation that defines the moving boundary as a function of time. For , and , while for , we have and .
For any , on taking the total logarithmic derivative with respect to time of the expression in (81), we obtain
and from this basic relation equated to zero, we may deduce that the velocity of the moving boundary is given by the simple expression
which we observe is entirely consistent with the above special cases arising from and . In deriving (83), we are assuming that the velocity of the boundary may be determined by simply assuming the similarity variable to be a constant and then differentiating. The question is whether the particle velocities as defined by and predicted by both the power-law energy momentum relations (23) and from the broader wave theory (in other words from the general expressions (38) or (56) and not just the constant similarity variable) are consistent with the boundary velocity as given by (83). In order for the moving boundary to be well defined from both sides, the particle velocity must be continuous across the boundary, although we may allow jumps in both energy and momentum.
The power-law energy-momentum relations (23) are functions of particle velocity, so that on assuming (83) for the velocity on the boundary, and in conjunction with (81) in the form
we might establish the following boundary values for energy and momentum:
which are evidently consistent with (83) through the definition , as indeed they must.
The question is whether the boundary values for the similarity solution (80) can be arranged so that the predicted velocity on the boundary coincides with (83) for any boundaries other than . On partially differentiating (81), we obtain
so that from the expressions
and the new relations (56), we might eventually deduce
and from these expressions, we may deduce that the velocity u is given by
which is consistent with (83) provided that on the moving boundary , the function satisfies the condition
This is the critical condition that the function must satisfy in order for the particle and wave theories to be meaningful at the common boundary . In principle, since m is arbitrary, it might always be possible to satisfy this condition.
Typically, the two constants and , say in the Bessel function solution for , are determined from a boundary constraint such as along with a jump condition which connects the boundary velocity with jumps in energy and momentum across the boundary. The jump or shock relations are referred to as the Rankine–Hugoniot conditions, and a concise account with examples may be found in [] (page 89). If the constant is set to zero, then generally, this leads to a transcendental equation for the determination of the constant , which is usually solved numerically. However, we do not pursue such matters further here.
As an aside and an interesting related issue, Equation (83) may be re-arranged to give
and we observe that the factor involved in this equation also arises in (26)2, leading to the condition , which together with (26)1 gives the interesting implication that along the moving boundary, the Lorentz invariant is a constant, namely . This interesting outcome leads to us asking whether the second Lorentz invariant is also constant along the moving boundary, which is indeed the case. Again, using both (26)1 and (23a) in conjunction with (83) and (81) in the form (84), we may eventually show that the constant values of and along the moving boundary are given by
giving rise to the condition along the boundary, where as usual refers to the constant rest energy.
Generally, (78) is a highly nonlinear second-order ordinary differential equation. However, it is apparent from (78) that the equation remains invariant under the one-parameter group of stretching transformations , where is the one-parameter. This means that the second-order differential equation may be reduced to one of first order. The formal calculation details are presented in Appendix A.
In brief, if we adopt as the new dependent variable, which we note is an invariant of the stretching group . Then, the second-order differential Equation (78) may be shown to reduce to the following first-order differential equation:
Further details are given in Appendix A. We note that although (93) is certainly of first order, it is not one of the readily solvable first-order differential equations. However, it is conceivable that particular values of the two constants m and may give rise to equations that are analytically solvable.
Some details are given in Appendix A for an approximate asymptotic solution to (78) of the form
which applies for small values of (), and where A, B, C, and n denote certain constants which are determined by substitution of (94) into (93) and equating the leading powers. We further comment that the boundary condition (90) for the function , for the function , becomes simply , and assuming that we may use this constant value in the differential Equation (93), we may deduce the following estimate of . Thus, altogether, we have
where is the constant defined by (79).
6. Summary and Conclusions
In this paper, we use the Lorentz invariant power-law relations
where denotes an arbitrary parameter, in conjunction with both the generalisation of the Planck–de Broglie relations and , namely
where denotes a new fundamental constant such that , and the generalisation of the Schrödinger operator relations, namely
where in this context, , to derive the following nonlinear generalisation of the Klein–Gordon equation
We have proposed that the nonlinear partial differential Equation (99) (also (59) and (60)), involving the arbitrary power-law parameter , as a generalisation of the Klein–Gordon Equation (20) or (21), arises from (99) in the special case . The partial differential Equation (99) is based upon the generalisation (98) of the Schrödinger operator relations, which are motivated from the basic idea proposed in [] that sub-luminal (particle) and superluminal (wave) characteristics both contribute to fundamental physical relations. Further, these operator relations are coupled with the Lorentz invariant energy-momentum power-law relations (96) proposed in [] and with the generalised Planck–de Broglie formulae (97) derived in [], where is a new fundamental constant. The generalised Planck–de Broglie formulae are supported by the family of Lorentz invariant energy relations (96) characterised by the arbitrary constant , and in this context, .
For nonlinear partial differential equations, there are no super-position principles available for generating general solutions, and apart from numerical solutions, often, the best that can be achieved analytically are particular solutions satisfying artificially constructed boundary data. The major characteristic of the resulting nonlinear partial differential Equation (99) (or (59) or (60)) is its invariance under a one-parameter group of stretching transformations. This stretching symmetry gives rise to a family of similarity solutions and means that the partial differential equation may be reduced to a nonlinear second-order ordinary differential equation, which may be then reduced to a highly nonlinear first-order ordinary differential equation.
Many of the nonlinear solution techniques proposed in [,,] are available to determine analytical solutions of (99). Here, we have examined three such solutions for , including (67), which arises from assuming a separable solution of the form , where , , , are determined from the partial differential Equation (59). We have also examined the simple travelling wave (70) with dispersion relation (72) or (74), noting that the two derivations of this relation are distinct. We have also noted that the structure of the partial differential Equation (59) must be regarded as unusual in the sense that the travelling wave (22) is the solution of a linear equation and generally would not apply to most nonlinear partial differential equations. Finally, we have examined the similarity solution (76), involving the arbitrary parameter m, which permits the reduction of the partial differential Equation (59) to the second-order ordinary differential Equation (78). The arbitrariness in the parameter m is a consequence of the invariance of (99) under the one-parameter group of transformations (75). We have shown that the special case has zeroth-order Bessel function solutions, while generally for , (78) may be reduced to the highly nonlinear first order ordinary differential Equation (93) for which we have provided in Appendix A some details for an approximate asymptotic solution of the form (94).
In terms of future work and interesting topics that have not been addressed, to date, we have not determined any numerical values from experimental results for the new constant and its relative magnitude with respect to Planck’s constant h. In addition, we have not determined any numerical solutions of either the nonlinear partial differential Equation (99) or the first-order nonlinear ordinary differential Equation (93), nor have we examined in any detail the major physical implications of the derived solutions of (99), which are given in Section 5. The larger questions of deriving the operator relations (38) from a more fundamental approach such as a variational framework or the relationship of (99), if any, with other nonlinear relativistic wave equations, such as the sine–Gordon equation, have also not so far been addressed.
Funding
This research received no external funding.
Data Availability Statement
The data is contained within the article.
Acknowledgments
The author is grateful to three referees who uniformly provided extensive, constructive, and helpful commentary which has materially improved the presentation.
Conflicts of Interest
The author declares no conflicts of interest.
Appendix A. Reduction to First-Order Differential Equation
In this Appendix, we present the formal details for the reduction of (78) to the first-order ordinary differential Equation (93). We first write (78) as
and then as
and noting that the substitution becomes
which together with the invariance of (77) under the one-parameter stretching group identifies a Euler-type differential equation for which we need to introduce the Euler variable so that we have
Equation (A2) now becomes
If we use as a working variable, noting that , then this equation becomes
and on changing the differential variable from to , this equation becomes
and using gives
On performing the differentiation, we obtain
which reduces to the following first-order differential Equation (93):
Although (A10) is not one of the standard solvable first-order ordinary differential equations, for small values of (), we may readily establish a simple approximate asymptotic solution to (A10) of the form , where A and n are constants that are determined by equating leading order terms. We find that and , where we are using the working variables and , noting the relation . In order to obtain meaningful determining equations for the constants, we need to assume an expansion of the form
which applies for small values of (). The constants B and C are determined simply by substitution of (A11) into (A10), making use of and , and then equating the leading powers. The final results for B and C are not particularly insightful, and we present instead some details for the asymptotic expansions from which the expressions in (A13) are determined. Thus,
With the constant A determined from , where n and are given by and , respectively, the determining equations for the constants B and C are found to be as follows:
and from which we may deduce expressions for B and C.
From (A4) and the approximate solution (A11), we obtain the following approximate first-order ordinary differential equation for the solution of as a function of or :
and for which a wide variety of solution procedures are available, including direct integration
where denotes the constant of integration, and also noting that the lower-order equation arising from the assumption generates a standard first-order Bernoulli differential equation with solution , where denotes the integration constant, although we do not pursue these possibilities further here.
Finally, it may be of some general interest to note that even for the case , the first-order ordinary differential Equation (A10) becomes
where . At face value, (A16) is a Riccati differential equation, which presents its own particular challengers. In this case, however, we know that the solution arises from , namely
where , primes here denote differentiation with respect to z, and noting that the solution involves only one essential arbitrary constant which is the ratio . In terms of z, the differential Equation (A16) becomes
and on differentiating (A17c), using the second-order differential equation for the modified Bessel functions, we may confirm that (A18) is correctly satisfied.
References
- Bjorken, J.D.; Drell, S.D. Relativistic Quantum Fields (Pure and Applied Physics); McGraw-Hill: New York, NY, USA, 1965. [Google Scholar]
- Dirac, P.A.M. Lectures on Quantum Mechanics; Courier Dover Publications: New York, NY, USA, 1964. [Google Scholar]
- Dirac, P.A.M. Principles of Quantum Mechanics; Clarendon Press: Oxford, UK, 1981. [Google Scholar]
- Gross, F. Relativistic Quantum Mechanics and Field Theory; John Wiley and Sons: New York, NY, USA, 2008. [Google Scholar]
- Peskin, M.E.; Schroeder, D.V. An Introduction to Quantum Field Theory; Addison-Wesley Advanced Book Program: New York, NY, USA, 1995. [Google Scholar]
- Roman, P. Introduction to Quantum Field Theory; John Wiley and Sons: New York, NY, USA, 1969. [Google Scholar]
- Hagan, K. Particles and Quantum Fields; World Scientific: New York, NY, USA, 2016. [Google Scholar]
- de Broglie, L. Waves and quanta. Nature 1923, 112, 540. [Google Scholar] [CrossRef]
- de Broglie, L. Ondes et quanta. Comptes Rendus 1923, 177, 507–510. [Google Scholar]
- de Broglie, L. Recherches sur la Theorie des Quanta. Ph.D. Thesis, Sorbonne University of Paris, Paris, France, 1924. [Google Scholar]
- Bohm, D.J.; Hiley, B.J. The de Broglie pilot wave theory and the further development of new insights arising out of it. Found. Phys. 1982, 12, 1001–1016. [Google Scholar] [CrossRef]
- Guemez, J.; Fiolhais, M.; Fernandez, L.A. The principle of relativity and the de Broglie relation. Am. J. Phys. 2016, 84, 443–447. [Google Scholar] [CrossRef]
- Houlik, J.M.; Rousseaux, G. “Non-relativistic” kinematics: Particles or waves. arXiv 2010, arXiv:1005.1762. [Google Scholar]
- Weinberger, P. Revisiting Louis de Broglie’s famous 1924 paper in the Philosophical Magazine. Philos. Mag. Lett. 2006, 86, 405–410. [Google Scholar] [CrossRef]
- Hill, J.M. A consistent extension of the Planck-de Broglie relations. Z. Fur Angew. Math. Und Phys. 2025, 76, 239. [Google Scholar] [CrossRef]
- Hill, J.M. Einstein’s energy and space isotropy. Z. Fur Angew. Math. Und Phys. 2022, 73, 65. [Google Scholar] [CrossRef]
- Hill, J.M. Mathematics of Particle-Wave Mechanical Systems; Springer: Berlin/Heidelberg, Germany, 2022; ISBN 978-3-031-19792-5/978-3-031-19793-2. [Google Scholar] [CrossRef]
- Aspect, A.; Villain, J. The birth of wave mechanics (1923–1926). Comptes Rendus Phys. 2017, 18, 583–585. [Google Scholar] [CrossRef]
- Sommerfeld, A. Wave Mechanics; Brose, H.L., Translator; Methuen Co., Ltd.: London, UK, 1930. [Google Scholar]
- Hill, J.M. On the formal origins of dark energy. Z. Fur Angew. Math. Und Phys. 2018, 69, 133–145. [Google Scholar] [CrossRef]
- Hill, J.M. Some further comments on special relativity and dark energy. Z. Fur Angew. Math. Und Phys. 2019, 70, 5–14. [Google Scholar] [CrossRef]
- Clark, H. A First Course in Quantum Mechanics; Van Nostrand Reinhold Co., Ltd.: London, UK, 1982; pp. 313–317. [Google Scholar]
- Migkas, K.; Schellenberger, G.; Reiprich, T.H.; Pacaud, F.; Ramos-Ceja, M.E.; Lovisari, L. Probing cosmic isotropy with a new X-ray galaxy cluster sample through the LX − T scaling relation. Astron. Astrophys. 2020, 636, A15. [Google Scholar] [CrossRef]
- Wilczynska, M.R.; Webb, J.K.; Bainbridge, M.; Barrow, J.D.; Bosman, S.E.; Carswell, R.F.; Dąbrowski, M.P.; Dumont, V.; Lee, C.-C.; Leite, A.C.; et al. Four direct measurements of the fine structure constant 13 billion years ago. Sci. Adv. Astrophys. 2020, 6, 9672. [Google Scholar] [CrossRef] [PubMed]
- Suto, K. The Planck constant was not a universal constant. J. Appl. Math. Phys. 2020, 8, 456–463. [Google Scholar] [CrossRef]
- Barrow, J.D.; Magueijo, J. A contextual Planck parameter and the classical limit in quantum cosmology. Found. Phys. 2021, 51, 22. [Google Scholar] [CrossRef]
- Ames, W.F. Nonlinear Partial Differential Equations in Engineering I; Academic Press: New York, NY, USA, 1965. [Google Scholar]
- Ames, W.F. Nonlinear Partial Differential Equations in Engineering II; Academic Press: New York, NY, USA, 1972. [Google Scholar]
- Hill, J.M. Differential Equations and Group Methods for Scientists and Engineers; Studies in Advanced Mathematics; CRC Press: Boca Raton, FL, USA, 1992; ISBN 0-8493-4442-5. [Google Scholar]
- Hill, J.M. One-Dimensional Stefan Problems: An Introduction; Pitman Monographs and Surveys in Pure and Applied Mathematics; Longman: London, UK, 1987; ISBN 0-582-98823-3. [Google Scholar]
- Howison, S. Practical Applied Mathematics: Modelling, Analysis, Approximation; Cambridge Texts in Applied Mathematics; Cambridge University Press: Cambridge, UK, 2005; ISBN 10-0-521-84274-3. [Google Scholar]
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content. |
© 2025 by the author. Licensee MDPI, Basel, Switzerland. This article is an open access article distributed under the terms and conditions of the Creative Commons Attribution (CC BY) license (https://creativecommons.org/licenses/by/4.0/).