A recently proposed approach for avoiding the ultraviolet divergence of Hamiltonians with particle creation is based on interior-boundary conditions (IBCs). The approach works well in the non-relativistic case, i.e., for the Laplacian operator. Here, we study how the approach can be applied to Dirac operators. While this has successfully been done already in one space dimension, and more generally for codimension-1 boundaries, the situation of point sources in three dimensions corresponds to a codimension-3 boundary. One would expect that, for such a boundary, Dirac operators do not allow for boundary conditions because they are known not to allow for point interactions in 3D, which also correspond to a boundary condition. Indeed, we confirm this expectation here by proving that there is no self-adjoint operator on a (truncated) Fock space that would correspond to a Dirac operator with an IBC at configurations with a particle at the origin. However, we also present a positive result showing that there are self-adjoint operators with an IBC (on the boundary consisting of configurations with a particle at the origin) that are away from those configurations, given by a Dirac operator plus a sufficiently strong Coulomb potential.

Hamiltonians for quantum theories with particle creation and annihilation are often plagued by ultraviolet divergence.1–3 For defining such a Hamiltonian in a rigorous way, one might employ an ultraviolet cutoff corresponding to smearing out the source of particle creation over a positive volume, or in some cases, one can obtain a renormalized Hamiltonian by taking a limit of the cutoff Hamiltonian in which the volume of the source tends to zero.4–7 Another more recent approach is based on interior-boundary conditions (IBCs)8,9 and yields directly (i.e., without taking a limit) a Hamiltonian suitable for a point source. Here, the wave function ψ is a function on the configuration space

Q=N=0Q1N
(1)

of a variable number of particles that can move in the one-particle space Q1, for example

Q1=R3\{0}
(2)

if particles can be created at a point source fixed at the origin 0R3. In this example, the boundary Q of Q consists of configurations with at least one particle at the origin,

Q=N=0(Q1N)=N=0(x1,,xN)(R3)N:xi=0 for  some  i{1,,N}.
(3)

The IBC is a condition on ψ relating values on the boundary and values in the interior, more precisely, relating the values on two configurations q, q′ that differ by the creation (respectively, annihilation) of a particle at the origin,

q=(x1,,xi=0,,xN)Q1N
(4a)

and

q=(x1,,xi1,xi+1,,xN)Q1N1.
(4b)

As shown by Lampart et al.,10 a Hamiltonian with particle creation at the origin can be defined rigorously in this way in the non-relativistic case without UV divergence or the need for renormalization. Our goal here is to examine in what way and to what extent this approach can be extended to the Dirac equation.

There are already two works about IBCs for the Dirac equation: Schmidt et al.11 explored what IBCs for the Dirac equation can look like on a codimension-1 boundary. However, in our case, boundary (3) has codimension 3. Lienert and Nickel12 developed a quantum field theory (QFT) model in one space dimension using Dirac particles and an IBC that allows two particles to merge into one or one particle to split in two. In contrast, we consider here space dimension 3.

We present here a negative result and a positive one. The negative result (Theorem 1) asserts, roughly speaking, that in three space dimensions, there exists no self-adjoint Hamiltonian for the configuration space as in (1)–(3) (or even the truncated one allowing only N = 0 and N = 1) that acts like the free Dirac Hamiltonian away from the boundary but involves particle creation. Put differently, the free Dirac equation cannot be combined with IBCs in three dimensions. The fact is analogous to the known impossibility of point interaction (δ potentials) for the Dirac equation in three dimensions (see, e.g., Refs. 1315, Theorem 1.1), in particular, since the point interaction is described by a boundary condition,16,34 and for an external field that acts nontrivially only at 0R3, the relevant boundary for this boundary condition is precisely (3), or one sector thereof. Our proof of Theorem 1 makes use of the impossibility theorems for point interaction to deduce the impossibility of IBCs in this situation. We formulate some variants of Theorem 1 as Theorems 3 and 4.

The positive result (Theorem 6) concerns a way in which, nevertheless, a self-adjoint Hamiltonian with particle creation can be rigorously defined by means of an IBC and the Dirac equation for the configuration space as in (1)–(3); it is based on adding a potential. Specifically, we show in Theorem 6 that if a Coulomb potential of sufficient strength, centered at the origin and acting on each of the particles in the model, is added to the action of the Hamiltonian away from the boundary, then a self-adjoint version of the Hamiltonian exists that involves an IBC and leads to particle creation and annihilation at the origin, analogously to the non-relativistic case with the Laplacian operator. The IBC is analogous to the known IBCs for codimension-1 boundaries.11 We formulate the result for a truncated Fock space with only the N = 0 and N = 1 sectors.

The results are formulated in detail in Sec. II. They can be expressed as statements about self-adjoint extensions. That is because, when considering a truncated Fock space (NNmax) and configuration space, for functions ψ that vanish in a neighborhood of the boundary in the top sector Nmax and also vanish in all lower sectors, we know how the desired Hamiltonian should act: in the same way as the Dirac operator Hfree of Nmax particles (respectively, with a Coulomb potential). Let D° be the space of these functions (not dense). Thus, the desired Hamiltonian H is a self-adjoint extension of a (symmetric but not closed) operator H° = H|D° = Hfree|D°. (By the way, when we say “self-adjoint,” we always mean that the operator is densely defined.) The no-go result (Theorem 1) will show that H° has only one self-adjoint extension, viz., Hfree; that means that it is not possible to implement particle creation and annihilation at just one point 0, using IBCs or otherwise. Theorem 6 will take H° to include a suitable Coulomb potential and provide a self-adjoint extension (even several ones) featuring particle creation and annihilation. It remains to be seen whether and how IBCs can be employed in more realistic models of relativistic QFTs.

The IBC Hamiltonians provided by Theorem 6 are neither translation invariant nor rotation invariant, but that was only to be expected: the model cannot be translation invariant, given that the source is fixed at the origin, and the emission of a spin-12 particle by a spinless source cannot conserve angular momentum and, thus, cannot be rotation invariant. (An alternative proof is given in Sec. II D.)17 

Here are further comments on the literature. Some forms of interior-boundary conditions were considered, not necessarily with the UV problem in mind, early on by Moshinsky.18–20 Self-adjoint Hamiltonians based on IBCs and the Laplacian were first rigorously defined by Thomas21 and Yafaev.22 Lampart et al.10 extended these results to the full Fock space and showed that the non-relativistic IBC Hamiltonian agrees, up to addition of a constant, with the one obtained through UV cutoff and renormalization. Lampart and Schmidt23 further extended the proofs to moving sources in two space dimensions, and Lampart,24 in three dimensions. Keppeler and Sieber25 studied the one-dimensional non-relativistic case. Bohmian trajectories for IBC Hamiltonians were defined and studied in Refs. 26, 27 and 40. Other kinds of boundary conditions come up in the study of detection times (see Ref. 35). Further studies of IBC Hamiltonians and their properties include Refs. 2833.

In Sec. II, we describe our results. In Sec. III, we collect the proofs.34,35

For comparison, it will be useful to recapitulate some aspects of the non-relativistic case in three dimensions with a single point source fixed at the origin.10 For simplicity, we consider a truncated Fock space of spinless particles,

HnrN=0NmaxHnr(N)N=0NmaxS±L2(R3,C)N
(5)

with ⊕ the orthogonal sum of Hilbert spaces and S±⋯ the image of the symmetrization operator S+ or anti-symmetrization operator S (subscript nr is for “non-relativistic case”). There is a five-parameter family of IBC Hamiltonians; some members of this family can be regarded as involving an external zero-range potential at the origin in addition to the particle source; there is a two-parameter subfamily that can be regarded as involving no such potential, i.e., as being a pure particle source. The remaining parameters are the energy E0 that must be expended for creating a particle and the strength g of particle creation. Let us fix values E0 > 0 and gC\{0} and call the corresponding operator Hnr. Hnr is a self-adjoint operator in Hnr; let Dnr denote its domain of self-adjointness. Functions ψ=(ψ(0),,ψ(Nmax)) in Dnr satisfy the IBC

limr0rψ(N+1)(x1,,xN,rω)=gm2π2N+1ψ(N)(x1,,xN)
(6)

(suitably understood, and with r ↘ 0 meaning the limit from the right) for N = 0, …, Nmax − 1 and every unit vector ωR3. Another operator to compare to is the free Hamiltonian Hnrfree, which acts on functions ψ from its domain

DnrfreeN=0NmaxS±H2(R3N,C)
(7)

(with H2 the second Sobolev space) according to

Hnrfreeψ(N)=j=1NH1nr,jψ(N)
(8)

with H1nr,j the one-particle Hamiltonian

H1nr=E022mΔ
(9)

acting on particle j.

Now, we want to express that away from the boundary, Hnr acts like Hnrfree. However, the IBC enforces that if ψ(N) ≠ 0 for some N ≥ 0, then every higher sector, ψ(N) with N′ > N, must be nonzero (and, in fact, unbounded) in a neighborhood of the boundary Q1N. In contrast, ψ can lie in Dnr if ψ(Nmax) vanishes in a neighborhood of the boundary and ψ(N) = 0 for all N < Nmax; in fact, ψwill lie in Dnr if ψ(Nmax)Cc(Q1Nmax,C) (where Cc means smooth functions with compact support) and ψ(N) = 0 for all N < Nmax,

DnrDnr{0}{0}Cc(Q1Nmax,C).
(10)

Note that since 0 was excluded from Q1, “compact support” entails that the support stays away from the boundary. The symbol ⊕, when applied to sets that are not Hilbert spaces, should be understood as the Cartesian product, resulting in a subset of Hnr. On Dnr, Hnr acts like the free Hamiltonian,

HnrDnr=HnrfreeDnr=:Hnr.
(11)

That is both (Hnr, Dnr) and (Hnrfree,Dnrfree) are self-adjoint extensions of (Hnr,Dnr). Note that Dnr is not a dense subspace of Hnr (whereas Dnr and Dnrfree are). The condition that (Hnr, Dnr) is an extension of (Hnr,Dnr) expresses that in the highest sector, particle creation can occur only at the origin. In passing, we remark that Yafaev22 showed for Nmax = 1 that all self-adjoint extensions of (Hnr,Dnr) belong to the five-parameter family of IBC Hamiltonians (which includes, as a subfamily, the Hamiltonians without particle creation but with point interaction at the origin).

Now, we turn to the Dirac case. We define the truncated Fock space

HN=0NmaxH(N)N=0NmaxS±L2(R3,C4).
(12)

(Although spin-12 particles are fermions in nature, we cover here also the mathematical case of Dirac particles that are bosons.) The one-particle Hamiltonian is the Dirac Hamiltonian

H1=icα+mc2β
(13)

with mass m ≥ 0, which is self-adjoint on D1=H1(R3,C4) (first Sobolev space). Let HNfree be the free N-particle Hamiltonian, which acts according to

HNfreeψ=j=1NH1jψ,
(14)

D(HNfree) its domain of self-adjointness in L2(R3N,(C4)N), and

Hfreeψ(N)=HNfreeψ(N)
(15)

on the domain

Dfree=N=0NmaxS±D(HNfree).
(16)

Hfree is self-adjoint and is the (truncated) “second quantization” of H1 in H.

The desired IBC Hamiltonian (H, D), or, in fact, any Hamiltonian that agrees with Hfree except for particle creation and annihilation at the origin, must be a self-adjoint extension of (H°, D°) with

D°={0}{0}S±CcQ1Nmax,(C4)Nmax
(17)

and

H°HfreeD°.
(18)

Let H(<Nmax)H(0)H(Nmax1).

Theorem 1.

LetNmax > 0. For every self-adjoint extension (H, D) of (H°, D°), the highest sector decouples from the other sectors; that is,His block diagonal with respect to the decompositionH=H(<Nmax)H(Nmax).

We give all proofs in Sec. III. To paraphrase the conclusion of Theorem 1, the time evolution generated by H involves no exchange between ψ(Nmax) and the other sectors; no particle creation or annihilation occurs toward or from the Nmax-sector; in particular, ψ(Nmax) is time independent. So, Theorem 1 implies that there is no IBC Hamiltonian for the Dirac equation in three space dimensions, as long as no further element such as potentials, space–time curvature, or other particles is introduced.

Theorem 1 is obtained by combining two theorems, Theorem 2 and Theorem 3. The former is a specialized form of a theorem of Svendsen.14 

Theorem 2.

LetN ≥ 1 andMR3Nthe union of finitely manyCsubmanifolds of equal codimensionc. LetH\Mbe the restriction ofHNfreetoCc(R3N\M,(C4)N). Then,H\Mis essentially self-adjoint inL2(R3N,(C4)N)if and only ifc ≥ 2. In particular, the restriction ofHNfreetoCc(Q1N,(C4)N)is essentially self-adjoint, and its restriction toS±Cc(Q1N,(C4)N)is essentially self-adjoint inS±L2(R3N,(C4)N).

The last sentence follows by taking

M=(x1,,xN)R3N:xi=0 for  some  i
(19a)
=i=1N(x1,,xN)R3N:xi=0,
(19b)

so c = 3 and R3N\M=Q1N. This theorem excludes point interaction for the free Dirac equation in three space dimensions. Note, however, that Theorem 1 is not a direct corollary of Theorem 2 because our Hilbert space H is not L2(R3N,Ck) but contains further sectors, and D° is not dense. What we need is the following statement, a kind of generalization of Theorem 1.

Theorem 3.
LetNmax > 0, let(H̃,D̃)be essentially self-adjoint inH(Nmax), and let now
D°{0}{0}D̃H
(20)
andH°:D°Hbe given by
H°0,,0,ψ(Nmax)0,,0,H̃ψ(Nmax).
(21)
For every self-adjoint extension (H, D) of (H°, D°), the highest sector decouples from the other sectors; in fact, with respect to the decompositionH=H(<Nmax)H(Nmax),D=D(<Nmax)D(Nmax)andHis block diagonal with blocksH(<Nmax)andH(Nmax), where(H(Nmax),D(Nmax))is the unique self-adjoint extension of(H̃,D̃).

In this section and Sec. II C, we focus on the case of two sectors, i.e., Nmax = 1. Our positive result is about examples of Dirac Hamiltonians in 3D with particle creation by means of IBCs. These Hamiltonians are based on the one-particle Hamiltonian

H1=icα+mc2β+ν|x|,νcq,
(22)

which consists of the free Dirac Hamiltonian plus a Coulomb potential of strength q (i.e., q is the product of the charge at x and the charge at the origin). We will show in Theorem 6 that for 3/2<|q|<1, there exist IBC Hamiltonians. We conjecture that also for |q| ≥ 1, IBC Hamiltonians exist. On the other hand, the following theorem, a generalization of Theorem 1 in the case Nmax = 1, shows that for |q|3/2, no IBC Hamiltonian exists.

Theorem 4.
LetH=CL2(R3,C4),D°={0}Cc(Q1,C4), and letH° act onψ = (0, ψ(1)) ∈ D° likeH1as in(22),
H°ψ=0,H1ψ(1)
(23)
with|q|3/2. For every self-adjoint extension (H, D) of (H°, D°), the two sectors decouple; that isHis block diagonal. In fact,D=CDH1̄withH1̄the closure ofH1, and forψD,
Hψ=E00ψ(0),H1̄ψ(1)
(24)
with some constantE00R.

It is known (e.g., Ref. 36, Proposition A1) that for |q|<3/2, DH1̄=H1(R3,C4) (first Sobolev space), whereas for |q|=3/2, the domain is bigger than the first Sobolev space. Theorem 4 follows by means of Theorem 3 from the following known theorem (Ref. 37, Theorem 6.9).38 

Theorem 5.

InL2(R3,C4), the operatorH1as in(22)is essentially self-adjoint onCc(Q1,C4)if and only if|q|3/2.

Remark 1.
It is also known15 that
H1=icα+mc2β+V(x)
(25)
with a matrix-valued potential V such that each component Vij satisfies the bound
|Vij(x)|qc|x|+b
(26)
with constants b > 0 and 0 < q ≤ 1/2 is essentially self-adjoint. By Theorem 3, Theorem 4 still applies if (22) is replaced by (25) and (26) with 0 < q ≤ 1/2.
Cassano and Pizzichillo39 improved upon (26) in the case of general Coulomb-type potentials
V(x)=1|x|ν+μβ+λiαx|x|β=:Vel(x)+Vsc(x)+Vam(x),
(27)
but allowing for an additional scalar and anomalous magnetic contribution. More precisely, by exploiting the behavior of functions in the domain at the origin, they showed that H1 in (25) with V given by (27) is essentially self-adjoint if and only if
minkZ\{0}(ck+λ)2+μ2ν2c22,
which reproduces Theorem 5 for μ = λ = 0.

It will be helpful again to consider first the non-relativistic IBC Hamiltonian (Hnr, Dnr), now for Nmax = 1, so Hnr=CL2(R3,C). We report a few facts:10 for every ψDnr, the upper sector is of the form

ψ(1)(x)=c1|x|1+c0|x|0+o(|x|0)
(28)

as x0, with (uniquely defined) “short distance coefficients” c1,c0C. ψDnr satisfies the IBC (6), which can be written in the form

c1=gψ(0),
(29)

and the Hamiltonian acts on ψ like

(Hnrψ)(0)=g*c0,
(30a)
(Hnrψ)(1)(x)=E022mΔ ψ(1)(x)for x0.
(30b)

[At x = 0, Δψ(1), when understood in the distributional sense, includes a delta distribution stemming from the |x|−1 contribution in (28).]

Now, we turn again to the Dirac case with

H=CL2(R3,C4)
(31)

and one-particle operator H1 as in (22). In the following, we take H1 to be an operator on the domain

D1=Cc(Q1,C4)L2(R3,C4)
(32)

with adjoint (H1*,D1*) in L2(R3,C4).

The notation

Φmj,κj±
(33)

is common for certain functions that form an orthonormal basis of L2(S2,C4) (where S2 means the unit sphere in R3) and are simultaneous eigenvectors of J2, K, J3 with J = L + S the total angular momentum and K = β(2S · L + 1) the “spin–orbit operator.” Their explicit definition in terms of spherical harmonics can be found in, e.g., Ref. 15, p. 126. The symbols mj and κj are the traditional names of their indices.

Theorem 6.
LetH=CL2(R3,C4),D°={0}Cc(Q1,C4), andH° be given by(23)with3/2<|q|<1. SetB1q2, and note that0<B<12. ChoosegC\{0}and
(m̃j,κ̃j)A12,1,12,1,12,1,12,1.
(34)
Then, there is a self-adjoint extension (H, D) of (H°, D°) with the following properties:
  1. Particle creation occurs; i.e., the two sectors do not decouple [His not block diagonal in the decomposition(31)].

  2. For everyψD, the upper sector is of the form
    ψ(1)(x)=cfm̃jκ̃jx|x||x|1B+(mj,κj)Ac+mjκjfmjκj+x|x||x|1+B+o(|x|1/2)
    (35)
    asx0with (uniquely defined) short distance coefficientsc,c+mjκjCand particular functionsfmjκj±:S2C4given by
    fmjκj+=(1+qB)Φmjκj+(1+q+B)Φmjκj,
    (36a)
    fmjκj=(1+q+B)Φmjκj+(1+qB)Φmjκj.
    (36b)
  3. EveryψDobeys the IBC
    c=gψ(0),
    (37)
    andHacts onψDaccording to
    (Hψ)(0)=g*4B(1+q)c+m̃jκ̃j,
    (38a)
    (Hψ)(1)(x)=icα+mc2β+q|x|ψ(1)(x)(x0).
    (38b)

Equivalently, the IBC (37) could be written as

limr0r1+Bψ(1)(rω)=gfm̃jκ̃j(ω)ψ(0)
(39)

for every ωS2.

Remark 2.

It seems plausible that an analogous IBC Hamiltonian can be set up for Nmax > 1 by applying the same terms on each particle sector.

Remark 3.
The IBC Hamiltonian described in Theorem 6 belongs to a whole family of IBC Hamiltonians in which (37) and (38a) are replaced by
a1c+a2c+m̃jκ̃j=gψ(0)
(40)
and
(Hψ)(0)=g*(a3c+a4c+m̃jκ̃j),
(41)
respectively, with real constants a1, …, a4 satisfying
a1a4a2a3=4B(1+q).
(42)
As we will show in the Proof of Theorem 6, each of these choices defines a self-adjoint operator. The Hamiltonian defined through (37) and (38a) corresponds to a1 = 1, a2 = 0, a3 = 0, and a4 = 4B(1 + q).

Remark 4.

In our construction of H, one angular momentum sector of L2(R3,C4), indexed by (m̃j,κ̃j), gets coupled to the 0-particle sector. If we took the 0-particle sector to have more than one dimension, say H(0)=C4 instead of C, then we could couple several angular momentum sectors listed in (34) to several mutually orthogonal subspaces of the 0-particle sector H(0). However, this mathematical possibility does not seem physically natural.

Remark 5.

It would be interesting to investigate also the case |q| ≥ 1. We expect that IBC Hamiltonians exist also in that case because hmjκj (an angular momentum block of H1, see Sec. III) is known to have multiple self-adjoint extensions also in that case. However, we do not have a proof.

Remark 6.

It is also of interest to define a |ψ|2-distributed jump process for the Bohmian configuration in analogy to the processes defined in Ref. 26 for non-relativistic IBC Hamiltonians. This issue is addressed in a separate work.40 

The Hamiltonian H provided by Theorem 6 is not rotationally symmetric, for example, because the subspace Km̃jκ̃j that plays a special role for H is not invariant under J1 or J2, and, thus, not under the action of the rotation group (more precisely, of its covering group). One might think of coupling all four angular momentum sectors Kmjκj with (mj,κj)A to the 0-particle sector in a symmetric way, but actually, that does not help.

Theorem 7.

LetHand (H°, D°) be as in Theorems 4 and 6 with anyqR. None of the self-adjoint extensions of (H°, D°) with particle creation is rotationally symmetric.41 

The proof is based on the following fact that we also prove in Sec. III.

Lemma 1.

The only vector inL2(R3,C4)invariant under rotations [i.e., under the representation of the covering group ofSO(3)] is the zero vector.

Proof of Theorem 1.

Let D̃=S±CcQ1Nmax,(C4)Nmax and H̃ the restriction of HNfree to D̃. By Theorem 2, (H̃,D̃) is essentially self-adjoint in H(Nmax). Now, Theorem 3 yields the statement of Theorem 1.□

Proof of Theorem 3.
As a shorthand notation, we write
H<H(<Nmax),H=H(Nmax),
(43)
and correspondingly, H<, D<, H=, and D=. Let Γ(A) denote the graph of an operator [A, D(A)] and Γ(A)̄ its closure in HH, which is the graph of the closure of A, ΓĀ. Since H̃ is essentially self-adjoint, its closure H̃̄ is its self-adjoint extension H=. Since (H°, D°) is symmetric (and H̃ is densely defined), it is closable with closure H°̄,DH°̄, and we get that
Γ(H)=Γ(H)̄Γ(H°)̄=0f0H̃f:fD̃̄=0f0g:fgΓ(H̃)̄
(44a)
=0f0g:fgΓ(H=)=0f0H=f:fD=,
(44b)
where 0 means the zero of H<. In particular,
DDH°̄={0}D=.
(45)
Now, define the “adjoint domain” of the not densely defined operator H°̄ as
D*H°̄=ϕH:ηHψDH°̄:ϕ|H°̄ψH=η|ψH
(46)
and note that
DD*H°̄.
(47)
By (45), ψ in (46) is of the form (0, f) with fD=, so (writing ϕ= = g and η= = h)
D*(H°̄)=H<gH=:hH=fD=:g|H=fH==h|fH=
(48a)
=H<D=
(48b)
by self-adjointness of H=. We, thus, obtain the chain of inclusions
{0}D=DH<D=.
(49)
This entails further that
D=D<D= with D<ψ<:ψ<ψ=D.
(50)
Indeed, writing vectors now as rows, for any (ψ<, ψ=) ∈ D, we have that ψ<D< by definition of D< and ψ=D= by (49), so (ψ<, ψ=) ∈ D<D=. Conversely, if ψ<D< and ψ=D=, then by definition of D<, there is ϕ=H= such that (ψ<, ϕ=) ∈ D, but then, ϕ=D= by (49), so ψ=ϕ=D= since D= is a subspace, so (0, ψ=ϕ=) ∈ D by (49), so (ψ<, ψ=) = (ψ<, ϕ=) + (0, ψ=ϕ=) ∈ D since D is a subspace.
Now, we turn to the action of the operator H and claim that for every ϕ = (ϕ<, ϕ=) ∈ D<D=,
(Hϕ)==H=ϕ=
(51)
regardless of the choice of ϕ<D<. Indeed, from (47), we obtain that
ϕ<ϕ==ϕDη<η==ηHψ<ψ==ψD(H°̄):(Hϕ)=|ψ=H==ψ<=0Hϕ|ψH=H s.a.ϕ|HψH=H°̄Hϕ|H°̄ψH=46η|ψH=ψ<=0η=|ψ=H==48aϕ=|H=ψ=H==H= s.a.H=ϕ=|ψ=H=.
(52)
By (45), (Hϕ)=|fH==H=ϕ=|fH= for every fD=. Since D= is dense, (51) follows.
Now, it follows further from (51) that there is an operator H<:D<H< such that
(Hϕ)<=H<ϕ<
(53)
regardless of ϕ=. Indeed, by setting ϕ= = 0, we obtain from (51) that (H(ϕ<,0))==0 and define
Hϕ<0=H<ϕ<0.
(54)
From (44), it follows that
H0ϕ==0H=ϕ=.
(55)
Thus,
Hϕ<ϕ==Hϕ<0+H0ϕ==H<ϕ<0+0H=ϕ==H<ϕ<H=ϕ=
(56)
for all ϕD = D<D=; that is H is block diagonal. This completes the proof.□

Proof of Theorem 4.

Use Theorems 3 and 5 with Nmax = 1, H̃=H1 as in (22), D̃=Cc(R3\{0},C4). Since H<=C, the only possibilities for D< are D< = {0} and D<=C, and since D is dense, only D<=C remains. In particular, H< is multiplication by some real constant E00.□

Proof of Theorem 6.
We begin by reviewing the well known decomposition of Hilbert space H(1)=L2(R3,C4) in terms of the functions Φmj,κj±. By passing to spherical coordinates and denoting with d2ω the surface measure of the unit sphere S2, we obtain the canonical isomorphism
U:H(1)L2((0,),C,dr)L2(S2,C4,d2ω)
(57)
by setting, for each ψH(1),
(Uψ)(r,ω)=rψ(rω).
(58)
Under this transformation, the Dirac–Coulomb operator H1 as in (22) takes the form (Ref. 15, p. 125)
UH1U=iαrr+1r1rβK+mβ+qr,
(59)
where αr = er · α is the radial component of α and K = β(2S · L + 1) is the spin–orbit operator consisting of the spin operator S=i4α×α and the angular momentum operator L = x × (−i∇). Denoting with J = L + S the total angular momentum, one can show15 that K commutes with J2 and with the third component J3 of J. The functions Φmj,κj± form a joint eigenbasis of J2, K, and J3 with eigenvalues j(j + 1), κj, and mj, respectively, thus providing the following orthogonal decomposition:
L2(S2,C4,dΩ)=jN0+12mj=jjκj=±j+12Kmjκj
(60)
with
Kmjκj=span(Φmjκj+,Φmjκj).
(61)
In this basis, we have that (Ref. 15, Lemma 4.13)
αr=0ii0,β=1001,iαrβ=0110,
(62)
and we quote the following fact.

Lemma 2.
(Ref.15 , Theorem 4.14).Uas in(58)mapsCc(Q1,C4)to (but not onto)
j,mj,κjCc(0,),C2Kmjκj,
(63)
and (the closure of)H1is unitarily equivalent to (the closure of)UH1U, which is block diagonal with respect to(63)with blocks
hmj,κj=m+qrr+κjrr+κjrm+qr.
(64)

In the following, we will say “angular momentum sector” to Kmjκj. Our construction of the IBC Hamiltonian H proceeds for each angular momentum sector separately. For one sector, the one chosen in (34), we will couple hm̃jκ̃j to the 0-particle sector of our mini-Fock space H; all other angular momentum sectors will decouple. That is, H will be block diagonal relative to the sum decomposition

HĤ(j,mj,κj)(ȷ̃,m̃j,κ̃j)L2((0,))Kmjκj
(65)

(note that j is determined by κj through j=|κj|12), but not relative to

Ĥ=H(0)L2((0,))Km̃jκ̃j.
(66)

To this end, we need a self-adjoint extension for every hmjκj with (mj,κj)(m̃j,κ̃j).

Lemma 3
(Refs.42 and43 , Propositions 1.2 and 2.2–2.4). The operator[hmjκj,Cc((0,))Kmjκj]is essentially self-adjoint if and only ifq2κj214. As a consequence, for3/2<|q|<1, the only angular momentum sectors for whichhmjκjis not essentially self-adjoint are those mentioned in(34). Furthermore, for3/2<|q|<1, those sectors that are not essentially self-adjoint possess, among an infinitude of self-adjoint extensions, a distinguished one,hDmjκjthat is uniquely characterized by the property that for all functionsϕin the domain, kinetic and potential energy are separately finite,|hDmjκjqr|1/2ϕ<and(|q|r)1/2ϕ<. Functionsϕin the domain of the distinguished extension obey the asymptotics
ϕ(r,ω)=c+mjκjfBmjκj+(ω)rB+o(r1/2)
(67)
as r ↘ 0.

For the three sectors in A mentioned in (34) but different from (m̃j,κ̃j), we choose the distinguished extension, and for those not mentioned there (i.e., with |κj| ≥ 2), the extension is unique. These extensions can be combined to form an extension of H° on all summands but the first in (65). So, it remains to construct the block Ĥ of H acting in Ĥ, which is where the coupling between particle sectors takes place. Correspondingly,

DD̂DdistinguishedĤĤ=H
(68)

with D̂ to be determined. For brevity, we set

KKm̃jκ̃j and hhm̃jκ̃j.
(69)

We regard h as defined on Cc((0,))K. Another known fact is the following.

Lemma 4
(Ref.43 , Theorem 2.6). Every functionϕin the domainD(h*) of the adjointh* ofhobeys, asr ↘ 0, the asymptotics44 
ϕ(r,ω)=cfm̃jκ̃j(ω)rB+c+fm̃jκ̃j+(ω)rB+o(r1/2)
(70)
withf±as in(36); the coefficientsc,c+=c+m̃jκ̃jCare uniquely determined byϕ, and all combinations(c,c+)C2occur for someϕ.

For higher angular momentum sectors (|κj| ≥ 2), functions in the domain of the unique self-adjoint extension of hmjκj are o(r1/2) [because the extension is hmjκj̄, and the domain of that contains only o(r1/2) functions by Ref. 43, Proposition 2.4] and, thus, contribute only o(|x|−1/2) in (35).

We will directly consider the more general form (40) of the IBC, which contains (37) as a special case. We define the domain D̂ of Ĥ as follows, containing functions satisfying the IBC (40)

D̂(ψ(0),ψ(1))Ĥ:ψ(1)D(h*) and (40).
(71)

We define the action of Ĥ [according to (38b) and (41)] as

(Ĥψ)(0)=g*(a3c+a4c+m̃jκ̃j),
(72a)
(Ĥψ)(1)=h*ψ(1).
(72b)

D̂ is dense in Ĥ because (i) in D(h*), there exist functions with arbitrary values of c and c+, so for ψD̂, any desired complex number can occur as ψ(0); (ii) since D(h)=Cc((0,))K is contained in D(h*) and dense in L2((0,))K, the set {(ψ(0),ψ(1)+ϕ):ϕD(h)}D̂ is dense in {ψ(0)}L2((0,))K. Together, (i) and (ii) imply that D̂ is dense.

We now prove that Ĥ is a symmetric operator on D̂. Let ψD̂ with short distance coefficients c and c+, and ϕD̂ with short distance coefficients d and d+. Then,

ϕ,HψĤHϕ,ψĤ=ϕ(0),g*(a3c+a4c+)C+ϕ(1),h*ψ(1)L2((0,),K)g*(a3d+a4d+),ψ(0)Ch*ϕ(1),ψ(1)L2((0,),K)
(73)
=a1d+a2d+,a3c+a4c+Ca3d+a4d+,a1c+a2c+C+ϕ(1),h*ψ(1)L2((0,),K)h*ϕ(1),ψ(1)L2((0,),K).
(74)

Now, note that h* can be written as h* = −rr + M(r), where M(r) is a Hermitian operator KK for each r > 0. Thus, the last two terms of (74) can be written as

ϕ(1),h*ψ(1)L2((0,),K)h*ϕ(1),ψ(1)L2((0,),K)=0drϕ(1)(r),(iαr)rψ(1)(r)K+rϕ(1)(r),(iαr)ψ(1)(r)K
(75)
=0drrϕ(1)(r),(iαr)ψ(1)(r)K
(76)
=limr0ϕ(1)(r),(+iαr)ψ(1)(r)K
(77)
=limr0dfrB+d+f+rB+o(r1/2),c(iαr)frB+c+(iαr)f+rB+o(r1/2)K
(78)
=limr0d*cf,iαrfKr2B+d*c+f,iαrf+K+d+*cf+,iαrfKr0+o(r0).
(79)

Since in the orthonormal basis Φ±=Φm̃jκ̃j± of K, the explicit form of αr is given by (62), any vector v=xΦ++yΦK with coefficients x,yC has

v,iαrvK=(x*y*)0110xy=x*yy*x=2Im(x*y).
(80)

Since by (36), f± has real coefficients relative to Φ±, we have that

f±,iαrf±K=0.
(81)

Thus, the r−2B term in (79) vanishes. Moreover, by (36) and (62) again,

f,iαrf+K=4B(1+q),f+,iαrfK=4B(1+q).
(82)

Thus,

ϕ(1),h*ψ(1)L2((0,),K)h*ϕ(1),ψ(1)L2((0,),K)=d+*c4B(1+q)d*c+4B(1+q).
(83)

Therefore, putting together (74) and (83),

ϕ,HψHHϕ,ψH=a1a3d*c+a2a3d+*c+a1a4d*c++a2a4d+*c+a1a3d*ca2a3d*c+a1a4d+*ca2a4d+*c++d+*cd*c+4B(1+q)
(84)
=(d+*cd*c+)(a1a4+a2a3+4B(1+q))
(85)
=0
(86)

if (42) holds.

In order to see that Ĥ is also self-adjoint, it remains to verify that D̂=DĤ*. To this end, we first note that

D̂DĤ*CD(h*).
(87)

Any given ϕCD(h*) lies in DĤ* if and only if there is ηĤ such that for every ψD̂,

η,ψĤ=ϕ,ĤψĤ.
(88)

If we write again c, c+ for the short-distance coefficients of ψ(1) and d, d+ for those of ϕ(1), then condition (88) is equivalent to

η,ψĤ=ϕ(0),(Ĥψ)(0)C+ϕ(1),h*ψ(1)L2((0,),K)
(89)
=(41),(83)ϕ(0)*g*(a3c+a4c+)+h*ϕ(1),ψ(1)L2((0,),K)(d*c+d+*c)4B(1+q)
(90)
=(42)gϕ(0)*(a3c+a4c+)+h*ϕ(1),ψ(1)L2((0,),K)d*d+*cc+a1a3a2a4
(91)
=gϕ(0)*(a3c+a4c+)+h*ϕ(1),ψ(1)L2((0,),K)d*a1+d+*a2d*a3+d+*a4ca1+c+a2ca3+c+a4
(92)
=(a1d+a2d+)+gϕ(0)*(a3c+a4c+)+h*ϕ(1),ψ(1)L2((0,),K)+a3d*+a4d+*a1c+a2c+
(93)
=(a1d+a2d+)+gϕ(0)*(a3c+a4c+)+h*ϕ(1),ψ(1)L2((0,),K)+g*(a3d+a4d+),ψ(0)C.
(94)

The only way this can be true for all ψD̂ is that

η(0)=g*(a3d+a4d+),
(95a)
η(1)=h*ϕ(1),
(95b)

and the product in the first line of (94) vanishes. For one thing, we obtain from this an expression for Ĥ*ϕ=η. Moreover, since the product in the first line of (94) consists of one factor depending on ϕ and one depending on ψ, and since the product needs to vanish for every ψD̂ but the second factor will not, the first factor has to vanish for every ϕD(Ĥ*). As a consequence, ϕ also needs to satisfy the IBC (40), i.e.,

a1d+a2d+=gϕ(0),
(96)

so that we arrive at

D̂=DĤ*
(97)

and Ĥ (and, thus, H) is self-adjoint. This completes the proof of Theorem 6.□

Remark 7.

Here is an alternative argument for the part after (86), after it has been shown that H is symmetric. While this alternative argument does not show that H is self-adjoint, it shows that H possesses a self-adjoint extension, which suffices for proving Theorem 6. The argument is based on the von Neumann theorem about conjugations (Ref. 45, Theorem X.3), which asserts the following: letHbe a Hilbert space. An antilinear mapC:HH[i.e., such thatC(αϕ + βψ) = α* + β*] is called a conjugation if it is norm-preserving andC2 = I. Let [A, D(A)] be a densely defined symmetric operator, and suppose that there exists a conjugationCwithC(D(A)) ⊆ D(A) andAC = CA. Then,Ahas equal deficiency indices and, therefore, has self-adjoint extensions.

In our case, we use that the differential expression (64) for h has only real entries. We assume for the argument that the coupling constant g is real; if this is not the case, it can be arranged through a unitary transformation of H that merely changes the phase of ψ(0) by the phase of g. So, we take C:ĤĤ to be complex conjugation of ψ(0) and of the coefficients of ψ(1)L2((0,))K relative to Φm̃jκ̃j±. Since the coefficients a1, …, a4 are real, H commutes with C, and the von Neumann theorem applies.

Remark 8.

In Ref. 39, the authors established short-distance asymptotics of wave functions for the case of general Coulomb-type potentials (27), very similarly to Lemmas 3 and 4. Having their analog at hand, constructing a similar IBC Hamiltonian for certain parameter ranges of ν,μ,λR from (27) is reasonably straightforward.

Proof of Lemma 1.

Suppose ψ was a nonzero vector in L2(R3,C4) invariant under the representation of the covering group of SO(3). A rotation (about any axis) through 360° is a particular element g of the covering group that acts on vectors x in R3 as the identity and on spinors as −I with I the identity. Thus, g maps ψ to −ψ, and if ψ is invariant, it must vanish.□

Proof of Theorem 7.

If H were a self-adjoint extension of H° that involves particle creation, then the initial Fock vector ψ0=(ψ0(0),ψ0(1))=(1,0) (i.e., the Fock vacuum) would evolve by some time t > 0 to a non-vacuum state, i.e., one with nonzero ψt(1), that would be invariant under rotations. Since rotations do not mix particle number sectors, also ψt(1) by itself would be invariant under rotations. However, by Lemma 1 no such state exists.□

Note added in proof. After the completion of this work, Binz and Lampart posted a preprint46 in which they described a general abstract framework for the construction of IBC Hamiltonians. The one given in Theorem 6 can also be cast in this form.

J.H. gratefully acknowledges the partial financial support by the ERC Advanced Grant “RMTBeyond” under Grant No. 101020331.

The authors have no conflicts to disclose.

Joscha Henheik: Formal analysis (equal); Investigation (equal); Writing – original draft (equal); Writing – review & editing (equal). Roderich Tumulka: Formal analysis (equal); Investigation (equal); Writing – original draft (equal); Writing – review & editing (equal).

Data sharing is not applicable to this article as no new data were created or analyzed in this study.

1.
J.
Dereziński
, “
Van Hove Hamiltonians—Exactly solvable models of the infrared and ultraviolet problem
,”
Ann. Henri Poincare
4
,
713
738
(
2003
).
2.
J. R.
Oppenheimer
, “
Note on the theory of the interaction of field and matter
,”
Phys. Rev.
35
,
461
477
(
1930
).
3.
L.
van Hove
, “
Les difficultés de divergences pour un modèle particulier de champ quantifié
,”
Physica
18
,
145
159
(
1952
).
4.
J.
Glimm
and
A.
Jaffe
,
Quantum Field Theory and Statistical Mechanics
(
Birkhäuser
,
Basel
,
1985
).
5.
J.
Glimm
and
A.
Jaffe
,
Quantum Physics—A Functional Integral Point of View
(
Springer
,
New York
,
1987
).
6.
T. D.
Lee
, “
Some special examples in renormalizable field theory
,”
Phys. Rev.
95
,
1329
1334
(
1954
).
7.
E.
Nelson
, “
Interaction of nonrelativistic particles with a quantized scalar field
,”
J. Math. Phys.
5
,
1190
1197
(
1964
).
8.
S.
Teufel
and
R.
Tumulka
, “
Hamiltonians without ultraviolet divergence for quantum field theories
,”
Quantum Stud.: Math. Found.
8
,
17
35
(
2021
); arXiv:1505.04847.
9.
S.
Teufel
and
R.
Tumulka
, “
Avoiding ultraviolet divergence by means of interior-boundary conditions
,” in
Quantum Mathematical Physics
, edited by
F.
Finster
,
J.
Kleiner
,
C.
Röken
, and
J.
Tolksdorf
(
Birkhäuser
,
Basel
,
2016
), pp.
293
311
; arXiv:1506.00497.
10.
J.
Lampart
,
J.
Schmidt
,
S.
Teufel
, and
R.
Tumulka
, “
Particle creation at a point source by means of interior-boundary conditions
,”
J. Math. Phys., Anal., Geom.
21
,
12
(
2018
); arXiv:1703.04476.
11.
J.
Schmidt
,
S.
Teufel
, and
R.
Tumulka
, “
Interior-boundary conditions for many-body Dirac operators and codimension-1 boundaries
,”
J. Phys. A: Math. Theor.
52
,
295202
(
2019
); arXiv:1811.02947.
12.
M.
Lienert
and
L.
Nickel
, “
Multi-time formulation of particle creation and annihilation via interior-boundary conditions
,”
Rev. Math. Phys.
32
(
2
),
2050004
(
2020
); arXiv:1808.04192.
13.
S.
Albeverio
,
F.
Gesztesy
,
R.
Høegh-Krohn
, and
H.
Holden
,
Solvable Models in Quantum Mechanics
(
Springer
,
Berlin
,
1988
).
14.
E. C.
Svendsen
, “
The effect of submanifolds upon essential self-adjointness and deficiency indices
,”
J. Math. Anal. Appl.
80
,
551
565
(
1981
).
15.
B.
Thaller
,
The Dirac Equation
(
Springer
,
Heidelberg
,
1991
).
16.
H.
Bethe
and
R.
Peierls
, “
Quantum theory of the diplon
,”
Proc. R. Soc. London, Ser. A
148
,
146
156
(
1935
).
17.

More generally, the emission of a spin-12 particle by a source of any spin cannot conserve angular momentum because (i) the creation term in the Hamiltonian morally reads r,r,sgrrsar(x)bs(x)ar(x) with ar the creation operator of the emitting particle, bs the creation operator of the emitted particle, and grrs complex coefficients with s a Dirac spinor index and r and r′ indices referring to another representation space of the rotation group SO(3) or its covering group and (ii) there is no rotation invariant array of coefficients grrs. In contrast, the emission of a spin-1 particle by a source of spin 12can conserve angular momentum, as the coefficients gssμ = γssμ are rotation invariant.

18.
M.
Moshinsky
, “
Boundary conditions for the description of nuclear reactions
,”
Phys. Rev.
81
,
347
352
(
1951
).
19.
M.
Moshinsky
, “
Boundary conditions and time-dependent states
,”
Phys. Rev.
84
,
525
532
(
1951
).
20.
M.
Moshinsky
, “
Quantum mechanics in Fock space
,”
Phys. Rev.
84
,
533
(
1951
).
21.
L. E.
Thomas
, “
Multiparticle Schrödinger Hamiltonians with point interactions
,”
Phys. Rev. D
30
,
1233
1237
(
1984
).
22.
D. R.
Yafaev
, “
On a zero-range interaction of a quantum particle with the vacuum
,”
J. Phys. A: Math. Gen.
25
,
963
978
(
1992
).
23.
J.
Lampart
and
J.
Schmidt
, “
On Nelson-type Hamiltonians and abstract boundary conditions
,”
Commun. Math. Phys.
367
,
629
663
(
2019
); arXiv:1803.00872.
24.
J.
Lampart
, “
A nonrelativistic quantum field theory with point interactions in three dimensions
,”
Ann. Henri Poincare
20
,
3509
3541
(
2019
); arXiv:1804.08295.
25.
S.
Keppeler
and
M.
Sieber
, “
Particle creation and annihilation at interior boundaries: One-dimensional models
,”
J. Phys. A: Math. Theor.
49
,
125204
(
2016
); arXiv:1511.03071.
26.
D.
Dürr
,
S.
Goldstein
,
S.
Teufel
,
R.
Tumulka
, and
N.
Zanghì
, “
Bohmian trajectories for Hamiltonians with interior-boundary conditions
,”
J. Stat. Phys.
180
,
34
73
(
2020
); arXiv:1809.10235.
27.
H.-O.
Georgii
and
R.
Tumulka
, “
Some jump processes in quantum field theory
,” in
Interacting Stochastic Systems
, edited by
J.-D.
Deuschel
and
A.
Greven
(
Springer-Verlag
,
Berlin
,
2004
), pp.
55
73
; arXiv:math.PR/0312326.
28.
B.
Galvan
, “
Quantum field theory without divergence: The method of the interaction operators
,” arXiv:1607.03876 (
2016
).
29.
M.
Moshinsky
and
G. L.
Laurrabaquio
, “
Relativistic interactions by means of boundary conditions: The Breit–Wigner formula
,”
J. Math. Phys.
32
,
3519
3528
(
1991
).
30.
J.
Schmidt
, “
On a direct description of pseudorelativistic Nelson Hamiltonians
,”
J. Math. Phys.
60
,
102303
(
2019
); arXiv:1810.03313.
31.
J.
Schmidt
and
R.
Tumulka
, “
Complex charges, time reversal asymmetry, and interior-boundary conditions in quantum field theory
,”
J. Phys. A: Math. Theor.
52
,
115301
(
2019
); arXiv:1810.02173.
32.
R.
Tumulka
, “
Bohmian mechanics at space-time singularities. I. Timelike singularities
,”
J. Geom. Phys.
145
,
103478
(
2019
); arXiv:0708.0070.
33.
R.
Tumulka
, “
Interior-boundary conditions for Schrödinger operators on codimension-1 boundaries
,”
J. Phys. A: Math. Theor.
53
,
155201
(
2020
); arXiv:1808.06262.
34.
G.
Dell’Antonio
,
R.
Figari
, and
A.
Teta
, “
A brief review on point interactions
,” in
Inverse Problems and Imaging
, Lecture Notes in Mathematics Vol. 1943 (
Springer
,
Berlin
,
2008
), pp.
171
189
.
35.
R.
Tumulka
, “
Detection time distribution for the Dirac equation
,” arXiv:1601.04571 (
2016
).
36.
M.
Esteban
,
M.
Lewin
, and
É.
Séré
, “
Domains for Dirac–Coulomb min-max levels
,”
Rev. Mat. Iberoam.
35
(
3
),
877
924
(
2019
).
37.
J.
Weidmann
,
Spectral Theory of Ordinary Differential Operators
(
Springer
,
Heidelberg
,
1987
).
38.
M.
Gallone
, “
Self-adjoint extensions of Dirac operator with Coulomb potential
,” in
Advances in Quantum Mechanics
, edited by
G.
Dell’Antonio
and
A.
Michelangeli
(
Springer
,
2017
), pp.
169
185
; arXiv:1710.02069.
39.
B.
Cassano
and
F.
Pizzichillo
, “
Self-adjoint extensions for the Dirac operator with Coulomb-type spherically symmetric potentials
,”
Lett. Math. Phys.
108
,
2635
2667
(
2018
); arXiv:1710.08200.
40.
J.
Henheik
and
R.
Tumulka
, “
Creation rate of Dirac particles at a point source
,” arXiv:2211.16606 (
2022
).
41.

Although (H°, D°) is rotationally symmetric, since the domain D° is not dense in H, the self-adjoint extensions of (H°, D°) with particle creation are not obtained by simply taking the closure. Hence, as we show, rotational invariance (might) get lost.

42.
H.
Hogreve
, “
The overcritical Dirac–Coulomb operator
,”
J. Phys. A: Math. Theor.
46
,
025301
(
2012
).
43.
M.
Gallone
and
A.
Michelangeli
, “
Self-adjoint realisations of the Dirac–Coulomb Hamiltonian for heavy nuclei
,”
Anal. Math. Phys.
9
,
585
616
(
2019
); arXiv:1706.00700.
44.

Note that the meaning of symbols such as ψ(1) differs by a factor of r [coming from (58)] depending on whether we consider the left- or the right-hand side of the isomorphism (65). This is why the exponents are ±B in (70) but −1 ± B in (35).

45.
M.
Reed
and
B.
Simon
,
Methods of Modern Mathematical Physics II: Fourier Analysis, Self-Adjointness
(
Academic Press
,
1975
).
46.
T.
Binz
and
J.
Lampart
, “
An abstract framework for interior-boundary conditions
,” arXiv:2103.17124 (
2021
).