This paper was converted on www.awesomepapers.org from LaTeX by an anonymous user.
Want to know more? Visit the Converter page.

Quantum measurement process with an ideal detector array

Michael Zirpel, mz@mzirpel.de
Abstract

Any observable with finite eigenvalue spectrum can be measured using a multiport apparatus realizing an appropriate unitary transformation and an array of detector instruments, where each detector operates as an indicator of one possible value of the observable. The study of this setup in the frame of von Neumann’s quantum mechanical measurement process has a remarkable result: already after the interaction of the measured system with the detector array without collapse, exactly one detector is indicating a detection. Each single detector indicates either 0 or 1 detection, and no superposition can be attributed to it.
Keywords:  Measurement process \cdot Detector array \cdot Definite outcome \cdot Preferred base \cdot Wave function collapse

1 Introduction

Reck et al. (1994) demonstrated that any finite dimensional unitary transformation can be realized in the laboratory using a lossless multiport apparatus constructed with beam splitters, mirrors, and phase shifters. Furthermore, any finite dimensional observable can be measured realizing an appropriate unitary transformation, attaching a particle source to the input port and particle detectors to the output ports. In this case, each detector operates as an indicator of one possible value of the observable.

Measurement instruments of this type are used in many modern versions of quantum mechanical experiments. The authors list Einstein-Podolsky-Rosen (Einstein et al., 1935; Bohm and Aharonov, 1957) and Bell test (Bell, 1964; Clauser et al., 1969) experiments, as well as variants of Stern-Gerlach experiments (Gerlach and Stern, 1922), and quantum cryptography devices (Ekert, 1991). A famous example is Aspect’s Bell test experiment (Aspect et al., 1982) with photons, which was repeated in many variants, even with neutrons (Hasegawa et al., 2003). Several other experiments with neutrons (in line with their photon counterparts) are described in (Sponar et al., 2021), with many additional references.

In the following, such measurement instruments are considered theoretically, with a focus on the detectors, which are treated quantum mechanically in the frame of von Neumann’s measurement process (v. Neumann, 1932; Wheeler and Zurek, 1983). Any observable with finite eigenvalue spectrum can be measured with an array of detector instruments using an appropriate measurement interaction, described by a unitary transformation, which fulfills the conditions for a repeatable ideal measurement. It is well-known, that this interaction of the measured system and the measurement instrument ends with an entangled state, which exhibits the “measurement problem”: how to get a definite outcome. Schrödinger’s cat, waiting in a killing machine triggered by the measurement outcome, is iconic for this situation: is it dead, alive, or in a superposition of these states? For the Copenhagen interpretation (Heisenberg, 1958), the measurement must be completed by some final reduction or collapse (Busch et al., 1991) to get a definite outcome. However, with the detector array the entangled state already has some special properties, which are usually assumed to arise with the collapse: exactly one detector is indicating a detection; each single detector indicates either 0 or 1 detection, and no superposition can be attributed to it.

2 Von Neumann’s measurement process

Von Neumann (1932) described the measurement process quantum mechanically as a short time interaction between the measured system SS and a measurement instrument MM, using Hilbert space S\mathcal{H}_{S} and M\mathcal{H}_{M}, respectively, as state space, and the tensor product SM\mathcal{H}_{S}\otimes\mathcal{H}_{M} for the composite system SMSM. The instrument has to display the measurement outcome by entering one of the pairwise orthogonal pointer states φ1,φ2,M\varphi_{1},\varphi_{2},...\in\mathcal{H}_{M}, which correspond to the eigenvalues a1,a2,a_{1},a_{2},...\in\mathbb{R} of the measured observable A(S)A\in\mathcal{B}(\mathcal{H}_{S}) (assumed to be bounded and to have a pure, non-degenerate eigenvalue spectrum). The initial state of the composite system SMSM before the interaction and its final state after that are connected by a unitary transformation USM(SM)U_{SM}\in\mathcal{B}(\mathcal{H}_{S}\otimes\mathcal{H}_{M}), which must fulfill two conditions to constitute a repeatable ideal measurement of the observable AA: any eigenstate αk\alpha_{k} of AA stays unchanged, and the instrument must enter the corresponding pointer state φk\varphi_{k}, indicating the measurement of the eigenvalue aka_{k}, i.e.

USM(αkφ0)=αkφk,U_{SM}\left(\alpha_{k}\otimes\varphi_{0}\right)=\alpha_{k}\otimes\varphi_{k}\,, (1)

with φ0\varphi_{0} as the initial state of the instrument. As a consequence, the interaction of the measurement instrument in initial state φ0\varphi_{0} with a system in state

ψ=kckαk,\psi=\sum_{k}c_{k}\alpha_{k}\>,

with ckc_{k}\in\mathbb{C}, transforms the composite system into the entangled state

USM(ψφ0)=kckαkφk,U_{SM}\left(\psi\otimes\varphi_{0}\right)=\sum_{k}c_{k}\alpha_{k}\otimes\varphi_{k}\,,

with probability pk=|ck|2p_{k}=\bigl{|}c_{k}\bigr{|}^{2} to observe the pointer φk\varphi_{k}, according to Born’s rule. This superposition of products of eigenstates αk\alpha_{k} and corresponding pointer states φk\varphi_{k} exhibits the “measurement problem”, how to get a definite outcome, and is a reason for different interpretations of quantum mechanics (Mittelstaedt, 1998). According to the Copenhagen interpretation, however, this state terminates only the “premeasurement” (Busch et al., 1991). The measurement must be completed by some additional reduction or collapse, giving a definite outcome in a random, irreversible transition

kckαkφkpdαdφd\sum_{k}c_{k}\alpha_{k}\otimes\varphi_{k}\;\overset{p_{d}}{\curvearrowright}\;\alpha_{d}\otimes\varphi_{d}

where φd\varphi_{d} is the pointer state indicating the definite outcome ada_{d}, αd\alpha_{d} the corresponding eigenstate of the measured system and pd=|cd|2p_{d}=\bigl{|}c_{d}\bigr{|}^{2} the transition probability. Different authors associate this transition with different circumstances, e.g. the recognition of the measurement outcome by a conscious observer (Wigner, 1961) or its registration by a classical device (Bohr, 1958). Some non-Copenhagen interpretations (e.g. the pilot wave interpretation (Bohm, 1952a, b), the many-worlds interpretation (Everett, 1957)) deny such an additional transition at all and assume that the final premeasurement state already constitutes the end of the measurement.

The state αd\alpha_{d} of the system after reduction is an eigenstate of AA, so a repetition of the measurement will give the same outcome ada_{d}. However, as Pauli (1933) already noticed, there are ideal measurements for which an immediate repetition gives not the same result. The measurement scheme of the multiport apparatus with detectors (Reck et al., 1994) is of this type. Von Neumann’s conditions (1) must be relaxed to describe such non-repeatable ideal measurements (Busch et al., 1991): an eigenstate αk\alpha{}_{k} will be transformed into another state αkS\alpha^{\prime}_{k}\in\mathcal{H}_{S}, while the instrument enters the corresponding pointer state

USM(αkφ0)=αkφk.U_{SM}\left(\alpha{}_{k}\otimes\varphi_{0}\right)=\alpha^{\prime}_{k}\otimes\varphi_{k}\,. (2)

For the multiport apparatus we can write αk=Uαk\alpha^{\prime}_{k}=U\alpha{}_{k}, with the unitary transformation U(S)U\in\mathcal{B}(\mathcal{H}_{S}) determined by the apparatus. Here, these αk\alpha^{\prime}_{k} are not eigenstates of AA but of the observable

A=UAU1A^{\prime}=UAU^{-1} (3)

measured repeatably by the detectors. The collapse of the final premeasurement state gives then

kckαkφkpdαdφd\sum_{k}c_{k}\alpha^{\prime}_{k}\otimes\varphi_{k}\;\overset{p_{d}}{\curvearrowright}\;\alpha^{\prime}_{d}\otimes\varphi_{d} (4)

indicating the measurement outcome ada_{d}, with the system in the state αd\alpha^{\prime}_{d}.

3 Premeasurement with an ideal detector

In the following, an ideal detector is considered as an instrument for measuring an indicator observable (binary observable) which is represented by a projection operator E(S)E\in\mathcal{B}(\mathcal{H}_{S}) with E=E=E2E=E^{\dagger}=E^{2}, eigenvalues σ(A)={0,1}\sigma(A)=\{0,1\}, and the complement E¯=1E\overline{E}=1-E. A suitable detector instrument DD has at least 22 orthogonal pointer states φ0,φ1D\varphi_{0},\varphi_{1}\in\mathcal{H}_{D} to display the measurement result. With Hilbert space D=2\mathcal{H}_{D}=\mathbb{C}^{2}, the detector instrument can be a spin-1/2-system or a qubit.

The mathematical expression of von Neumann’s conditions (1) for the interaction USD(SD)U_{SD}\in\mathcal{B}(\mathcal{H}_{S}\otimes\mathcal{H}_{D}) must be generalized to include repeatable ideal measurements of degenerate observables (Lüders, 1951). For an arbitrary vector ψS\psi\in\mathcal{H}_{S}, EψE\psi and E¯ψ\overline{E}\psi are (unnormalized) eigenvectors of EE (or the zero vector) with eigenvalues 11 or 0,0, respectively. So, the conditions are

USD(Eψφ0)=Eψφ1U_{SD}\left(E\psi\otimes\varphi_{0}\right)=E\psi\otimes\varphi_{1} (5)
USD(E¯ψφ0)=E¯ψφ0U_{SD}\left(\overline{E}\psi\otimes\varphi_{0}\right)=\overline{E}\psi\otimes\varphi_{0}

for all ψS\psi\in\mathcal{H}_{S}.

Example 1

Premeasurement with a detector.

a) A vector of the Hilbert space S2\mathcal{H}_{S}\otimes\mathbb{C}^{2} can be written as 2-component column vector (similar as a spinor). In this notation, the matrix

VSD=(E¯EEE¯)(S2),V_{SD}=\left(\begin{array}[]{cc}\overline{E}&E\\ E&\overline{E}\end{array}\right)\in\mathcal{B}(\mathcal{H}_{S}\otimes\mathbb{C}^{2}),

with VSD=VSD1=VSDV_{SD}=V_{SD}^{-1}=V_{SD}^{\dagger}, fulfills the conditions for a repeatable ideal measurement (5).
b) If the measured system SS is a spinless particle, with S=2(3)\mathcal{H}_{S}=\mathcal{L}^{2}(\mathbb{R}^{3}), an indicator observable for position detection is the indicator function IV:3{0,1}I_{V}:\mathbb{R}^{3}\rightarrow\{0,1\}, with IV(r)=1rVI_{V}(r)=1\Leftrightarrow r\in V, of a Borel set VB(3)V\in B(\mathbb{R}^{3}) representing the detection volume (in position representation), with

VSD=(1IVIVIV1IV)(2(3)2).V_{SD}=\left(\begin{array}[]{cc}1-I_{V}&I_{V}\\ I_{V}&1-I_{V}\end{array}\right)\in\mathcal{B}(\mathcal{L}^{2}(\mathbb{R}^{3})\otimes\mathbb{C}^{2})\,.

c) If the measured system SS is a qubit (Nielsen and Chuang, 2000), with S=2,\mathcal{H}_{S}=\mathbb{C}^{2}, a indicator observable for a measurement in the computational base is

E~=|11|=(0001),E~¯=|00|=(1000),\tilde{E}=\bigl{|}1\bigr{\rangle}\bigl{\langle}1\bigr{|}=\left(\begin{array}[]{cc}0&0\\ 0&1\end{array}\right),\overline{\tilde{E}}=\bigl{|}0\bigr{\rangle}\bigl{\langle}0\bigr{|}=\left(\begin{array}[]{cc}1&0\\ 0&0\end{array}\right),

and with its matrix representation one gets a 4×44\times 4 matrix for the measurement transformation

V~SD=(1000000100100100),\tilde{V}_{SD}=\left(\begin{array}[]{cc|cc}1&0&0&0\\ 0&\text{0}&0&1\\ \hline\cr 0&0&1&0\\ 0&1&0&0\end{array}\right),

which is a representation of a CNOT circuit with 2 qubits (fig. 1).

Refer to caption
Figure 1: Premeasurement with CNOT circuit

4 Premeasurement with an ideal detector array

Any observable A(S)A\in\mathcal{B}(\mathcal{H}_{S}) with a finite eigenvalue spectrum σ(A)={a1,,an}\sigma(A)=\{a_{1},...,a_{n}\}\subset\mathbb{R} can be written as a linear combination of projection operators Ek(S)E_{k}\in\mathcal{B}(\mathcal{H}_{S}) onto the eigenspaces

A=k=1nakEk,A=\sum_{k=1}^{n}a_{k}E_{k}\,,

with

EjEk=EkEj=δj,k1E_{j}E_{k}=E_{k}E_{j}=\delta_{j,k}1 (6)

for all j,kIn={1,,n}j,k\in I_{n}=\{1,...,n\}, and

k=1nEk=1.\sum_{k=1}^{n}E_{k}=1\,. (7)

The projection operators E1,,EnE_{1},...,E_{n} are indicator observables for the corresponding eigenvalues and can be measured using ideal detector instruments D1,,DnD_{1},...,D_{n}, assuming a bijection

EkDk.E_{k}\leftrightarrows D_{k}\,. (8)

An array of these nn detectors can be used as a measurement instrument for the observable AA. With pointer states φ0(k)\varphi_{0}^{(k)}, φ1(k)\varphi_{1}^{(k)} of detector DkD_{k}, the pointer states of this instrument MM are

φ~0=φ0(1)φ0(n),\mathrm{\tilde{\varphi}}_{0}=\mathrm{\varphi_{0}^{(1)}}\otimes...\otimes\varphi_{0}^{(n)}\,,
φ~k=φ0(1)φ1(k)φ0(k+1)φ0(n)\mathrm{\tilde{\varphi}}_{k}=\mathrm{\varphi_{0}^{(1)}}\otimes...\otimes\varphi_{1}^{(k)}\otimes\varphi_{0}^{(k+1)}\otimes...\otimes\varphi_{0}^{(n)}

for kInk\in I_{n}, in Hilbert space M=D(1)D(n)=Dn\mathcal{H}_{M}=\mathcal{H}_{D}^{(1)}\otimes...\otimes\mathcal{H}_{D}^{(n)}=\mathcal{H}_{D}^{\otimes n}. The measurement result is the index kk of the detector with the pointer state φ1(k)\varphi_{1}^{(k)}, equivalent to the pointer state φ~k\mathrm{\tilde{\varphi}}_{k} of the whole array, indicating the value aka_{k} of the observable AA. Here, von Neumann’s conditions (1) for an repeatable ideal measurement of observable AA are

USM(Ekψφ~0)=Eψkφ~kU_{SM}(E_{k}\psi\otimes\tilde{\varphi}_{0})=E{}_{k}\psi\otimes\tilde{\varphi}_{k} (9)

for all ψS\psi\in\mathcal{H}_{S} and kInk\in I_{n}.

In the rest of this paragraph, a construction for this unitary transformation USM(SM)U_{SM}\in\mathcal{B}(\mathcal{H}_{S}\otimes\mathcal{H}_{M}) is given, using the detector transformation USD(k)(SD(k))U_{SD}^{(k)}\in\mathcal{B}(\mathcal{H}_{S}\otimes\mathcal{H}_{D}^{(k)}) from (5) for each single detector DkD_{k}. With (6) we can write (5)

USD(k)(Ekψφ0(j))=Ekψφδj,k(j)U_{SD}^{(k)}\left(E_{k}\psi\otimes\varphi_{0}^{(j)}\right)=E_{k}\psi\otimes\varphi_{\delta_{j,k}}^{(j)} (10)

for all j,kInj,k\in I_{n}, and USMU_{SM} can be constructed as product

USM=UM(1)UM(n)U_{SM}=U_{M}^{(1)}\cdot...\cdot U_{M}^{(n)}

where

UM(k)=1(1)USD(k)1(k+1)1(n)U_{M}^{(k)}=1^{(1)}\otimes...\otimes U_{SD}^{(k)}\otimes 1^{(k+1)}...\otimes 1^{(n)}

is the representation of USD(k)U_{SD}^{(k)} in the product space SM.\mathcal{H}_{S}\otimes\mathcal{H}_{M}. It is obvious, that these operators commute

UM(j)UM(k)=1(1)USD(j)USD(k)1(n)=UM(k)UM(j)U_{M}^{(j)}U_{M}^{(k)}=1^{(1)}\otimes...U_{SD}^{(j)}\otimes...U_{SD}^{(k)}\otimes...\otimes 1^{(n)}=U_{M}^{(k)}U_{M}^{(j)}

for all j,kInj,k\in I_{n}. To demonstrate that USMU_{SM} fulfills von Neumann’s conditions for the eigenstates EkψE_{k}\psi of AA, we multiply the initial state of the composite system

Φ0=Ekψφ0(1)φ0(n)=Ekψφ~0\varPhi_{0}=E_{k}\psi\otimes\mathrm{\varphi_{0}^{(1)}}\otimes...\otimes\varphi_{0}^{(n)}=E_{k}\psi\otimes\tilde{\varphi}_{0}

for each kInk\in I_{n} successively with UM(1),,UM(n)U_{M}^{(1)},...,U_{M}^{(n)} using (10)

Φ1=UM(1)Φ0=Ekψφ0(1)φ0(k)φ0(n)=Ekψφ~0\varPhi_{1}=U_{M}^{(1)}\varPhi_{0}=E_{k}\psi\otimes\mathrm{\varphi_{0}^{(1)}}\otimes...\otimes\varphi_{0}^{(k)}\otimes...\otimes\varphi_{0}^{(n)}=E_{k}\psi\otimes\tilde{\varphi}_{0}
...
Φk=UM(k)Φk1=Ekψφ0(1)φ1(k)φ0(k+1)φ0(n)=Ekψφ~k\varPhi_{k}=U_{M}^{(k)}\varPhi_{k-1}=E_{k}\psi\otimes\mathrm{\varphi_{0}^{(1)}}\otimes...\otimes\varphi_{1}^{(k)}\otimes\varphi_{0}^{(k+1)}\otimes...\otimes\varphi_{0}^{(n)}=E_{k}\psi\otimes\tilde{\varphi}_{k}
...
USMΦ0=Φn=UM(n)Φn1=Ekψφ0(1)φ1(k)φ0(k+1)φ0(n)=Ekψφ~k,U_{SM}\varPhi_{0}=\varPhi_{n}=U_{M}^{(n)}\varPhi_{n-1}=E_{k}\psi\otimes\mathrm{\varphi_{0}^{(1)}}\otimes...\otimes\varphi_{1}^{(k)}\otimes\varphi_{0}^{(k+1)}\otimes...\otimes\varphi_{0}^{(n)}=E_{k}\psi\otimes\tilde{\varphi}_{k}\,,

which gives (9).

Example 2

Premeasurement with a 2-detector array.

a) With the indicator observables E1,E2(S)E_{1},E_{2}\in\mathcal{B}(\mathcal{H}_{S}) the matrix representation of VSDV_{SD} from Example 1 gives

VSM=(VSD(1)1(2))(1(1)VSD(2))=(0E1E20E100E2E200E10E2E10),V_{SM}=(V_{SD}^{(1)}\otimes 1^{(2)})(1^{(1)}\otimes V_{SD}^{(2)})=\left(\begin{array}[]{cccc}0&E_{1}&E_{2}&0\\ E_{1}&0&0&E_{2}\\ E_{2}&0&0&E_{1}\\ 0&E_{2}&E_{1}&0\end{array}\right),
VSM(ψ000)=(0E1ψE2ψ0),V_{SM}\left(\begin{array}[]{c}\psi\\ 0\\ 0\\ 0\end{array}\right)=\left(\begin{array}[]{c}0\\ E_{1}\psi\\ E_{2}\psi\\ 0\end{array}\right),

which fulfills (9).
b) If the measured system SS is a particle, the Hilbert space can be restricted to a finite box B(3)B\in\mathcal{B}(\mathbb{R}^{3}), with S=2(B)\mathcal{H}_{S}=\mathcal{L}^{2}(B). Then, the completeness condition (7) for the partition of the position space V1V2=BV_{1}\cup V_{2}=B can be fulfilled with two finite detection volumes and the corresponding indicator functions E1=IV1E_{1}=I_{V_{1}}, E2=IV2E_{2}=I_{V_{2}} in position representation. Such partitions give a simple model of a tracking chamber.
c) If the measured system is a qubit, the measurement interaction for a measurement in the computational base with E~1=|11|\tilde{E}_{1}=\bigl{|}1\bigr{\rangle}\bigl{\langle}1\bigr{|} and E~2=|00|\tilde{E}_{2}=\bigl{|}0\bigr{\rangle}\bigl{\langle}0\bigr{|} can be described with the resulting 8×88\times 8 matrix for V~SM\tilde{V}_{SM} or, equivalently, with an 11-to-22 decoder circuit (fig. 2).

Refer to caption
Figure 2: Premeasurement with 1-to-2 decoder circuit

5 Properties of the final premeasurement state

It is a consequence of (7) and the conditions (9), that for an arbitrary initial state ψS\psi\in\mathcal{H}_{S} of the measured system SS the final premeasurement state of the composite system SMSM is

USM(ψφ~0)=U((k=1nEk)ψφ~0)SM=k=1nEψkφ~k.U_{SM}\left(\psi\otimes\tilde{\varphi}_{0}\right)=U{}_{SM}\bigl{(}\bigl{(}\sum_{k=1}^{n}E_{k}\bigr{)}\psi\otimes\tilde{\varphi}_{0}\bigr{)}=\sum_{k=1}^{n}E{}_{k}\psi\otimes\tilde{\varphi}_{k}\>.

To include non-repeatable ideal measurements with the multiport apparatus (2), we discuss the more general final premeasurement state

Φ=k=1nEψkφ~k\varPhi=\sum_{k=1}^{n}E^{\prime}{}_{k}\psi\otimes\tilde{\varphi}_{k} (11)

where Ek=U1EkUE^{\prime}_{k}=U^{-1}E_{k}U is a projection operator for all kInk\in I_{n}, defined by the inverse transformation of (3).

This entangled state of the system and the detector array has several interesting properties.

5.1 Exactly one detection

An observable of the detector array is the number of detectors indicating a detection N(M)N\in\mathcal{B}(\mathcal{H}_{M}). With the eigenvalue spectrum σ(N)={0,1,,n}\sigma(N)=\{0,1,...,n\} it can be written as

N=k=0nkQk,N=\sum_{k=0}^{n}kQ_{k}\,,

with projection operators Qk(M)Q_{k}\in\mathcal{B}(\mathcal{H}_{M}) onto the closed subspaces spanned by the vectors

Vk={φi1(1)φin(n)Mi1,,in{0,1},i1++in=k}.V_{k}=\left\{\varphi_{i_{1}}^{(1)}\otimes...\otimes\varphi_{i_{n}}^{(n)}\in\mathcal{H}_{M}\mid i_{1},...,i_{n}\in\{0,1\},i_{1}+...+i_{n}=k\right\}\,.

It is obvious that φ~0V0\mathrm{\tilde{\varphi}}_{0}\in V_{0} is an eigenvector of NN with eigenvalue 0 and φ~1,,φ~nV1\mathrm{\tilde{\varphi}}_{1},...,\mathrm{\tilde{\varphi}}_{n}\in V_{1} are eigenvectors with eigenvalue 11. Consequently, the final premeasurement state Φ\mathrm{\varPhi} (11) is an eigenstate of 1SN1_{S}\otimes N with eigenvalue 1

(1SN)Φ=k=1nEkψNφ~k=k=1nEkψφ~k=Φ.\left(1_{S}\otimes N\right)\mathrm{\varPhi}=\sum_{k=1}^{n}E^{\prime}_{k}\psi\otimes N\tilde{\varphi}_{k}=\sum_{k=1}^{n}E^{\prime}_{k}\psi\otimes\tilde{\varphi}_{k}=\mathrm{\varPhi\,.}

So, the observable 1SN1_{S}\otimes N has in state Φ\mathrm{\varPhi} the dispersion free expectation value 11. This means, exactly one detector is indicating a detection.

5.2 Preferred pointer base states

The biorthogonal Schmidt decomposition of an entangled state is in general not unique (Schlosshauer, 2004; Ekert and Knight, 1995). Consequently, the same entangled state can arise as a final premeasurement state measuring incompatible observables. However, with the final premeasurement state Φ\mathrm{\varPhi} (11), only eigenstates of the number operator with eigenvalue 11 can be used as factors for the detector array. This can be shown considering the alternative Schmidt decomposition

Φ=jdjβjχj,\varPhi=\sum_{j}d_{j}\beta_{j}\otimes\chi_{j}\,,
(1SN)Φ=Φ(1_{S}\otimes N)\varPhi=\varPhi\;\Rightarrow
jdjβjNχj=djβjχj\sum_{j}d_{j}\beta_{j}\otimes N\chi_{j}=\sum d_{j}\beta_{j}\otimes\chi_{j}\;\Rightarrow
jdjβj(Nχjχj)=0.\sum_{j}d_{j}\beta_{j}\otimes(N\chi_{j}-\chi_{j})=0\,.

Because of the pairwise orthogonality, each summand must be zero. Therefore, each χj\chi_{j} has to be an eigenstate of NN with eigenvalue 11.

There is no factorization of such an eigenstate with other single detector states than the pointer states φ0\varphi_{0},φ1\varphi_{1}. To see this, let φ~Dn\tilde{\varphi}\in\mathcal{H}_{D}^{\otimes n} be an eigenvector of the number operator with eigenvalue 11 and

φ~=χψ,\tilde{\varphi}=\chi\otimes\psi\,,

a factorization with ψ=aφ0+bφ1D\psi=a\varphi_{0}+b\varphi_{1}\in\mathcal{H}_{D}. Then, we can write

χ=cχ0+dχ1,\chi=c\chi_{0}+d\chi_{1}\,,

with N(χ0φ0)=0N\left(\chi_{0}\otimes\varphi_{0}\right)=0 and N(χ1φ0)=χ1φ0N\left(\chi_{1}\otimes\varphi_{0}\right)=\chi_{1}\otimes\varphi_{0}, and

χψ=acχ0φ0+adχ1φ0+bcχ0φ1+bdχ1φ1\chi\otimes\psi=ac\chi_{0}\otimes\varphi_{0}+ad\chi_{1}\otimes\varphi_{0}+bc\chi_{0}\otimes\varphi_{1}+bd\chi_{1}\otimes\varphi_{1}
N(χψ)=0+adχ1φ0+bcχ0φ1+2bdχ1φ1.N(\chi\otimes\psi)=0+ad\chi_{1}\otimes\varphi_{0}+bc\chi_{0}\otimes\varphi_{1}+2bd\chi_{1}\otimes\varphi_{1}\,.

Non-trivial equality is only possible with c=0c=0 and b=0b=0 or with d=0d=0 and a=0a=0. Since the factorization of the state

Φ=jcjβjχj=jcjβjχ~jψ\varPhi=\sum_{j}c_{j}\beta_{j}\otimes\chi_{j}=\sum_{j}c_{j}\beta_{j}\otimes\tilde{\chi}_{j}\otimes\psi

is a factorization of all χj\chi_{j}, it is only possible with ψφ0\psi\propto\varphi_{0} or ψφ1\psi\propto\varphi_{1}.

For a single detector DkD_{k} of the array, with kInk\in I_{n}, this has the consequence that no superposition of the pointer states φ0(k)\varphi_{0}^{(k)} and φ1(k)\varphi_{1}^{(k)} can be attributed to the final premeasurement state Φ\mathrm{\varPhi} (11) as a factor of a Schmidt decomposition. The pointer base states φ0(k),φ1(k)\varphi_{0}^{(k)},\varphi_{1}^{(k)} are preferred (Zurek, 2003) and the reduced state of each detector DkD_{k} is a unique mixture of φ0(k)\varphi_{0}^{(k)} and φ1(k)\varphi_{1}^{(k)}.

5.3 Conditional expectations equal expectations after collapse

Other studies (Laura and Vanni, 2008) have shown, that for an ideal repeatable measurement, fulfilling von Neumann’s conditions, the final premeasurement state gives the same conditional probabilities for the measured system as the collapse of the wave function. This is also the case with the collapse of the premeasurement state after a non-repeatable ideal measurement (4).

To see this for state Φ\mathrm{\varPhi} (11) let PkMP_{k}\in\mathcal{H}_{M} be the indicator observable for detection by detector DkD_{k}, i.e. the projection operator onto φ~k\tilde{\varphi}_{k}. It commutes with all system observables. So, a common probability space is available for all these observables and the conditional probability of any event represented by the indicator observable F(S)F\in\mathcal{B}(\mathcal{H}_{S}) of the system, given Pk(M)P_{k}\in\mathcal{B}(\mathcal{H}_{M}), is well defined for all kInk\in I_{n}

p(FPk)=p(FPk)p(Pk)=Φ,FPkΦΦ,1SPkΦ=ψ,EkFEkψψ,Ekψp(F\mid P_{k})=\frac{p(F\wedge P_{k})}{p(P_{k})}=\frac{\bigl{\langle}\varPhi,F\otimes P_{k}\varPhi\bigr{\rangle}}{\bigl{\langle}\varPhi,1_{S}\otimes P_{k}\varPhi\bigr{\rangle}}=\frac{\bigl{\langle}\psi,E^{\prime}_{k}FE^{\prime}_{k}\psi\bigr{\rangle}}{\bigl{\langle}\psi,E^{\prime}_{k}\psi\bigr{\rangle}}

if ψ,Ekψ>0\bigl{\langle}\psi,E^{\prime}_{k}\psi\bigr{\rangle}>0. This is a consequence of Born’s rule. The state of the system after the complete measurement with collapse, given the outcome aka_{k}, is according to (4)

ψak=1ψ,EkψEkψ.\psi_{a_{k}}=\frac{1}{\sqrt{\bigl{\langle}\psi,E^{\prime}_{k}\psi\bigr{\rangle}}}E^{\prime}_{k}\psi\,.

For this state the probability of FF is

ψak,Fψak=ψ,EkFEkψψ,Ekψ=p(FPk)\bigl{\langle}\psi_{a_{k}},F\psi_{a_{k}}\bigr{\rangle}=\frac{\bigl{\langle}\psi,E^{\prime}_{k}FE^{\prime}_{k}\psi\bigr{\rangle}}{\bigl{\langle}\psi,E^{\prime}_{k}\psi\bigr{\rangle}}=p(F\mid P_{k})

and has the same value as the conditional probability of FF, given PkP_{k}, for the state Φ\varPhi. The same is valid for the conditional expectation value of any observable of the system.

Consequently, considering the Heisenberg picture, the isolated evolution of the system SS after the measurement interaction, given PkP_{k}, is the same as for the collapsed wave function after the measurement, given the outcome aka_{k}.

6 Conclusion

The entangled state of the system and the detector array after the ideal premeasurement of an observable has the remarkable property that the number of detectors indicating a detection is exactly one. This is a hint at a definite outcome of the measurement and resembles the collapse, since the conditional expectations, given a detection, are the same as for the collapsed state, given the corresponding outcome. Furthermore, no superposition state can be attributed to a single detector: the pointer base states and their mixtures are preferred.

Using Schrödinger’s cat illustration, one can say: if in a lossless multiport experiment a particle has different exits for all possible measurement outcomes, each one with an ideal detector connected to a killing machine with its own cat, then exactly one cat will be dead, and all others will be alive after a trial; there is no superposition of dead and alive – even after a pure unitary evolution without collapse.

These properties depend on an ideal measurement process where the detector array fulfills the completeness and orthogonality conditions (7), (6) with the bijection (8). Quantum computing circuits (example 2c) demonstrate the possibility of such a setting. In case of position detectors (example 2b), exactly positioned detection volumes without gap and overlap, filling the available space, are required, which seams only possible in a macroscopic quasi-classical limit, in line with Bohr’s Copenhagen (Bohr, 1958) interpretation.

Solely, these properties cannot constitute irreversible facts. That, however, could be accomplished through the decoherence of the final premeasurement state (Schlosshauer, 2004).

The results of our study may support interpretations of quantum mechanics without collapse. However, if some of the latter’s observable consequences are deducible from other rules, this will not mean a contradiction to the Copenhagen interpretation. Even, if the collapse postulate was rendered unnecessary, it could persist as useful rule of thumb.

References

  • Aspect et al. [1982] A. Aspect, J. Dalibard, and G. Roger. Experimental test of Bell’s inequalities using time-varying analyzers. Phys. Rev. Lett., 49, 1982. doi: 10.1103/PhysRevLett.49.91.
  • Bell [1964] J. S. Bell. On the Einstein-Podolsky-Rosen paradox. Physics, 1, 1964. doi: 10.1017/CBO9780511815676.004.
  • Bohm [1952a] D. Bohm. A suggested interpretation of the quantum theory in terms of ’hidden’ variables. I. Phys. Rev., 85, 1952a. doi: 10.1103/PhysRev.85.166.
  • Bohm [1952b] D. Bohm. A suggested interpretation of the quantum theory in terms of ’hidden’ variables. II. Phys. Rev., 85, 1952b. doi: 10.1103/PhysRev.85.180.
  • Bohm and Aharonov [1957] D. Bohm and Y. Aharonov. Discussion of Experimental Proof for the Paradox of Einstein, Rosen, and Podolsky. Phys. Rev., 108, 1957. doi: 10.1103/PhysRev.108.1070.
  • Bohr [1958] N. Bohr. Quantum Physics and Philosophy - Causality and Complementarity. In R. Klibansky, editor, Philosophy in the Mid-Century. La nuova Italia Editrice, Florence, 1958.
  • Busch et al. [1991] P. Busch, P. J. Lahti, and P. Mittelstaedt. The Quantum Theory of Measurement. Springer-Verlag, Berlin, 1991.
  • Clauser et al. [1969] J. F. Clauser, M. A. Horne, A. Shimony, and R. A. Holt. Proposed experiment to test local hidden-variable theories. Phys. Rev. Lett., 23, 1969. doi: 10.1103/PhysRevLett.23.880.
  • Einstein et al. [1935] A. Einstein, B. Podolsky, and N. Rosen. Can quantum-mechanical description of physical reality be considered complete? Phys. Rev., 47, 1935. doi: 10.1103/PhysRev.47.777.
  • Ekert [1991] A. K. Ekert. Quantum cryptography based on Bell’s theorem. Phys. Rev. Lett., 67, 1991. doi: 10.1103/PhysRevLett.67.661.
  • Ekert and Knight [1995] A. K. Ekert and P. L. Knight. Entangled quantum systems and the Schmidt decomposition. Am. J. Phys., 63, 1995. doi: 10.1119/1.17904.
  • Everett [1957] H. Everett. "relative state" formulation of quantum mechanics. Rev. Mod. Phys., 29, 1957. doi: 10.1103/RevModPhys.29.454.
  • Gerlach and Stern [1922] W. Gerlach and O. Stern. Der experimentelle Nachweis der Richtungsquantelung im Magnetfeld. Z. Physik, 9, 1922. doi: 10.1007/BF01326983.
  • Hasegawa et al. [2003] Y. Hasegawa, R. Loidl, G. Badurek, M. Baron, and H. Rauch. Violation of a bell-like inequality in single-neutron interferometry. Nature, 425, 2003. doi: 10.1038/nature01881.
  • Heisenberg [1958] W. Heisenberg. Physics and Philosophy. Penguin Classics, Reprint, 1958.
  • Laura and Vanni [2008] R. Laura and L. Vanni. Conditional probabilities and collapse in quantum measurements. Int. Journal of Theor. Phys., 47, 2008. doi: 10.1007/s10773-008-9672-7.
  • Lüders [1951] G. Lüders. Über die Zustandsänderung durch den Messprozess. Ann. d. Physik, 1951. doi: 10.1002/andp.200610207.
  • Mittelstaedt [1998] P. Mittelstaedt. The Interpretation of Quantum Mechanics and the Measurement Process. Cambridge University Press, 1998.
  • Nielsen and Chuang [2000] M. Nielsen and I. L. Chuang. Quantum Computation and Quantum Information. Cambridge University Press, 2000.
  • Pauli [1933] W. Pauli. Die allgemeinen Prinzipien der Wellenmechanik. Springer-Verlag, Berlin, 1933.
  • Reck et al. [1994] M. Reck, A. Zeilinger, H. J. Bernstein, and P. Bertani. Experimental realization of any discrete unitary operator. Phys. Rev. Lett., 73, 1994. doi: 10.1103/PhysRevLett.73.58.
  • Schlosshauer [2004] M. Schlosshauer. Decoherence, the measurement problem, and interpretation of quantum mechanics. Rev. Mod. Phys., 76, 2004. doi: 10.1103/RevModPhys.76.1267.
  • Sponar et al. [2021] S. Sponar, R. I. P. Sedmik, and M. Pitschmann. Tests of fundamental quantum mechanics and dark interactions with low-energy neutrons. Nat. Rev. Phys., 3, 2021. doi: 10.1038/s42254-021-00298-2.
  • v. Neumann [1932] J. v. Neumann. Mathematische Grundlagen der Quantenmechanik. Springer-Verlag, Berlin, 1932.
  • Wheeler and Zurek [1983] J. A. Wheeler and W. H. Zurek, editors. Quantum Theory and Measurement. Princeton University Press, Princeton, 1983.
  • Wigner [1961] E. P. Wigner. Remarks on the mind body question. In I. J. Good, editor, The Scientist Speculates. William Heinemann, Ltd., London, 1961.
  • Zurek [2003] W. H. Zurek. Decoherence, einselection, and the quantum origins of the classical. Rev. Mod. Phys., 75, 2003. doi: 10.1103/RevModPhys.75.715.