Expand this Topic clickable element to expand a topic
Skip to content
Optica Publishing Group

Using state tomography for characterizing input principal modes in optically scattering medium

Open Access Open Access

Abstract

We propose two new methods to measure principal modes, or Eisenbud-Wigner-Smith eigenstates in optically scattering medium. Both methods use similar techniques as in quantum state tomography, and are based on direct measurement of temporal delays. The first method requires N2 different input launching conditions, and only the mean signal delay of these input states are needed to obtain full information of the principal modes. When the mode delay differences are large and all modes are non-degenerate, a second method can be used, which only requires 3N − 2 input launching conditions.

© 2016 Optical Society of America

1. Introduction

Coherent light scattering in space and time through disordered medium is a very interesting topic. Due to the inherent inhomogeneity within the amorphous medium through which light travels, the coherent scattering causes spatial and temporal speckle patterns. These speckles limit the resolution in various optical systems, such as the spatial resolution in optical imaging or data transmission rate in optical communication.

In terms of temporal scattering, there exists a set of modes called principal modes, or Eisenbud-Wigner-Smith eigenstates [1–3], that despite spatiotemporal scattering, always arrive at the output temporally unscattered. The existence of these principal modes in multi-mode fiber (MMF) was first proposed in 2005 [4], in analogy to the principal states of polarization (PSPs) in single-mode fiber [5, 6]. The principal modes represent the paths through the medium with the maximum possible spatial coherence [7]. A large amount of work ensued [8–11], but primarily focused on the theoretical investigation of the principal modes, in terms of modeling, numerical simulation and statistics on the mode delay distribution.

Experimentally creating and characterizing these special modes will have a major impact on understanding the coherent light scattering phenomenon, as well as higher order dispersion and nonlinearity of these principal modes [12,13]. It will also be the first step for any potential applications using the principal modes [14–17]. The only experimental demonstration of creating and characterizing principal modes in MMFs so far were only made recently [7, 18] due to the high accuracy required of the spatial and spectral information characterizing these highly scattering states. In these paper, the authors measured the full transfer matrix between the input and output of the MMFs, as well as its first order wavelength derivative, and then calculated the principal modes based on this information. On the other hand, since principal modes correspond to the temporal unscattered modes, it is possible to measure principal modes using exclusively temporal measurements. A mode dependent signal delay method was recently proposed [19]. There the authors showed that just by measuring the mean signal time delays for N2 different predetermined inputs, it is possible to completely identify all the principal modes.

In this paper, we utilize the similarity between principal mode characterization and quantum state tomography [20,21], and derive a more straightforward algorithm for measuring principal states using the temporal delay of N2 predetermined input modes. An analytic expression for the full mode delay operator based on these delays is given for an arbitrary N. We also show that if higher order dispersion is negligible, and the output is composed of discrete pulses, one only needs 3N − 2 inputs to fully identify all the principal modes. We will use numerical simulation to demonstrate how the methods work in a 3-mode fiber.

2. Group delay operator

The Maxwell’s equations, in the absence of sources and currents and under the paraxial approximation, are reduced to the Helmholtz equation [22]

T2A(x)i2kzA(x)=0
This equation has mathematically the same form as the Schrödinger equation in quantum mechanics, and thus often referred to as the (2+1 dimensional) linear Schrödinger equation. As a result, many concepts in quantum mechanics, such as the Hilbert space and the Dirac formalism can be directly applied to study the paraxial electromagnetic wave propagation.

For a narrow band optical signal centered at frequency ω, propagating in a multi-mode fiber (MMF) that supports N propagating modes, the light propagation due to mode coupling within the fiber can be described as [4]

|ψout(ω)=U^(ω)|ψin
where |ψin〉 and |ψout〉 are N dimensional vectors representing the normalized electric field pattern of the input and output field; Û(ω) is an N dimensional unitary transformation representing mode propagation and mode coupling.

In analogy to the principal states of polarization (PSPs) in single mode fibers (SMFs) [5], the principal modes in MMFs are defined such that given an input principal mode |pin〉, the corresponding output principal mode |pout〉 is constant with respect to frequency ω to the first order, up to a certain phase, i.e.,

ω(eiϕ(ω)|pout(ω))=0
Plugging in Eq. (2), one immediately obtains
F^(ω)|pin=τ|pin
where (ω) = −(ω)ωÛ(ω) is the mode delay operator, and τ = ωϕ(ω). Since Û(ω) is unitary, it is straightfoward to show that (ω) is a Hermitian operator [4]. Moreover, just as in the case of SMFs [6], the mean signal delay of an arbitrary input signal is the expected value of (ω) on that state [19].
τgdttEout(t)Eout(t)dtEout(t)Eout(t)=idωEout(ω)ωEout(ω)dωEout(ω)Eout(ω)=iψout(ω)|ω|ψout(ω)ψout(ω)|ψout(ω)=iψin|U^(ω)ωU^(ω)|ψin1=ψin|F^(ω)|ψin
This means that the Hermitian operator (ω) is indeed the operator that corresponds to the physical quantity of the group delay, and the input principal modes are non other than its eigenmodes {|piin}. In this case, the expected value for a quantum operator corresponds to the mean value of a classical operator. For non-degenerate , the eigenmodes are always orthonormal; for degenerate , one can still find an orthonormal principal mode sets. In addition, their corresponding output states |piout=U^|piin also forms an orthonormal basis at the output end of the fiber [4], since
pjout|piout=pjinU^U^|piin=pjin|piin=δij
If an input principal mode is launched into the fiber, it comes out at the output end of the fiber unscattered, even though it could still experience temporal broadening or splitting during the propagation. On the other hand, this also means that measuring all the principal modes and their respective delays equates to a full tomography on the operator (ω).

3. Tomography using mean delays

The first method we propose is based on the similarity between quantum state tomography using projection measurement [20,21] and operator tomography using mean value measurement. In quantum state tomography, the density matrix (a Hermitian matrix) is measured by projecting it onto a fixed set of selected input states. Here in the mean delay characterization, we measure the mode delay operator (a Hermitian matrix) by using the mean signal delays (equivalent to projection) of a set of input modes.

Below we detail the mathematics of the operator tomography method. An arbitrary Hermitian matrix in N dimension is in an N2 dimensional vector space over the real number . Suppose we have N2 Hermitian matrices {Γμ} (each one an N × N matrix) that satisfies

Γμ=Γμ
Tr[ΓμΓν]=δμ,ν
Then an arbitrary Hermitian matrix F can be expanded using this basis set
F=μ=1N2FμΓμ
where Fμ = Tr[F · Γμ].

For the measurement of a physical quantity on a particular state |ψi〉, the mean value is

fiψi|F^|ψi=Tr[F|ψiψi|]=Tr[FPi]=μFμTr[ΓμPi]=μFμBiμ
where Pi is the projection operator of state |ψi〉 and the coefficients of the linear equations
Biμ=Tr[ΓμPi]
If we have a full set of N2 different input states {|ψi〉}, and B (an N2 × N2 matrix) is reversible (the condition of which is discussed below), then from Eq. (9) we have Fμ=i=1N2(B1)μifi and
F=μFμΓμ=μ(i(B1)μifi)Γμ=i(μ(B1)μiΓμ)fi=iMifi
where
Mi=μ=1N2(B1)μiΓμ

Here {Mi} matrices directly connect the measurement results fi to the target matrix F. If we can efficiently calculate these {Mi} matrices, tomography of F is as simple as multiplying them by the measurement results {fi} and add them up together, as indicated in Eq. (11).

Since each measurement provides one real number, and the total degrees of freedom for an arbitrary Hermitian matrix is N2, one needs a minimum of N2 different measurements.

Theorem1: B = Tr[Γμ · Pi] is reversible iff {Pi} are linearly independent

Proof: B = Tr[Γμ · Pi] is reversible means if iαiBiμ=0 for all μ, then αi = 0.

The left hand side: for all μ, 0=iαiBiμ=iαiTr[ΓμPi]=Tr[ΓμiαiPi]=Tr[Γμα]

Then using Eq. (8), we have α=μTr[αΓμ]Γμ=μ0Γμ=0.

Thus B is reversible is equivalent to if αiαiPi=0 then αi = 0, which means {Pi} are linearly independent ▪

As a result, as long as we can choose N2 input states {|ψi〉} so that the corresponding projection operators {Pi} are linearly independent, we can use Eq. (10) to calculate B and Eq. (12) to calculate {Mi} and Eq. (11) to calculate the matrix F. Here we propose a simple universal set of {|ψi〉} that has N2 states and works for an arbitrary N:{|ψ1(k),|ψ2(k1,k2),|ψ3(k1,k2)}, where

{|ψ1(k)=|k,1kN|ψ2(k1,k2)=12(|k1+|k2),1k1<k2N|ψ3(k1,k2)=12(|k1+i|k2),1k1<k2N
The same set has been used in phase retrieval in MIMO. [23,24] It is straightforward to verify that {Pi} that corresponds to this set are linearly independent.

Theorem2: {Mi} does not depend on the choice of {Γμ}, but only on the choice of the measurement set {Pi}

Proof: Suppose there is another complete set {Γ′μ}. Since both {Γμ} and {Γ′μ} are complete, they are connected by a linear reversible transformation

Γμ=νAμνΓν
The corresponding
Biμ=Tr[ΓμPi]=Tr[νAμνΓνPi]=νAμνBiν
Since Aµν is reversible, it can be easily verified that
(B1)νi=μ(B1)μi(A1)μν
Thus, the corresponding
Mi=ν(B1)νiΓνi=νμ(B1)μi(A1)μν.αAναΓα=μα(B1)μiδμαΓα=μ(B1)μiΓμ=Mi

As a result, any {Γμ} that satisfies Eqs. (7a) and (7b) will work, and will generate the same {Mi}. Once again, we propose one universal set that works for an arbitrary N:{Γ1,ij(k),Γ2,ij(k1,k2),Γ3,ij(k1,k2)}, where

{Γ1,ij(k)=δikδjk,1kNΓ2,ij(k1,k2)=12(δik1δjk2+δik2δjk1),1k1<k2NΓ3,ij(k1,k2)=i2(δik1δjk2δik2δjk1),1k1<k2N
It is straightforward to verify that this set satisfies all the requirements.

As a matter of fact, since we have such simple analytic expressions for {|ψi〉} and {Γμ}, it turns out {Mi} also have simple analytic expressions for an arbitrary N:{M1,ij(k),M2,ij(k1,k2),M3,ij(k1,k2)}, where

{M1,ij(k)=δikδjk+[(12+i2)δi<j+(12i2)δi>j](δik+δjk),1kNM2,ij(k1,k2)=δik1δjk2+δik2δjk1,1k1<k2NM3,ij(k1,k2)=i(δik1δjk2δik2δjk1),1k1<k2N
Thus, one does not even need to do any costly matrix inversion operations to calculate the mode delay operator.

To summarize, the process of using the mean signal delay method to find all the principal modes is as follows: first, one prepares N2 different input modes {|ψi〉} according to Eq. (13). One can work under any orthonormal basis. One such basis could be the eigen modes of the fiber. Second, one launches these N2 input modes into the fiber to measure their respective mean signal delay fi. Third, one uses Eq. (19) to calculate {Mi} and Eq. (11) to calculate F. Last, one calculates the eigenvectors and eigenvalues of F, which are the principal modes under the chosen basis and their respective delays. To verify that these are indeed the principal modes, one can recreate the superposition states according to the eigenvectors, under the same basis, and launch them into the same fiber. These modes should experience no temporal spreading to the first order of mode dispersion.

We numerically simulate the tomography process in Appendix A for N = 3.

Error estimation

Here we consider the propagation of error during the input state preparation and mean signal delay measurement to the principal states calculation. In the ideal case with no error, the matrix for the group delay operator (ω) (Eq. (11)) is reproduced here as

F=iMifi
where fi = 〈ψi||ψi〉. Suppose there is an error in the state preparation, and the actual input states launched are instead |ψ′i〉 = |ψi〉 + δ|ψi〉. Also, we assume that there are errors related to the mean signal delay measurement δfmes. As a result, the measured values for fis are (ignoring higher order terms)
fi=ψi|F^|ψi+δfmes=ψi|F^|ψi+ψi|F^(δ|ψi)+c.c.+δfmesfi+δfi
where δfi ≡ 〈ψi| (δ|ψi〉) + c.c. + δfmes. The measured group delay operator matrix becomes
F=iMifi=iMifi+iMiδfiF+δF
where δFiMiδfi. The real principal modes |pi〉 are the eigenstates of the group delay operator and therefore satisfies |pi〉 = τi |pi〉. (Here for simplicity, we drop the superscript that designates the input end of the fiber) On the other hand, the measured principal modes |p′i〉 satisfies F̂′|p′i〉 = τ′i|p′i〉. Using non-degenerate perturbation theory, the calculated principal modes and their corresponding group delay is
τi=τi+pi|δF^|pi+O(δ2)
|pi=|pi+jipj|δF^|piτiτj|pj+O(δ2)
The standard deviations of the group delays for the real principal modes |pi〉 are zero, because these are the eigen states of the group delay operator. The variance of the signal delay for the measured principal modes |p′i〉 are calculated as
Var(τi)=pi|F^2|pi|pi|F^|pi|2=τi2(1+pi|(δ|pi)+c.c.+O(δ2))|τi(1+pi|(δ|pi)+c.c.+O(δ2))|2=τi2(1+O(δ2))|τi(1+O(δ2))|2=O(δ2)
where we use 〈pi | (δ|pi〉) = 0. Therefore, the standard deviation of signal delay is σ(τ′i) = O(δ).

4. Tomography using projection operators

In section 3, we implement the full tomography on the operator using the mean signal delays. However, when the differences of group delays between principal modes are big enough, i.e., the output principal modes are temporally separated, we can resolve each principal mode component. In this case, the signal strength corresponding to the time delay of the jth principal mode for the ith input is

Ai,j=|pjin|ψi|2
As a result, with each input, we can measure its projection onto every principal modes. Thus, the question of the minimum number of input states necessary for the tomography turns into the minimum number of projection measurements needed to identify one particular state, up to an arbitrary phase. This problem is equivalent to the rank-1 positive-operator-valued measurement for pure states in N dimension. Even though coherent N-mode optical field is in a (N−1)-dimensional complex projective space CP(N − 1) [25], it is proven that one needs at least 3N − 2 measurements to determine unequivocally the state [26,27]. Here we explicitly construct 3N − 2 input states, and show that projection measurements on these states are enough to identify unequivocally an arbitrary pure state, thus any principal modes.
{|ψi}={|k,1kN12(|1+|k),1<kN12(|1+i|k),1<kN

Assuming that the arbitrary principal state is decomposed in the eigen modes as

|p=kck|k
From the first N measurements we obtain
fk=|k||p|2=|ck|2,1kN
Since there is an arbitrary phase, we designate c1 to be real and positive. Thus
c1=f1
For the next 2(N−1) measurements, we obtain
gk=|12(1|+k|)|p|2=12(f1+fk+2c1Re[ck]),1<kN
hk=|12(1|ik|)|p|2=12(f1+fk+2c1Im[ck]),1<kN
Using these measurements, one can calculate the amplitude of the kth component by
ck=fkexp(iϕk),1<kN
where ϕk = arg [(2gkf1fk) + i (2hkf1fk)]. For this method to work, |c1| cannot be too small, i.e., there has to be one mode with good overlap with all the principal modes. Finally, the group delay operator is reconstructed as
F^=iτi|pipi|

A numerical simulation of this method is shown in Appendix B.

5. Discussion

To experimentally implement our method, one needs a single spatial light modulator to launch the specified input states, and a fast photodetector to temporally resolve the optical signal. In the transfer matrix method, such as the one employed in [7] and [18], the phase information between different modes is retrieved through a reference beam in terms of homodyne detection. The interferometric stability degrades over long distance, and active stabilization might be necessary, making the setup more complicated. In our method, the phase information is encoded within the N2 input states (self-referenced), thus no interferometer is needed.

The speed of the photodetector and the temporal width of the input pulse limit the temporal resolution, and thus the minimum measurable delay difference. If the mode delay differences of a group of principal modes are smaller than the temporal resolution, they can be regarded as degenerate, and any orthonormal basis in this subspace can be used as principal modes. Transmitting any one of these will result in minimal temporal delay spreading. Ultimately, the minimum pulse width (the inverse of bandwidth) is limited by higher order dispersion.

One thing we would like to point out here is that, the mode delay operator, though highly related, does not provide the full information of the transfer matrix, and vice versa. For example, one cannot calculate the transfer matrix from the mode delay matrix only. As a result, our method does not measure the output principal modes. If one need to identify both input and output modes, one needs to make additional measurements, by launching the input principal modes, and characterizing the corresponding output modes.

Compared with previous signal delay method [19], our method is more straightforward to derive. Since we now have an analytic expression for the mode delay operator, we no longer need to do the costly matrix inversion of a N2 × N2 matrix. Moreover, it prevents spreading numeric errors due to ill-conditioned matrices. Also, our definition of the mode delay operator has non-zero trace, where the trace is the mean signal delay. Therefore, we can measure the mean delay without additional measurements.

If the pulses are well seperated, our second method further reduces the number of necessary measurements from quadratic N to linear N. In cases where the measurement speed is critical, for example when the environment is changing quickly, which casues the principal modes to change quickly, this becomes a crucial saving.

Our methods are described in the context of mode coupling in multi-mode fiber. However, they are applicable to many cases, as long as all the spatial modes that are coupled together are included.

6. Conclusion

In this paper, we use the mathematical similarity between characterizing mode delay operator and quantum state tomography, and develop two new methods to characterize the principal modes. Our methods require smaller number of input states compared with mainstream methods, are more robust in terms of measurement errors, and should be more economical to implement.

Appendix A: Tomography using mean delays for N=3

Here, we numerically simulate the tomography process for N = 3 modes to illustrate our method.

We construct an arbitrary 3 × 3 Hermitian matrix for the group delay operator

F=(10.2i0.20.2i00.3+0.4i0.20.30.4i0.5)
We can calculate its eigenvalues, i.e., the mode delays of the principal modes {τk } (listed in Table 1). Our goal is to recreate the matrix F using measurement results of mean signal delay of chosen input states.

Tables Icon

Table 1. Simulated temporal delays of the principal modes using mean delays

First, we construct N2 = 9 launch modes according to Eq. (13) and get

{|ψi}={(100),(010),(001),(12120),(12012),(01212),(12i20),(120i2),(012i2)}

Second, we simulate the output signal of the ith input signal according to

sigi(t)=k|pk|ψi|2rk(t)
where rk (t) is the temporal response of kth principal mode and assumed to be
rk(t)=12πσkexp[(tτk)22σk2]+
σk is the pulse width and is assumed to be 0.25. Each data point is also added a small error of Gaussian distribution with zero mean and standard deviation of 0.03. The simulated output signal is shown in Figure 1. The mean signal delay is then calculated as
fi=dttsigi(t)dtsigi(t)
and we get
{fi}={1.0370.014,0.5390.524,0.053,0.572,0.718,0.272,0.122}
Third, we calculate {Mi} according to Eq. (19) and get
{Mi}={(11+i21+i21i2001i200),(01+i201+i211+i201i20),(001+i2001+i21+i21i21),(010100000),(001000100),(000001010),(0i0i00000),(00i000i00),(00000i0i0)}
Plug this into Eq. (11), the measured matrix
F=(1.0370.002+0.192i0.196+0.023i0.0020.192i0.0140.310+0.384i0.1960.023i0.3100.384i0.539)
Compare this with our starting matrix F (Eq. (31)), the method reproduces the original matrix well.

 figure: Fig. 1

Fig. 1 Simulated signal delay for 9 different input modes: sigi(t)=k|pk|ψi|2rk(t), where rk(t)=12πσkexp[(tτk)22σk2]+. τk is its mode delay of kth principal mode, σk = 0.25 is the pulse width, is the measurement error.

Download Full Size | PDF

Last, we calculate the eigenvalues {τ′k} (listed in Table 1) and eigenvectors {|p′k〉} of F′. The measurement errors on mode delays are approximately 2%. We also include the actual mean delays {τ″k = 〈p′k||p′k〉} if these measured “principal states” are launched into the fiber, as well as the standard deviation of the mode delay σ′k = (〈p′k|2|p′k〉 − |〈p′k||p′k〉|2)1/2. They are in the range of 0.03 – 0.04. In comparison, an arbitrary state has the standard deviation on the order of 1.

Appendix B: Tomography using projection operators for N=3

Here we numerically simulate the projection method. Since N = 3, our method should use 7 different input configurations for the complete tomography.

We assume the same mode delay matrix F as in Eq. (31). Also, since our selected input sets according to Eq. (23) is a subset of the inputs in the mean signal delay method, we use the same input data as well.

We global fit the simulated signal using sigi(t)=k=13Ai,k2πσkexp[(tτk)22σk2] to obtain mode group delay for the three principal modes {τ′k} (listed in Table 3). The corresponding weights of each input Ai,k are listed in Table 2. During the fitting procedure, the parameters τk and σk are shared between different i. We also restrict all Ai,k to be non-negative, and k=13Ai,k=1 for i = 1 . . . 7.

Tables Icon

Table 2. Extracted Ai,j from the global fitting

Tables Icon

Table 3. Simulated temporal delays of the principal modes using projection operators

With these, we reconstruct the three principal modes using Eqs. (26) and (29), where we obtain

{|pk}={(0.9460.027+0.262i0.1840.041i),(0.3120.1650.801i0.426+0.230i),(0.0760.505+0.120i0.6650.532i)}
Due to the measurement and numerical errors, the obtained principal modes are not exactly orthogonal to each other. For example, |〈p′1|p′3〉| = 0.017. To obtain the orthogonal principal modes, we recalculate F according to Eq. (30) and get
F=(1.0370.002+0.192i0.196+0.023i0.0020.192i0.0140.310+0.384i0.1960.023i0.3100.384i0.539)
We also recalculate its eigenvalues {τ″k} and eigenvectors {|p″k〉} and again results are listed in Table 3.

Acknowledgments

The authors thank William A. Wood, Ioannis Roudas, and Robert Modavis for helpful discussions.

References and links

1. L. Eisenbud, Thesis (1948).

2. E. P. Wigner, “Lower limit for the energy derivative of the scattering phase shift,” Phys. Rev. 98, 145–147 (1955). [CrossRef]  

3. F. T. Smith, “Lifetime matrix in collision theory,” Phys. Rev. 118, 349–356 (1960). [CrossRef]  

4. S. Fan and J. M. Kahn, “Principal modes in multimode waveguides,” Opt. Lett. 30, 135 (2005). [CrossRef]   [PubMed]  

5. C. D. Poole and R. E. Wagner, “Phenomenological approach to polarisation dispersion in long single-mode fibres,” Electron. Lett. 22, 1029 (1986). [CrossRef]  

6. J. P. Gordon and H. Kogelnik, “PMD fundamentals: Polarization mode dispersion in optical fibers,” PNAS 97, 4541–4550 (2000). [CrossRef]   [PubMed]  

7. J. Carpenter, B. J. Eggleton, and J. Schroder, “Observation of eisenbud-wigner-smith states as principal modes in multimode fibre,” Nat. Photon. 9, 751–757 (2015). [CrossRef]  

8. M. B. Shemirani, W. Mao, R. A. Panicker, and J. M. Kahn, “Principal modes in graded-index multimode fiber in presence of spatial- and polarization-mode coupling,” J. Lightwave Technol. 27, 1248–1261 (2009). [CrossRef]  

9. K.-P. Ho and J. M. Kahn, “Statistics of group delays in multimode fiber with strong mode coupling,” J. Lightwave Technol. 29, 3119–3128 (2011). [CrossRef]  

10. C. Antonelli, A. Mecozzi, M. Shtaif, and P. J. Winzer, “Stokes-space analysis of modal dispersion in fibers with multiple mode transmission,” Opt. Express 20, 11718–33 (2012). [CrossRef]   [PubMed]  

11. S. O. Arik, D. Askarov, and J. M. Kahn, “Effect of mode coupling on signal processing complexity in mode-division multiplexing,” J. Lightwave Technol. 31, 423–431 (2013). [CrossRef]  

12. S. Betti, F. Curti, B. Daino, G. De Marchis, E. Iannone, and F. Matera, “Evolution of the bandwidth of the principal states of polarization in single-mode fibers,” Opt. Lett. 16, 467 (1991). [CrossRef]   [PubMed]  

13. M. B. Shemirani and J. M. Kahn, “Higher-order modal dispersion in graded-index multimode fiber,” J. Lightwave Technol. 27, 5461–5468 (2009). [CrossRef]  

14. G. Li, N. Bai, N. Zhao, and C. Xia, “Space-division multiplexing: the next frontier in optical communication,” Adv. Opt. Photonics 6, 413 (2014). [CrossRef]  

15. T. Cizmar and K. Dholakia, “Exploiting multimode waveguides for pure fibre-based imaging,” Nat. Comm. 3, 1027 (2012). [CrossRef]  

16. M. Ploschner, T. Tyc, and T. Cizmar, “Seeing through chaos in multimode fibres,” Nat. Photon. 9, 529–535 (2015). [CrossRef]  

17. Y. Weng, E. Ip, Z. Pan, and T. Wang, “Advanced spatial-division multiplexed measurement systems propositions-from telecommunication to sensing applications: A review,” Sensors 16, 1387 (2016). [CrossRef]  

18. W. Xiong, P. Ambichl, Y. Bromberg, B. Redding, S. Rotter, and H. Cao, “Spatiotemporal control of light transmission through a multimode fiber with strong mode coupling,” Phys. Rev. Lett. 117, 053901 (2016). [CrossRef]   [PubMed]  

19. G. Milione, D. A. Nolan, and R. R. Alfano, “Determining principal modes in a multimode optical fiber using the mode dependent signal delay method,” J. Opt. Soc. Am. B 32, 143 (2014). [CrossRef]  

20. D. F. V. James, P. G. Kwiat, W. J. Munro, and A. G. White, “Measurement of qubits,” Phys. Rev. A 64, 052312 (2001). [CrossRef]  

21. M. Paris and J. Rehacek, Quantum State Estimation, Lecture Notes in Physics (Springer-Verlag, 2004). [CrossRef]  

22. A. Forbes, Laser Beam Propagation : Generation and Propagation of Customized Light (CRC, 2014). [CrossRef]  

23. A. Agmon and M. Nazarathy, “Broadcast MIMO over multimode optical interconnects by modal beamforming,” Opt. Express 15, 13123 (2007). [CrossRef]   [PubMed]  

24. S. O. Arik and J. M. Kahn, “Direct-detection mode-division multiplexing in modal basis using phase retrieval,” Opt. Lett. 41, 4265–4268 (2016). [CrossRef]   [PubMed]  

25. W. A. Wood, W. Miller, and M. Mlejnek, “A geometrical perspective on the coherent multimode optical field and mode coupling equations,” IEEE J. Quant. Electron 51, 1–6 (2015). [CrossRef]  

26. J. Finkelstein, “Pure-state informationally complete and “really” complete measurements,” Phys. Rev. A 70, 052107 (2004). [CrossRef]  

27. S. T. Flammia, A. Silberfarb, and C. M. Caves, “Minimal informationally complete measurements for pure states,” Found. Phys. 35, 1985–2006 (2005). [CrossRef]  

Cited By

Optica participates in Crossref's Cited-By Linking service. Citing articles from Optica Publishing Group journals and other participating publishers are listed here.

Alert me when this article is cited.


Figures (1)

Fig. 1
Fig. 1 Simulated signal delay for 9 different input modes: sig i ( t ) = k | p k | ψ i | 2 r k ( t ), where r k ( t ) = 1 2 π σ k exp [ ( t τ k ) 2 2 σ k 2 ] + . τk is its mode delay of kth principal mode, σk = 0.25 is the pulse width, is the measurement error.

Tables (3)

Tables Icon

Table 1 Simulated temporal delays of the principal modes using mean delays

Tables Icon

Table 2 Extracted Ai,j from the global fitting

Tables Icon

Table 3 Simulated temporal delays of the principal modes using projection operators

Equations (45)

Equations on this page are rendered with MathJax. Learn more.

T 2 A ( x ) i 2 k z A ( x ) = 0
| ψ out ( ω ) = U ^ ( ω ) | ψ in
ω ( e i ϕ ( ω ) | p out ( ω ) ) = 0
F ^ ( ω ) | p in = τ | p in
τ g d t t E out ( t ) E out ( t ) d t E out ( t ) E out ( t ) = i d ω E out ( ω ) ω E out ( ω ) d ω E out ( ω ) E out ( ω ) = i ψ out ( ω ) | ω | ψ out ( ω ) ψ out ( ω ) | ψ out ( ω ) = i ψ in | U ^ ( ω ) ω U ^ ( ω ) | ψ in 1 = ψ in | F ^ ( ω ) | ψ in
p j out | p i out = p j in U ^ U ^ | p i in = p j in | p i in = δ i j
Γ μ = Γ μ
Tr [ Γ μ Γ ν ] = δ μ , ν
F = μ = 1 N 2 F μ Γ μ
f i ψ i | F ^ | ψ i = Tr [ F | ψ i ψ i | ] = Tr [ F P i ] = μ F μ Tr [ Γ μ P i ] = μ F μ B i μ
B i μ = Tr [ Γ μ P i ]
F = μ F μ Γ μ = μ ( i ( B 1 ) μ i f i ) Γ μ = i ( μ ( B 1 ) μ i Γ μ ) f i = i M i f i
M i = μ = 1 N 2 ( B 1 ) μ i Γ μ
{ | ψ 1 ( k ) = | k , 1 k N | ψ 2 ( k 1 , k 2 ) = 1 2 ( | k 1 + | k 2 ) , 1 k 1 < k 2 N | ψ 3 ( k 1 , k 2 ) = 1 2 ( | k 1 + i | k 2 ) , 1 k 1 < k 2 N
Γ μ = ν A μ ν Γ ν
B i μ = Tr [ Γ μ P i ] = Tr [ ν A μ ν Γ ν P i ] = ν A μ ν B i ν
( B 1 ) ν i = μ ( B 1 ) μ i ( A 1 ) μ ν
M i = ν ( B 1 ) ν i Γ ν i = ν μ ( B 1 ) μ i ( A 1 ) μ ν . α A ν α Γ α = μ α ( B 1 ) μ i δ μ α Γ α = μ ( B 1 ) μ i Γ μ = M i
{ Γ 1 , i j ( k ) = δ i k δ j k , 1 k N Γ 2 , i j ( k 1 , k 2 ) = 1 2 ( δ i k 1 δ j k 2 + δ i k 2 δ j k 1 ) , 1 k 1 < k 2 N Γ 3 , i j ( k 1 , k 2 ) = i 2 ( δ i k 1 δ j k 2 δ i k 2 δ j k 1 ) , 1 k 1 < k 2 N
{ M 1 , i j ( k ) = δ i k δ j k + [ ( 1 2 + i 2 ) δ i < j + ( 1 2 i 2 ) δ i > j ] ( δ i k + δ j k ) , 1 k N M 2 , i j ( k 1 , k 2 ) = δ i k 1 δ j k 2 + δ i k 2 δ j k 1 , 1 k 1 < k 2 N M 3 , i j ( k 1 , k 2 ) = i ( δ i k 1 δ j k 2 δ i k 2 δ j k 1 ) , 1 k 1 < k 2 N
F = i M i f i
f i = ψ i | F ^ | ψ i + δ f mes = ψ i | F ^ | ψ i + ψ i | F ^ ( δ | ψ i ) + c . c . + δ f mes f i + δ f i
F = i M i f i = i M i f i + i M i δ f i F + δ F
τ i = τ i + p i | δ F ^ | p i + O ( δ 2 )
| p i = | p i + j i p j | δ F ^ | p i τ i τ j | p j + O ( δ 2 )
Var ( τ i ) = p i | F ^ 2 | p i | p i | F ^ | p i | 2 = τ i 2 ( 1 + p i | ( δ | p i ) + c . c . + O ( δ 2 ) ) | τ i ( 1 + p i | ( δ | p i ) + c . c . + O ( δ 2 ) ) | 2 = τ i 2 ( 1 + O ( δ 2 ) ) | τ i ( 1 + O ( δ 2 ) ) | 2 = O ( δ 2 )
A i , j = | p j in | ψ i | 2
{ | ψ i } = { | k , 1 k N 1 2 ( | 1 + | k ) , 1 < k N 1 2 ( | 1 + i | k ) , 1 < k N
| p = k c k | k
f k = | k | | p | 2 = | c k | 2 , 1 k N
c 1 = f 1
g k = | 1 2 ( 1 | + k | ) | p | 2 = 1 2 ( f 1 + f k + 2 c 1 Re [ c k ] ) , 1 < k N
h k = | 1 2 ( 1 | i k | ) | p | 2 = 1 2 ( f 1 + f k + 2 c 1 Im [ c k ] ) , 1 < k N
c k = f k exp ( i ϕ k ) , 1 < k N
F ^ = i τ i | p i p i |
F = ( 1 0.2 i 0.2 0.2 i 0 0.3 + 0.4 i 0.2 0.3 0.4 i 0.5 )
{ | ψ i } = { ( 1 0 0 ) , ( 0 1 0 ) , ( 0 0 1 ) , ( 1 2 1 2 0 ) , ( 1 2 0 1 2 ) , ( 0 1 2 1 2 ) , ( 1 2 i 2 0 ) , ( 1 2 0 i 2 ) , ( 0 1 2 i 2 ) }
sig i ( t ) = k | p k | ψ i | 2 r k ( t )
r k ( t ) = 1 2 π σ k exp [ ( t τ k ) 2 2 σ k 2 ] +
f i = d t t sig i ( t ) d t sig i ( t )
{ f i } = { 1.037 0.014 , 0.539 0.524 , 0.053 , 0.572 , 0.718 , 0.272 , 0.122 }
{ M i } = { ( 1 1 + i 2 1 + i 2 1 i 2 0 0 1 i 2 0 0 ) , ( 0 1 + i 2 0 1 + i 2 1 1 + i 2 0 1 i 2 0 ) , ( 0 0 1 + i 2 0 0 1 + i 2 1 + i 2 1 i 2 1 ) , ( 0 1 0 1 0 0 0 0 0 ) , ( 0 0 1 0 0 0 1 0 0 ) , ( 0 0 0 0 0 1 0 1 0 ) , ( 0 i 0 i 0 0 0 0 0 ) , ( 0 0 i 0 0 0 i 0 0 ) , ( 0 0 0 0 0 i 0 i 0 ) }
F = ( 1.037 0.002 + 0.192 i 0.196 + 0.023 i 0.002 0.192 i 0.014 0.310 + 0.384 i 0.196 0.023 i 0.310 0.384 i 0.539 )
{ | p k } = { ( 0.946 0.027 + 0.262 i 0.184 0.041 i ) , ( 0.312 0.165 0.801 i 0.426 + 0.230 i ) , ( 0.076 0.505 + 0.120 i 0.665 0.532 i ) }
F = ( 1.037 0.002 + 0.192 i 0.196 + 0.023 i 0.002 0.192 i 0.014 0.310 + 0.384 i 0.196 0.023 i 0.310 0.384 i 0.539 )
Select as filters


Select Topics Cancel
© Copyright 2024 | Optica Publishing Group. All rights reserved, including rights for text and data mining and training of artificial technologies or similar technologies.