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

Propagation of temporal mode multiplexed optical fields in fibers: influence of dispersion

Open Access Open Access

Abstract

Exploiting two interfering fields which are initially in the same temporal mode but with the spectra altered by propagating through different fibers, we characterize how the spectral profiles of temporal modes change with the fiber induced dispersion by measuring the fourth-order interference when the order number and bandwidth of temporal modes are varied. The experiment is done by launching a pulsed field in different temporal modes into an unbalanced Mach-Zehnder interferometer, in which the fiber lengths in two arms are different. The results show that the mode mismatch of two interfering fields, reflected by the visibility and pattern of interference, is not only dependent upon the amount of unbalanced dispersion but also related to the order number of temporal mode. In particular, the two interfering fields may become orthogonal under a modest amount of unbalanced dispersion when the mode number of the fields is k ≥ 2. Moreover, we discuss how to recover the spectrally distorted temporal mode by measuring and compensating the transmission induced dispersion. Our investigation paves the way for further investigating the distribution of temporally multiplexed quantum states in fiber network.

© 2021 Optica Publishing Group under the terms of the Optica Open Access Publishing Agreement

1. Introduction

The capacity and performance of communication system can be increased and improved by encoding information on the multiple degree of freedom of optical fields. For example, the multiplexing technique of wavelength, polarization and spatial modes greatly increases the information capacity of fiber optical communication system [1]. For quantum information science (QIS), multiplexing is also an essential technology, which makes the quantum systems more scalable and practical due to increased channel capacity and tolerance to noise [24]. While the mode basis in wavelength, polarization and spatial degree of freedom have been widely exploited in the past [1,3,4], the temporal modes (TMs) which provide a convenient orthogonal basis for studying light pulses of any temporal and spectral shape have not attracted much attention until recently [2,5].

An example of such an orthogonal temporal mode (TM) basis is a set of Hermite-Gaussian modes. The first three Hermite-Gaussian modes of a TM basis are illustrated in Fig. 1(a) and 1(b). Because of the relation between time and frequency, frequency modes occupied by a TM increase with mode order number $k$ (see Fig. 1(c)). So the TM multiplexing, at the cost of wavelength division multiplexing, can not increase channel capacity. This is similar to the spatial mode multiplexing in free space [6]. However, TM analysis does offer a much more straightforward way with intrinsically decoupled modes in describing pulse-pumped parametric processes [2,7], by which a variety of quantum states of light have been experimentally generated and used in quantum information processing, including quantum communication, quantum simulation, and quantum metrology.

 figure: Fig. 1.

Fig. 1. First three Hermite-Gaussian modes in (a) the frequency domain $f_k(\omega )$ and (b) the time domain $\phi _k(t)$. (c) Frequency modes occupied by a temporal mode $\Delta \omega _k$ (normalized to $\Delta \omega _0$) varies with the order number k.

Download Full Size | PDF

It was recognized recently that the TMs of electromagnetic fields form a new framework for QIS [2], since they intrinsically span a high dimensional Hilbert space and lend themselves to integration into existing fiber communication networks. So far, a lot of efforts have been made to develop the tools for implementing photonic quantum information in TMs framework, such as the generation of quantum states by pulsed pumped parametric process for carrying information, and the manipulation of TMs to fulfill the function of multiplexing and demultiplexing. However, distribution of the temporally multiplexed quantum states in long distance optical fiber, which is a key for the success of quantum information processing in TMs framework, has not been done yet. It is generally accepted that for a set of orthogonal TMs basis, the walk-off effect can be neglected because the TMs are centering at the same frequency and the group velocities for different orders of TM are equal. However, dispersion of transmission medium will inevitably affect the shape of pulsed field. As a result, the devices designed for accomplishing the functions of TMs demultiplexing and detection will be affected. In this paper, as a first step for long-distance distribution of quantum states in higher-order TM, we study how the dispersion of optical fiber influence the TMs of electromagnetic fields by using an unbalanced Mach-Zehnder interferometer (MZI) and measuring the visibility of fourth-order interference.

The rest of the paper is organized as follows. In Sec. 2., we introduce the theory of how to characterize the spectral change of TMs with Hong-Ou-Mandel type fourth-order interference [8,9]. After briefly reviewing the fourth-order interference between two pulsed interfering fields, we analyze the fourth-order interference of an unbalanced MZI when its input is in a single TM but with different orders. In Sec. 3., we describe our experiments carried out by using the unbalanced MZI, in which the fiber lengths of two arms are different. We show how the transmission fiber induced dispersion affects the spectral profiles of the TM with different order numbers by measuring the pattern and visibility of fourth-order interference. In Sec. 4., we discuss how to measure the dispersion of transmission medium by using the fourth-order interference of unbalanced MZI, which can be used to cancel the dispersion induced spectral change of TM. Finally, we conclude in Sec. 5.

2. Theory

It is well-established that the visibility of the Hong-Ou-Mandel (HOM) type fourth-order interference depends on the degree of mode matching between two interfering fields [810]. Using this property of HOM interference, we can characterize how the spectra of temporal modes (TMs) change with the dispersion induced by transmission media.

Consider the scheme of a HOM interferometer shown in Fig. 2(a). For the sake of brevity, suppose the fields are one-dimensional so we can absorb the position variable with time and only consider the temporal variable $t$. Assuming the two input fields of the beam splitter (BS), $E$ and $\bar {E}$, are initially in the same temporal mode (TM) but with the spectra of $\bar {E}$ altered by propagating through a medium with length corresponding to multiple pulse separations, the two input fields can be written as

$$E(t)=\sum_{j} E_j(t-j\Delta T), ~~~\bar{E}(t)=\sum_{j} \bar{E}_j(t-(j+N)\Delta T),$$
where $\Delta T$ is the separation between adjacent pulses, and $N$ is the number of delayed pulses. Assume each pulse of the pulse train has an identical TM $\phi _k(t)$ but its amplitude and phase may fluctuate from pulse to pulse due to finite coherence time. Then $E_j(t)$ can be expressed as:
$$E_j(t) = A_je^{i\varphi_j} \phi_k(t),$$
where
$$\phi_k(t)=\int^\infty_{-\infty}f_k(\omega)e^{{-}i\omega t}d\omega,$$
satisfying $\int dt |\phi _k(t)|^2=1$ is the $k$-th order TM with the duration of a single pulse $\delta t\ll \Delta T$. A set of temporal modes (TMs) are orthogonal with respect to a frequency (time) integral [2]:
$$\int^\infty_{-\infty}dt\phi_k^*(t)\phi_l(t)=\int^\infty_{-\infty}d\omega f_k^*(\omega) f_l(\omega)=\delta_{k,l}.$$

The field propagating through the delay $N\Delta T$ can be written as

$$\bar{E}_j(t) = A_je^{i\varphi_j} \bar{\phi}_k(t)$$
with
$$\bar{\phi}_k(t)=\int^\infty_{-\infty}\bar{f}_k(\omega)e^{{-}i\omega t}d\omega, ~~~\bar{f}_k(\omega)=f_k(\omega)e^{{-}i\varphi(\omega) },$$
where $\varphi (\omega )$ denotes the phase shift induced by the delay of transmission medium.

 figure: Fig. 2.

Fig. 2. (a) Schematic diagram of a Hong-Ou-Mandel (HOM) interferometer. (b) Conceptual representation of the scheme for studying the influence of dispersion on temporal mode by measuring HOM type fourth-order interference of an unbalanced Mach-Zehnder interferometer. DL, delay line; BS, 50:50 beam splitter; D, detector; PZT, piezoelectric transducer; SMF, single-mode fiber.

Download Full Size | PDF

In Fig. 2(a), the fields emerging at two outputs of BS are given by:

$$E_1'(t)=[E(t)+\bar{E}(t+\Delta)]/\sqrt{2}, ~~~E_2'(t)=[\bar{E}(t+\Delta)-E(t)]/\sqrt{2},$$
where $\Delta$ is the additional adjustable delay between the fields $E$ and $\bar {E}$, which is obtained by passing one input field through a delay line (DL). The photocurrent of the detector (D1 and D2) can be expressed as:
$$\begin{aligned} i_{D1}(t)=&\int_{T_R}d\tau k(t-\tau)E_{1}'^*(\tau)E_{1}'(\tau)\approx \frac{1}{2}\sum_j k(t-j\Delta T)I_j^{(1)},\\ i_{D2}(t)=&\int_{T_R}d\tau k(t-\tau)E_{2}'^*(\tau)E_{2}'(\tau)\approx \frac{1}{2}\sum_j k(t-j\Delta T)I_j^{(2)}, \end{aligned}$$
with
$$I_j^{(1)}\equiv I_j+I_{j-N}+\Gamma_{j,N}(\Delta)+\Gamma_{j,N}^*(\Delta), ~~~I_j^{(2)}\equiv I_j+I_{j-N}-\Gamma_{j,N}(\Delta)-\Gamma_{j,N}^*(\Delta),$$
where $\Gamma _{j,N}(\Delta )\equiv \int d\tau E_j^*(\tau )\bar {E}_{j-N}(\tau +\Delta )$, and $k(t)$ is the response function of the detector with a response time of $T_R$. For the fields with pulse width $\delta t$ much smaller than $\Delta T$ and $T_R$, we have the approximation: $\int d\tau k(t-\tau )E_j^*(\tau -j\Delta T)E_i(\tau -i\Delta T)=0$ if $i\neq j$, and $\int d\tau k(t-\tau )\sum _j|E_j(\tau -j\Delta T)|^2\approx \sum _jk(t-j\Delta T)I_j$ with $I_j\equiv \int d\tau |E_j(\tau )|^2=|A_j|^2=\int d\tau |\bar {E}_j(\tau )|^2$. Moreover, we can add a piezoelectric transducer (PZT) in DL (not shown in Fig. 1(a)) so that the phase difference between two interfering fields fluctuates more than $2n\pi$ ($n$ is an integer much greater than 1) over a period $T$ covering many pulses. In this case, the second order interference terms $\Gamma _{j,N}(\Delta )$ and $\Gamma ^*_{j,N}(\Delta )$ in Eq. (9) are averaged to zero over the period $T$, i.e., $\langle \Gamma _{j,N}(\Delta )\rangle _T=0$, where $\langle \rangle _T$ is a pulse-to-pulse average. At each output of BS, the average intensity can be expressed as
$$\langle i_{D1}\rangle_T=\langle i_{D2}\rangle_T=\frac{1}{2T}\int_T dt\sum_j k(t-j\Delta T)(I_j+I_{j-N})=R_pQ\langle I_j\rangle_T,$$
where $Q=\int dt k(t)$ is the total charge for one pulse, $R_p=1/\Delta T$ is the repetition rate of the pulsed field and $\langle I_j\rangle _T\equiv \frac {1}{M}\sum _j |A_j|^2$ is the measured average intensity per pulse with $M\equiv T/\Delta T$ denoting the number of pulses to average. In general, the pulse-to-pulse average time $T$ is much longer than the delay time $N\Delta T$ so that $N\ll M$ and $\langle I_j\rangle _T \approx \langle I_{j-N}\rangle _T$.

Coincidence measurement between two detectors is

$$R_c=\frac{1}{T}\int_T dt \int_{T_R} d(\Delta t) i_{D1}(t)i_{D2}(t+\Delta t) =\frac{1}{4}R_p Q^2\langle I_j^{(1)}I_{j}^{(2)}\rangle_T,$$
here, we assume the detectors can resolve different pulses so that $T_R<\Delta T$ and $k(t-i\Delta T)k(t+\Delta t-j\Delta T)=0$ if $i\neq j$, and the delay between the two detectors $\Delta t$ is $0$. Because of the fast scan of phase difference between $E$ and $\bar {E}$, the cross terms like $\langle I_j\Gamma _{j,N}\rangle _T$ and $\langle \Gamma ^2_{j,N}\rangle _T$ etc. are averaged to zero, the fourth-order correlation term in Eq. (11) can be written as
$$\langle I_j^{(1)}I_{j}^{(2)}\rangle_T=\langle I_j^2\rangle_T+\langle I_{j-N}^2\rangle_T+2\langle I_j I_{j-N}\rangle_T-2\langle |\Gamma_{j,N}(\Delta)|^2 \rangle_T=2C\left[1-\xi\mathcal{V} (\Delta)\right],$$
where
$$\mathcal{V} (\Delta)\equiv\left| \int dt \phi_k^*(t)\bar{\phi}_k(t+\Delta)\right|^2=\left| \int d\omega f_k^*(\omega)\bar{f}_k(\omega )e^{i\omega\Delta}\right|^2$$
describes the mode matching degree between the interfering fields $E$ and $\bar {E}$,
$$\xi \equiv\frac{\langle|A_j|^2|A_{j-N}|^2\rangle_T}{\langle|A_j|^4\rangle_T+\langle|A_j|^2|A_{j-N}|^2\rangle_T}$$
is determined by the intensities and photon statistics of two interfering fields, and $C\equiv \langle |A_j|^4\rangle _T+\langle |A_j|^2|A_{j-N}|^2\rangle _T$ is associated with the intensity fluctuation of the fields. Notice that this is a special case of a general scheme of unbalanced interferometers [11], the random phase is cancelled in quantity $|\Gamma _{j,N}(\Delta )|^2$ in Eq. (12) and the result does not rely on the coherence of the input field.

The visibility of fourth-order interference, i.e., the HOM dip, is

$$V=\frac{R_c(\Delta\rightarrow\infty)-R_c(\Delta=0)}{R_c(\Delta\rightarrow\infty)}=\xi \mathcal{V}(0).$$

In Eq. (15), we obviously have the overlapping integral $\mathcal {V} (0)=\int d\omega f_k^*(\omega )\bar {f}_k(\omega )=1$ if $f_k(\omega )=\bar {f}_k(\omega )$. In this condition, the visibility of fourth-order interference or the HOM dip $V$ is maximal, we have $V_{max}=\xi$. Equation (14) shows that $\xi$ depends on the relative intensity and photon statistics of $E$ and $\bar {E}$. Since the intensities of $E$ and $\bar {E}$ are assumed to be equal, we have $\xi =1, 1/2, 1/3$ for the two fields in single photon state, coherent state and thermal state, respectively [8,10].

In general, it is difficult to obtain a thermal state or single photon state in a single TM [2,9,12,13]. However, coherent state in single TM can be straightforwardly obtained by tailoring the output of a mode-locked laser with a wave shaper. To reveal how the transmission medium induced dispersion affects the spectra of TMs, hereinafter, we will focus on analyzing the model in Fig. 2(b), in which the two interfering fields of fourth-order interference are single mode coherent state. Under this condition, the photon statistics of each interfering field is Poisson distribution, i.e, we have $\langle |A_j|^2\rangle _T=\langle |A_{j-N}|^2\rangle _T=|\alpha |^2$, $\langle |A_j|^4\rangle _T=|\alpha |^4$ and $\xi =1/2$ in Eq. (14), where $|\alpha |^2$ describes the average photon number of $E$ and $\bar {E}$ fields.

In Fig. 2(b), the fields $E$ and $\bar {E}$ are obtained by splitting the field in coherent state with a 50/50 beam splitter (BS1). So the scheme formed by two BSs in Fig. 2(b) is an Mach-Zehnder interferometer (MZI). We study the spectra dependence of TM on dispersion when the difference of the fiber lengths in two arms $z$ corresponds to the delay of multiple pulses $N\Delta T$. The input field of MZI is in a single TM with order number $k$, the frequency components for the fields in two arms are related through the relation $\bar {f}_k(\omega )=f_k(\omega )e^{-i\beta (\omega )z }$, where $\beta (\omega )$ denotes the unbalanced dispersion induced by transmission fibers and can be described by the Taylor expansion

$$\beta(\omega)=\beta_0+\beta_1(\omega-\omega_0)+\frac{1}{2}\beta_2(\omega-\omega_0)^2+\cdots.$$
In Eq. (16), $\beta _m$ $(m=0,1,2\cdots )$ denotes the $m$-th order dispersion coefficient. The zero order dispersion term is a constant irrelevant to frequency. The first-order dispersion coefficient $\beta _1$ is related to the group-velocity through the relation $V_g=1/\beta _1$ and has no influence on the pulse shape, but the terms with $m\geq 2$ change the spectral or temporal feature of a pulsed field [14].

When the frequency bandwidth occupied by the TM is relatively narrow, the influence of higher order dispersion terms $\beta _m$ ($m\geq 3$) on the evolution of pulsed field is negligibly small. For the MZI with standard single mode fibers in each arm, we have $|\beta _2\Delta \omega ^2 z|\approx 1$ and $|\beta _3\Delta \omega ^3 z|\approx 0.004$ when $z=200$ m and the bandwidth of input field centering at $1550$ nm telecom band is about $1.5$ nm [14]. In this case, the approximation $\beta (\omega )z\approx \frac {1}{2}(\omega -\omega _0)^2\beta _2 z= \varphi (\omega )$ holds, and the relation between $\bar {f}_k(\omega )$ and $f_k(\omega )$ can be rewritten as

$$\bar{f}_k(\omega)=f_k(\omega)e^{{-}i\varphi(\omega) }.$$

To understand how the unbalanced dispersion in the MZI affects the mode matching of fourth-order interference, we simulate the mode overlapping factor $\mathcal {V} (0)$ (see Eq. (13)) when the TM of input field in Fig. 2(b) takes different order number. In the simulation, a family of Hermite-Gaussian functions [2,5]

$$f_k(\omega)=i\frac{1}{\sqrt{2^kk!}}H_k\left(\frac{\omega-\omega_0}{\sqrt{2}\Delta\omega}\right)u(\omega)$$
with
$$u(\omega)=\frac{1}{\sqrt{\Delta\omega}}\frac{1}{(2\pi)^{1/4}}exp\left[-\frac{(\omega-\omega_0)^2}{4\Delta\omega^2}\right]$$
denoting a Gaussian mean field mode, are successively substituted into Eqs. (17) and (13). In Eq. (19), $\omega _0=\int \omega |u(\omega )|^2d\omega$ and $\Delta \omega ^2=\int (\omega -\omega _0)^2|u(\omega )|^2d\omega$, are the mean value and variance of the field $u(\omega )$, respectively. It is straightforward to get the analytical expression of $\mathcal {V} (0)$ for TM with order number $k=0,1,2,3$:
$$\begin{aligned} \mathcal{V} (0)=&\frac{1}{(1+B^2)^{1/2}}~(k=0), ~~~\mathcal{V} (0)=\frac{1}{(1+B^2)^{3/2}}~(k=1),\\ \mathcal{V} (0)=&\frac{(2-B^2)^2}{4(1+B^2)^{5/2}}~(k=2), ~~~\mathcal{V} (0)=\frac{9(\frac{2}{3}-B^2)^2}{4(1+B^2)^{7/2}}~(k=3), \end{aligned}$$
where $B$ is related to the bandwidth of $0$-th order mode, second order dispersion $\beta _2$ and fiber length difference $z$ through the relation:
$$B=\Delta\omega^2\beta_2z.$$

The results in Fig. 3 are obtained by assuming the values of second order dispersion, the central wavelength, and bandwidth of $0$-th order mode are $\beta _2=-20$ $\mathrm {ps^2/km}$, $\lambda _0=2\pi c/\omega _0=1533$ nm and $\Delta \lambda _0=2\sqrt {2\ln 2}\lambda _0\Delta \omega /\omega _0=1$ nm, respectively, where $c$ is the speed of light in vacuum. Figure 3(a) shows that for the case of $k=0,1$, $\mathcal {V} (0)$ always decreases with the increase of fiber length difference $z$, we have $\mathcal {V} (0)\rightarrow 0$ for $z\rightarrow \infty$. However, $\mathcal {V} (0)=0$, indicating the mode of $E$ and $\bar {E}$ is orthogonal to each other, can be achieved at some finite $z$ when $k\geq 2$. According to Eq. (20), for the case of $k=2,3$, we have $\mathcal {V} (0)=0$ under the condition of $|B|=\sqrt {2}$ and $|B|=\sqrt {2/3}$, respectively, as illustrated in Fig. 3(b). The results in Fig. 3 indicate that the mode overlapping factor is not only associated with the length difference $z$ induced dispersion, but also related to the mode structure (or order number) of TM. For the higher order TM, the dependence of $\mathcal {V} (0)$ upon $z$ becomes complicated.

 figure: Fig. 3.

Fig. 3. The mode overlapping factor $\mathcal {V} (0)$ for the input temporal mode with order number of (a) $k=0,1$ and (b) $k=2,3$ when the fiber length difference in the two arms of Mach-Zehnder interferometer, $z$, is varied.

Download Full Size | PDF

3. Experiments and results

Our experimental setup is shown in Fig. 4. The input field of the unbalanced MZI $E_0$ having arbitrarily engineered TM profile is obtained by passing the output of a mode-locked fiber laser through a properly programmed wave shaper (WS, Finisar 4000A). The repetition rate and pulse duration of the laser are 50 MHz and 100 fs, respectively, and the central wavelength is in 1550 nm telecom band. To avoid the influence of self-phase modulation in transmission optical fiber on the spectral or temporal profile of the pulsed field [14], the output of WS is heavily attenuated to the level of about one photon per pulse. After splitting the field $E_0$ into two by using BS1, the unbalanced dispersion in MZI is induced by propagating the two outputs of BS1 through standard single mode fibers (SMF) with length difference of $\sim 200$ m. Fiber polarization controller (FPC) placed in one arm is used to ensure the polarization of $E$ and $\bar {E}$ fields are well matched at BS2. The relative intensities of the fields in each arm is respectively adjusted by variable optical attenuator (VOA1 or VOA2), so that the intensities of two input fields of BS2 are equal. The two outputs of BS2 are respectively measured by the single photon detectors, SPD1 and SPD2.

 figure: Fig. 4.

Fig. 4. Experimental setup. FPBS, fiber polarization beam splitter; WS, wave shaper; ATT, attenuator; BS1-BS2, 50:50 beam splitter; VOA1-VOA2, variable optical attenuator; PZT, piezoelectric transducer; DL, delay line; FPC, fiber polarization controller; SMF, single-mode fiber; SPD1-SPD2, single photon detector.

Download Full Size | PDF

The two InGaAs-based SPDs (Langyan SPD4V5) are operated in a gated Geiger mode. The 2.5-ns gate pulses coincide with the arrival of photons at SPDs. The response time of SPDs is about 1 ns, which is 100 times longer than the pulse duration of the detected field. For each SPD, the detection efficiency is about $10 \%$, and the photon detection rate is less than 0.1 photon/pulse. The electrical signals produced by the SPDs in response to the incoming photons are reshaped and acquired by a computer-controlled analog-to-digital (A/D) board. So the individual counting rate of two SPDs, $R_1$ and $R_2$, and two-fold coincidences acquired from different time bins can be determined because the A/D card records all counting events. The fourth-order interference of the fields $E$ and $\bar {E}$ is measured by the coincidence rate originated from the same time bin, $R_c$. During the measurement, the PZT mounted on DL is scanned at a rate of 30 Hz. Under this condition, the second-order coherence effect between $E$ and $\bar {E}$ fields is averaged out. The counting rates of $R_1$ and $R_2$ recorded in the period $T$ of one second stay constant, while the two-fold coincidence counting rate $R_c$ varies with $\Delta$. For clarity, the results of $R_c$ are normalized by the accidental coincidence rate $R_{acc}$, which is the coincidence originated from adjacent time bins and is equivalent to $R_c(\Delta \rightarrow \infty )$. According to Eq. (12), the normalized coincidence is related to the mode overlapping factor $\mathcal {V} (\Delta )$ through the relation:

$$N_c(\Delta)=\frac{R_c}{R_{acc}}=\frac{R_c/R_p}{(R_1/R_p)(R_2/R_p)}=1-\xi\mathcal{V} (\Delta),$$
where $\xi =1/2$ because the input of MZI is in coherent state and the ideal mode match between $\bar {E}$ and $E$ is achievable. Note that in Eq. (22), the counting rates of $R_1$, $R_2$ and $R_c$ recorded within the time period of one second are corrected by subtracting the dark counts of SPDs.

We first measure the coincidence rate $R_c$ by varying the delay $\Delta$ when the input field $E_0$ of the unbalanced MZI is respectively in a set of TM with different order number $k$. In the experiment, the central wavelength of $E_0$ is fixed at $\lambda _0=1533$ nm, and the full width at half maximum (FWHM) of the intensity spectrum $|f(\omega )|^2$ of 0-th order Gaussian shaped beam is $\Delta \lambda _0=1$ nm. Figures 5(a), 5(b) and 5(c) plot the data of normalized coincidence $N_c=R_c/R_{acc}$, which are obtained when the order number of TM is $k=0,1,2$, respectively. The solid curves in Fig. 5 are obtained by substituting the experimental parameters into Eqs. (13), (14), (17)–(19) and (22), showing the theory predictions agree well with the experimental results. We find $N_c(\Delta )$ in each case is symmetric around the zero-delay point $\Delta =0$, but the pattern depends on the mode order number. $N_c(\Delta )$ monotonously increases with $|\Delta |$ for $k=0$. However, there exists oscillation in the pattern of $N_c(\Delta )$ for higher order mode, and the number of the oscillation peaks in each side is the same as order number $k$ due to the multiple peak structure of $\phi _k(t)$. The visibility of HOM dip is $V=45\%, 37\%, 25\%$ for the case of $k=0,1,2$ respectively. Obviously, the visibility of HOM is always less than the ideal value $50\%$, indicating the spectral or TM matching between $E$ and $\bar {E}$ is altered by the unbalanced dispersion in the two arms of MZI. Moreover, one sees that the visibility $V$ decreases with the increase of order number $k$, which means that the degree of mode mismatch increases with $k$. This is because that for a set of TM, which forms a field-orthogonal basis, the frequency bandwidth occupied by the TM increases with the order number, as illustrated in Fig. 1(c). As a result, for a fixed fiber length difference $z$, the amount of unbalanced dispersion increases with $k$.

 figure: Fig. 5.

Fig. 5. The normalized coincidence $N_c$ as a function of $\Delta$ when the input field of the unbalanced MZI is in the single temporal mode with (a) $k=0$, (b) $k=1$, and (c) $k=2$, respectively. The solid curves in each plot are the theoretical simulations and the error bars of data are within the size of data points.

Download Full Size | PDF

We then repeat the measurement of $N_c(\Delta )$ when the frequency band occupied by input field $E_0$ is the same. In the experiment, the central wavelength of $E_0$ is still $\lambda _0=1533$ nm, but the FWHM of $E_0$ field, having the order number $k=0$, $k^{\prime }=1$ and $k^{\prime \prime }=2$ respectively belonged to different mode basis, is fixed at 1.5 nm. The data for the cases of $k=0$, $k^{\prime }=1$ and $k^{\prime \prime }=2$ is shown in Figs. 6(a), 6(b) and 6(c), respectively. Also, we simulate the results by substituting the experimental parameters into Eqs. (13), (14), (17)–(19) and (22), as shown by the solid curves in Fig. 6, which well agree with the data points. We find the visibility of HOM dip is $35\%$, $45\%$ and $46\%$ for $k=0$, $k^{\prime }=1$ and $k^{\prime \prime }=2$, respectively. Different from Fig. 5, here, the visibility $V$ increases with the order number. This phenomenon can be qualitatively explained by the multiple-peak structure of the TM. When the frequency bands occupied by different TMs are the same, the effective bandwidth of the TM with lower order number is greater than that with higher order number because the number of pecks in frequency and temporal domain increases with the order number. This leads to larger $|B|=|\beta _2 \Delta \omega ^2 z|$ or smaller visibility according to Eq. (20). So the mode mismatch increases with the decrease of order number of TM. However, according to the expression of $\mathcal {V} (0)$ in Eq. (20), we think this interpretation only works when the corresponding term $|B|=|\beta _2 \Delta \omega ^2 z|$ of higher order TM is far away from $\sqrt {2}$. With the increase of $|B|$, the dependence of visibility upon the bandwidth and mode structure will become complicated.

 figure: Fig. 6.

Fig. 6. The normalized coincidence $N_c$ as a function of $\Delta$ when the input temporal mode of the unbalanced MZI occupies the same frequency resources but the order of temporal mode is (a) $k=0$, (b) $k^{\prime }=1$, and (c) $k^{\prime \prime }=2$, respectively. The solid curves in each plot are the theoretical simulations and the error bars of data are within the size of data points.

Download Full Size | PDF

To further understand how the dispersion induced by transmission medium affect the spectral profile of TM, we measure $N_c$ as a function of $\Delta$ and accordingly deduce the visibility $V$ of fourth-order interference when the bandwidth of input field is changed. In this experiment, the central wavelength of input $E_0$ is still fixed at $\lambda _0=1533$ nm, but the FWHM of $E_0$ is changed from 1.5 nm to 6 nm. For each setting of FWHM, the spectra profile of $E_0$ is tailored into $k=0$, $k^{\prime }=1$, or $k^{\prime \prime }=2$ of different TM basis. As shown in Fig. 7, the blue squares, orange circles, and green triangles correspond to the data of $V$ for $k=0$, $k^{\prime }=1$ and $k^{\prime \prime }=2$, respectively. We also simulate the visibility $V$ by substituting the experimental parameters into Eqs. (15) and (20), as shown by the dotted, dashed and solid curves in Fig. 7. We find that when the bandwidth of input is relatively narrow, the visibility for $k=0$ is the lowest. With the increase of the bandwidth, the visibility for $k=0$ starts to become higher than that for the modes occupying the same frequency band but with higher order mode number. Moreover, for the case of $k=0$ and $k^{\prime }=1$, $V$ decreases with the increase of bandwidth. While for the case of $k^{\prime \prime }=2$, we have $V\approx 0$ for $\Delta \lambda _{k^{\prime \prime }=2}=4$ nm. The inset in Fig. 7 plots $N_c$ versus $\Delta$ for $\Delta \lambda _{k^{\prime \prime }=2}=4$ nm. One sees that although the pattern of $N_c(\Delta )$ shows the visibility $V$ is approaching $0$, there exists oscillation structure symmetrically distributed around the central point $\Delta =0$. The result indicates that when the FWHM of $E_0$ field is 4 nm, $E$ and $\bar {E}$ with $k^{\prime \prime }=2$ are about orthogonal at $\Delta =0$, but the overlap between the two interfering fields may increase when $\Delta$ is away from $\Delta =0$ due to the multiple peak structure of higher order TM. The results in Fig. 7 are in agreement with the calculations shown in Fig. 3.

 figure: Fig. 7.

Fig. 7. The visibility of HOM dip $V$ as the functions of the FWHM of $k$-th mode $\Delta \lambda _k$, and the order of TM $k$. The inset plots the normalized coincidence $N_c$ versus delay $\Delta$ when the bandwidth and order number of TM are $\Delta \lambda _k=4$ nm and $k^{\prime \prime }=2$, respectively.

Download Full Size | PDF

Finally, we modify the unbalanced MZI into a balanced one and measure $N_c$ as a function of $\Delta$ when the input field $E_0$ is the same as in observing Fig. 5. In this experiment, the balanced MZI with $z=0$ is realized by either inserting a 200 m long standard SMF in the arm of $E$ field or taking out the 200 m SMF in the arm of $\bar {E}$ field. In this case, we have $f_k(\omega )=\bar {f}_k(\omega )$, which means that the mode profiles of $E$ and $\bar {E}$ are identical. In Fig. 8, the data represented by the blue squares, orange circles, and green triangles are the measured $N_c(\Delta )$ for a set of TM with $k=0,1,2$, respectively. Again, the solid curves are simulation results obtained by substituting experimental parameters into Eqs. (13), (14), (17)–(19) and (22). As expected, the simulations agree well with experimental results. In each case, the visibility of observed HOM dip is $V=50\%$. The results imply that for homodyne detection (HD) used in the information processing [15], the propagation induced distortion of the spectral or temporal profile of probe field can be compensated by passing the local oscillator of HD through the same propagation medium. Otherwise, the distortion will cause a reduced detection efficiency owning to the mode mismatching.

 figure: Fig. 8.

Fig. 8. The normalized coincidence $N_c$ measured by varying the TM $\phi _k (k=0,1,2)$ and the relative delay $\Delta$ between $E$ and $\bar {E}$ with a balanced MZI. The solid curves in each plot are the theoretical simulations and the error bars of data are within the size of data points.

Download Full Size | PDF

4. Discussion

In Sec. 2 and Sec. 3, the agreement between the experimental data and theoretical expectation indicates the dispersion parameter of transmission fiber $\beta _m$ can be obtained by fitting the data of measured $N_c(\Delta )$. Indeed, it is well known that the pattern of the HOM type fourth-order interference depends on the indistinguishability of two interfering fields [8,9]. The broadening of HOM dip in the pattern of $N_c(\Delta )$ and the decrease of the visibility due to unbalanced dispersion have already been analyzed and observed when the two Gaussian shaped fields are in thermal state [16], coherent state [17] and single photon state [10,18], respectively. Moreover, it has been shown that the second-order dispersion coefficient of transmission fiber can be measured by characterizing the broadening effect of the fourth-order interference pattern [10,16,17]. However, the investigations previously reported focus on the Gaussian shaped fields, the relation between the fourth-order interference pattern and the interfering fields in higher order temporal modes has not been studied. Indeed, in addition to conducting the theoretical predictions by substituting $\beta _2$ given by the datasheet of SMF into Eqs. (13), (14), (17)–(19) and (22), we perform best fit to the data in each plot of Fig. 5 by leaving $\beta _2$ as a free parameter. The values of second order dispersion deduced from the fitting of Figs. 5(a), 5(b), 5(c) are $-21.3,-19.6,-19.9$ $\mathrm {ps^2/km}$, respectively. Clearly, the values of deduced $\beta _2$ slightly deviate from that given by datasheet ($\beta _2=-20~\mathrm {ps^2/km}$). The main reason responsible for the deviation is that the fiber equivalent optical length difference ($z=200$ m) between the two arms of the unbalanced MZI is sensitive to temperature fluctuation and other environmental factors. We are currently taking data by using the photon counting technique, so the time spent in obtaining the data is relatively long. For each set of data, obtained by scanning the delay $\Delta$ from about $-20$ ps to $+20$ ps, it takes more than $15$ minutes. During such a time period, the exact value of $z$ will slight vary with time, which will inevitably influence the evaluation of the dispersion coefficients.

Fitting the data in each plot in Fig. 5 gives a reasonably good deduction of second order dispersion, however, we think the 0-th order TM mode is more suitable for measuring the dispersion because the pattern of fourth-order interference for $k=0$ is Gaussian shaped. From Eqs. (22), (17)–(20), we find the normalized coincidence $N_c(\Delta )$ for the TM with $k=0$ is related with $B=\Delta \omega ^2\beta _2 z$ through the formula:

$$N_c(\Delta)=1-\frac{\xi}{(1+B^2)^{1/2}}\exp{(\frac{-\Delta\omega^2\Delta^2}{1+B^2})}=1-V\exp{(\frac{-\Delta\omega^2\Delta^2}{1+B^2})}.$$
According to the definition of the FWHM ($\Delta \tau$) of HOM dip $N_c(\pm \frac {\Delta \tau }{2})=\frac {N_c{(\Delta \rightarrow \infty )}+N_c{(\Delta =0)}}{2}=1-\frac {V}{2}$, it is straightforward to deduce the relation between $\Delta \tau$ and the visibility of fourth-order interference V:
$$\Delta\tau=\frac{2\xi\sqrt{\ln 2}}{\Delta\omega V}.$$
Therefore, in the process of measuring dispersion coefficient of transmission line in one arm of MZI, instead of characterizing the full pattern of $N_c(\Delta )$ to extract out its FWHM, we can simply evaluate the visibility by measuring $N_c(\Delta =0)$ (or $R_c(\Delta =0)$). Obviously, the procedure of measuring $N_c(\Delta =0)$ is much fast and simple. Moreover, it is worth noting that our measurement of fourth-order interference does not have to rely on the photon counting technique. In Eqs. (10) and (12), there is no theoretical restriction to the average intensity (or photon number) per pulse $\langle I \rangle _T$. If the intensities of the interfering fields are low enough to avoid the harmful effect induced by self-phase modulation but high enough to make the weak photocurrents of detectors D1 and D2 (see Fig. 2) measurable [19], the data processing process will be nearly real time. In this case, the time spent in figuring out the visibility of fourth-order interference can be greatly reduced. The rapid measurement process will help to mitigate the environmental impact, leading to an improved accuracy in $\beta _2$ measurement.

When the bandwidth of input field is narrow, the influence of $\beta _m$ with $m\geq 3$ on the interference pattern is negligible, we can deduce coefficient $\beta _2$ in one arm of unbalanced MZI from the visibility of fourth-order interference. The effect of $\beta _m$ with $m\geq 3$ will come in when we enlarge the bandwidth of input field and we can measure them by increasing the bandwidth [14,20]. However, the method is somewhat complicated and we will discuss the detail elsewhere. Once the different order dispersion coefficients $\beta _m$ ($m\geq 2$) of the transmission line with the channel centered at a given wavelength are available, the spectral distortion of the TM can be recovered by compensating the dispersion with a wave shaper [21,22].

Comparing with the methods which are based on the second-order interference and often used to measure the dispersion of transmission medium, such as the white-light interferometer method and the modulation phase shift method etc. [23,24], this method has the following features: (1) the value of all higher order dispersion coefficient $\beta _m$ ($m\geq 2$) at the central wavelength of $E_0$ field can be directly obtained; (2) the length of the transmission fiber to be measured can be in the range of a few meters to at least tens kilometres.

As seen from Eq. (13), the visibility only depends on the overlap of the temporal mode functions of two interfering fields but not on the phase correlation between them. Thus, it is not sensitive to coherence time of the original input field to the MZI. Indeed, HOM interference was observed between two totally independent pulsed fields with no phase relation at all [13,16]. So, this technique is not limited to the length of delay. Indeed, the unbalanced MZI scheme used here is a special case of a more general scheme of unbalanced interferometers [11] where the unbalanced optical path is much larger than the coherence length of the interfering field and yet we can still observe interference in coincidence measurement.

5. Conclusion

In conclusion, in order to demonstrate how the dispersion induced by transmission optical fiber influence the spectral profile of different order TMs, we study the fourth-order interference of an unbalanced MZI. In the MZI, the two interfering fields of HOM interferometer are originally from one single temporal mode field but respectively propagate through two pieces of optical fiber to achieve unbalanced dispersion. The amount of unbalanced dispersion can be changed by changing the fibre length in the two arms of MZI. Both the simulation and experimental results show that the interference pattern of HOM type fourth-order interference varies with the order number of TM. For a set of TM described by Hermite-Gaussian function, the interference pattern of the $0$-th order TM is Gaussian shaped; while for the higher order mode, there exist oscillation in each side of the zero-delay point, and the number of peaks in the oscillation of one side is the same as order number $k$. Moreover, we find the visibility of fourth-order interference reflects the mode mismatching induced by unbalanced dispersion, and the relation between the mode mismatching degree and the amount of unbalanced dispersion varies with the order number and frequency bandwidth of TM. The visibility monotonically decreases with the amount of unbalanced dispersion when the order number of TM is $k=0,1$. However, for the case of $k\geq 2$, the visibility can be equal to zero when the unbalanced dispersion is modest, which means that the two interfering fields become orthogonal. In addition, we find the fourth-order interference of unbalanced MZI with the input of a single temporal mode coherent state can be used to measure the dispersion of transmission optical fiber. In particular, by simply evaluating the deviation of visibility from the maximal value of $50 \%$, the dispersion coefficients can be directly measured. Our investigation is useful for further studying the distribution of temporally multiplexed quantum states in fiber network [2,15].

Funding

National Natural Science Foundation of China (91836302, 12074283, 11874279); Tianjin Science and Technology Program (18ZXZNGX00210).

Disclosures

The authors declare no conflicts of interest.

Data availability

Data underlying the results presented in this paper are not publicly available at this time but may be obtained from the authors upon reasonable request.

References

1. D. J. Richardson, “Filling the light pipe,” Science 330(6002), 327–328 (2010). [CrossRef]  

2. B. Brecht, D. V. Reddy, C. Silberhorn, and M. G. Raymer, “Photon temporal modes: a complete framework for quantum information science,” Phys. Rev. X 5(4), 041017 (2015). [CrossRef]  

3. M. Erhard, R. Fickler, M. Krenn, and A. Zeilinger, “Twisted photons: new quantum perspectives in high dimensions,” Light: Sci. Appl. 7(3), 17146 (2018). [CrossRef]  

4. C. Fabre and N. Treps, “Modes and states in quantum optics,” Rev. Mod. Phys. 92(3), 035005 (2020). [CrossRef]  

5. M. G. Raymer and I. A. Walmsley, “Temporal modes in quantum optics: then and now,” Phys. Scr. 95(6), 064002 (2020). [CrossRef]  

6. N. Zhao, X. Li, G. Li, and J. M. Kahn, “Capacity limits of spatially multiplexed free-space communication,” Nat. Photonics 9(12), 822–826 (2015). [CrossRef]  

7. H. P. Yuen, “Multimode two-photon coherent states and unitary representation of the symplectic group,” Nucl. Phys. B, Proc. Suppl. 6, 309–313 (1989). [CrossRef]  

8. C. K. Hong, Z. Y. Ou, and L. Mandel, “Measurement of subpicosecond time intervals between two photons by interference,” Phys. Rev. Lett. 59(18), 2044–2046 (1987). [CrossRef]  

9. Z. Y. Ou, J.-K. Rhee, and L. J. Wang, “Photon bunching and multiphoton interference in parametric down-conversion,” Phys. Rev. A 60(1), 593–604 (1999). [CrossRef]  

10. X. Ma, L. Cui, and X. Li, “Hong–ou–mandel interference between independent sources of heralded ultrafast single photons: influence of chirp,” J. Opt. Soc. Am. B 32(5), 946–954 (2015). [CrossRef]  

11. Z. Y. Ou and X. Li, “General considerations in unbalanced fourth-order interference,” https://arxiv.org/abs/2111.02107.

12. W. P. Grice and I. A. Walmsley, “Spectral information and distinguishability in type-ii down-conversion with a broadband pump,” Phys. Rev. A 56(2), 1627–1634 (1997). [CrossRef]  

13. J. Su, J. Li, L. Cui, X. Li, and Z. Y. Ou, “Interference between two independent multi-temporal-mode thermal fields,” Phys. Rev. A 99(1), 013838 (2019). [CrossRef]  

14. G. P. Agrawal, Nonlinear fiber optics (Springer, 2000).

15. N. Huo, Y. Liu, J. Li, L. Cui, X. Chen, R. Palivela, T. Xie, X. Li, and Z. Y. Ou, “Direct temporal mode measurement for the characterization of temporally multiplexed high dimensional quantum entanglement in continuous variables,” Phys. Rev. Lett. 124(21), 213603 (2020). [CrossRef]  

16. X. Ma, X. Li, L. Cui, X. Guo, and L. Yang, “Effect of chromatic-dispersion-induced chirp on the temporal coherence properties of individual beams from spontaneous four-wave mixing,” Phys. Rev. A 84(2), 023829 (2011). [CrossRef]  

17. Y.-R. Fan, C.-Z. Yuan, R.-M. Zhang, S. Shen, P. Wu, H.-Q. Wang, H. Li, G.-W. Deng, H.-Z. Song, L.-X. You, Z. Wang, Y. Wang, G.-C. Guo, and Q. Zhou, “Effect of dispersion on indistinguishability between single-photon wave-packets,” Photonics Res. 9(6), 1134–1143 (2021). [CrossRef]  

18. D.-G. Im, Y. Kim, and Y.-H. Kim, “Dispersion cancellation in a quantum interferometer with independent single photons,” Opt. Express 29(2), 2348–2363 (2021). [CrossRef]  

19. Z. Zhang, S. Mouradian, F. N. Wong, and J. H. Shapiro, “Entanglement-enhanced sensing in a lossy and noisy environment,” Phys. Rev. Lett. 114(11), 110506 (2015). [CrossRef]  

20. Z. Mazzotta, S. Cialdi, D. Cipriani, S. Olivares, and M. G. Paris, “High-order dispersion effects in two-photon interference,” Phys. Rev. A 94(6), 063842 (2016). [CrossRef]  

21. C. Pulikkaseril, L. A. Stewart, M. A. Roelens, G. W. Baxter, S. Poole, and S. Frisken, “Spectral modeling of channel band shapes in wavelength selective switches,” Opt. Express 19(9), 8458–8470 (2011). [CrossRef]  

22. J. Schröder, M. A. Roelens, L. B. Du, A. J. Lowery, S. Frisken, and B. J. Eggleton, “An optical fpga: Reconfigurable simultaneous multi-output spectral pulse-shaping for linear optical processing,” Opt. Express 21(1), 690–697 (2013). [CrossRef]  

23. J. Y. Lee and D. Y. Kim, “Versatile chromatic dispersion measurement of a single mode fiber using spectral white light interferometry,” Opt. Express 14(24), 11608–11615 (2006). [CrossRef]  

24. L. Cohen, “Comparison of single-mode fiber dispersion measurement techniques,” J. Lightwave Technol. 3(5), 958–966 (1985). [CrossRef]  

Data availability

Data underlying the results presented in this paper are not publicly available at this time but may be obtained from the authors upon reasonable request.

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 (8)

Fig. 1.
Fig. 1. First three Hermite-Gaussian modes in (a) the frequency domain $f_k(\omega )$ and (b) the time domain $\phi _k(t)$. (c) Frequency modes occupied by a temporal mode $\Delta \omega _k$ (normalized to $\Delta \omega _0$) varies with the order number k.
Fig. 2.
Fig. 2. (a) Schematic diagram of a Hong-Ou-Mandel (HOM) interferometer. (b) Conceptual representation of the scheme for studying the influence of dispersion on temporal mode by measuring HOM type fourth-order interference of an unbalanced Mach-Zehnder interferometer. DL, delay line; BS, 50:50 beam splitter; D, detector; PZT, piezoelectric transducer; SMF, single-mode fiber.
Fig. 3.
Fig. 3. The mode overlapping factor $\mathcal {V} (0)$ for the input temporal mode with order number of (a) $k=0,1$ and (b) $k=2,3$ when the fiber length difference in the two arms of Mach-Zehnder interferometer, $z$, is varied.
Fig. 4.
Fig. 4. Experimental setup. FPBS, fiber polarization beam splitter; WS, wave shaper; ATT, attenuator; BS1-BS2, 50:50 beam splitter; VOA1-VOA2, variable optical attenuator; PZT, piezoelectric transducer; DL, delay line; FPC, fiber polarization controller; SMF, single-mode fiber; SPD1-SPD2, single photon detector.
Fig. 5.
Fig. 5. The normalized coincidence $N_c$ as a function of $\Delta$ when the input field of the unbalanced MZI is in the single temporal mode with (a) $k=0$, (b) $k=1$, and (c) $k=2$, respectively. The solid curves in each plot are the theoretical simulations and the error bars of data are within the size of data points.
Fig. 6.
Fig. 6. The normalized coincidence $N_c$ as a function of $\Delta$ when the input temporal mode of the unbalanced MZI occupies the same frequency resources but the order of temporal mode is (a) $k=0$, (b) $k^{\prime }=1$, and (c) $k^{\prime \prime }=2$, respectively. The solid curves in each plot are the theoretical simulations and the error bars of data are within the size of data points.
Fig. 7.
Fig. 7. The visibility of HOM dip $V$ as the functions of the FWHM of $k$-th mode $\Delta \lambda _k$, and the order of TM $k$. The inset plots the normalized coincidence $N_c$ versus delay $\Delta$ when the bandwidth and order number of TM are $\Delta \lambda _k=4$ nm and $k^{\prime \prime }=2$, respectively.
Fig. 8.
Fig. 8. The normalized coincidence $N_c$ measured by varying the TM $\phi _k (k=0,1,2)$ and the relative delay $\Delta$ between $E$ and $\bar {E}$ with a balanced MZI. The solid curves in each plot are the theoretical simulations and the error bars of data are within the size of data points.

Equations (24)

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

E ( t ) = j E j ( t j Δ T ) ,       E ¯ ( t ) = j E ¯ j ( t ( j + N ) Δ T ) ,
E j ( t ) = A j e i φ j ϕ k ( t ) ,
ϕ k ( t ) = f k ( ω ) e i ω t d ω ,
d t ϕ k ( t ) ϕ l ( t ) = d ω f k ( ω ) f l ( ω ) = δ k , l .
E ¯ j ( t ) = A j e i φ j ϕ ¯ k ( t )
ϕ ¯ k ( t ) = f ¯ k ( ω ) e i ω t d ω ,       f ¯ k ( ω ) = f k ( ω ) e i φ ( ω ) ,
E 1 ( t ) = [ E ( t ) + E ¯ ( t + Δ ) ] / 2 ,       E 2 ( t ) = [ E ¯ ( t + Δ ) E ( t ) ] / 2 ,
i D 1 ( t ) = T R d τ k ( t τ ) E 1 ( τ ) E 1 ( τ ) 1 2 j k ( t j Δ T ) I j ( 1 ) , i D 2 ( t ) = T R d τ k ( t τ ) E 2 ( τ ) E 2 ( τ ) 1 2 j k ( t j Δ T ) I j ( 2 ) ,
I j ( 1 ) I j + I j N + Γ j , N ( Δ ) + Γ j , N ( Δ ) ,       I j ( 2 ) I j + I j N Γ j , N ( Δ ) Γ j , N ( Δ ) ,
i D 1 T = i D 2 T = 1 2 T T d t j k ( t j Δ T ) ( I j + I j N ) = R p Q I j T ,
R c = 1 T T d t T R d ( Δ t ) i D 1 ( t ) i D 2 ( t + Δ t ) = 1 4 R p Q 2 I j ( 1 ) I j ( 2 ) T ,
I j ( 1 ) I j ( 2 ) T = I j 2 T + I j N 2 T + 2 I j I j N T 2 | Γ j , N ( Δ ) | 2 T = 2 C [ 1 ξ V ( Δ ) ] ,
V ( Δ ) | d t ϕ k ( t ) ϕ ¯ k ( t + Δ ) | 2 = | d ω f k ( ω ) f ¯ k ( ω ) e i ω Δ | 2
ξ | A j | 2 | A j N | 2 T | A j | 4 T + | A j | 2 | A j N | 2 T
V = R c ( Δ ) R c ( Δ = 0 ) R c ( Δ ) = ξ V ( 0 ) .
β ( ω ) = β 0 + β 1 ( ω ω 0 ) + 1 2 β 2 ( ω ω 0 ) 2 + .
f ¯ k ( ω ) = f k ( ω ) e i φ ( ω ) .
f k ( ω ) = i 1 2 k k ! H k ( ω ω 0 2 Δ ω ) u ( ω )
u ( ω ) = 1 Δ ω 1 ( 2 π ) 1 / 4 e x p [ ( ω ω 0 ) 2 4 Δ ω 2 ]
V ( 0 ) = 1 ( 1 + B 2 ) 1 / 2   ( k = 0 ) ,       V ( 0 ) = 1 ( 1 + B 2 ) 3 / 2   ( k = 1 ) , V ( 0 ) = ( 2 B 2 ) 2 4 ( 1 + B 2 ) 5 / 2   ( k = 2 ) ,       V ( 0 ) = 9 ( 2 3 B 2 ) 2 4 ( 1 + B 2 ) 7 / 2   ( k = 3 ) ,
B = Δ ω 2 β 2 z .
N c ( Δ ) = R c R a c c = R c / R p ( R 1 / R p ) ( R 2 / R p ) = 1 ξ V ( Δ ) ,
N c ( Δ ) = 1 ξ ( 1 + B 2 ) 1 / 2 exp ( Δ ω 2 Δ 2 1 + B 2 ) = 1 V exp ( Δ ω 2 Δ 2 1 + B 2 ) .
Δ τ = 2 ξ ln 2 Δ ω V .
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.