skip to main content
research-article
Public Access

Differentially Private Real-Time Release of Sequential Data

Published:07 November 2022Publication History

Skip Abstract Section

Abstract

Many data analytics applications rely on temporal data, generated (and possibly acquired) sequentially for online analysis. How to release this type of data in a privacy-preserving manner is of great interest and more challenging than releasing one-time, static data. Because of the (potentially strong) temporal correlation within the data sequence, the overall privacy loss can accumulate significantly over time; an attacker with statistical knowledge of the correlation can be particularly hard to defend against. An idea that has been explored in the literature to mitigate this problem is to factor this correlation into the perturbation/noise mechanism. Existing work, however, either focuses on the offline setting (where perturbation is designed and introduced after the entire sequence has become available), or requires a priori information on the correlation in generating perturbation. In this study we propose an approach where the correlation is learned as the sequence is generated, and is used for estimating future data in the sequence. This estimate then drives the generation of the noisy released data. This method allows us to design better perturbation and is suitable for real-time operations. Using the notion of differential privacy, we show this approach achieves high accuracy with lower privacy loss compared to existing methods.

Skip 1INTRODUCTION Section

1 INTRODUCTION

The collection and analysis of sequential data are crucial for many applications, such as monitoring web browsing behavior, analyzing daily physical activities recorded by wearable sensors, and so on. Privacy concerns arise when data is shared with third parties, a common occurrence. Toward this end, differential privacy [Dwork 2006] has been widely used to provide a strong privacy guarantee; it is generally achieved by disclosing a noisy version of the underlying data so that changes in the data can be effectively obscured.

To achieve differential privacy in sharing sequential data, a simple approach is to add independent noise to the data at each time instant (Figure 1(a)). This is problematic because of the temporal correlation in the data (see Section 3). A number of studies have attempted to address this issue. For example, [Rastogi and Nath 2010] apply Discrete Fourier Transform (DFT) of the sequence and release a private version generated using inverse DFT with the perturbed DFT coefficients; [Wang et al. 2017] propose a correlated perturbation mechanism where the correlated noise is generated based on the autocorrelation of the original sequence; [Kellaris and Papadopoulos 2013] decompose the sequence into disjoint groups of similar data, and use the noisy averages of these groups to reconstruct the original sequence; [Xiao and Xiong 2015] construct a Hidden Markov Model (HMM) from the independent-noise-added data sequence, and release the sequence inferred from the HMM; the method proposed in [Fioretto and Van Hentenryck 2019] first reconstructs the non-sampled data from perturbed sampled points and then solves a convex optimization to improve accuracy. Ghane et al. [2019] propose Trajectory Generative Mechanism that generates synthetic trajectory data under differential privacy; it first learns a probabilistic graphical generative model that encodes the trajectories, synthetic trajectory is then generated privately based on it. Gursoy et al. [2018] develope a differentially private synthetic trajectory publisher (DP-Star) for spacial data; it first reduces each trajectory into a sequence of representative points via the minimum description length principle, a synthetic trajectory is then generated privately from these representative points. Chen et al. [2011] proposes a data-dependent sanitization algorithm that generates a differentially private release for trajectory data; it first constructs a noisy prefix tree to group the sequences with same prefixes into the same branch, sanitized data is then generated from it. This approach was further extended in [Chen et al. 2012] by using a variable n-gram model to sanitize general sequential data. [Hua et al. 2015] consider more general trajectories by removing the implicit assumption used in Chen et al. [2011]; , 2012] that the trajectories contain a lot of identical prefixes or n-grams. However, all of the above studies rely on the availability of the entire sequence, so they can only be applied offline as post-processing methods. Fan and Xiong [2014] are the closest to our work, where the sequence is adaptively sampled first; Kalman/particle filters are then used to estimate non-sampled data based on the perturbed sampled data. However, it requires a priori knowledge of the correlation of the sequence.

Fig. 1.

Fig. 1. Comparison of two data release methods: \( \lbrace z_t\rbrace _{t=1}^T \) is the true sequence, \( \lbrace x_t\rbrace _{t=1}^T \) the released private sequence, \( \widehat{Z}_t \) the estimate of \( z_t \) learned from \( x_{t-1} \) , and \( \lbrace n_t\rbrace _{t=1}^T \) the added noise.

Fig. 2.

Fig. 2. A two-step illustration of the proposed method: adding noise \( n_t \) to the convex combination of estimate \( \hat{Z}_{t}(x_{t-1}) \) and true value \( z_t \) gives the released \( x_t \) .

In this paper we start from sequential data that can be modeled by first-order autoregressive (AR(1)) processes. We consider Gaussian AR(1) process as an example but the idea can be generalized to all (weakly) stationary processes. Leveraging time-invariant statistical properties of the stationary process, our proposed approach in each time step estimates the unreleased, future data from that already released, using correlation learned over time and not required a priori. This estimate is then used, in conjunction with the actual data observed in the next time step, to drive the generation of the noisy, released version of the data (Figure 1(b)). Both theoretical analysis and empirical results show that our approach can release a sequence of high accuracy with less privacy loss.

Other related works: Our work focuses on user-level privacy, where each individual generates a sequence of data and we aim to guarantee its privacy at all times. In contrast, some works in the literature study event-level privacy, which only guarantees the privacy at one time step.1 For example, [Chan et al. 2011; Dwork et al. 2010] propose binary tree mechanism which at each time outputs the sum of the binary data points seen so far under event-level privacy. Perrier et al. [2018] generalizes these works to continuous data and developed a mechanism based on binary tree mechanism. However, their method is only applicable to data sequences that obey a light-tailed distribution (i.e., distribution whose tail lies below the exponential distribution) and cannot be operated fully online (i.e., a sufficiently large time-lag is required before continual release).

Another line of research aims to develop new privacy notions for correlated data. Among them, some focus on correlation between users [Yang et al. 2015; Liu et al. 2016; Zhu et al. 2015], while others focus on correlation between data at different time steps. For example, [Song et al. 2017] propose the notion of pufferfish privacy which captures data correlation using a Bayesian Network, and Markov Quilt Mechanism was built that releases a data instance at one time step under the proposed privacy notion. Cao et al. [2019] propose temporal privacy leakage(TPL) to quantify privacy leakage of a temporal correlated sequence; they assume the probabilistic correlations between data points in every two consecutive time steps are known and based on which develop an algorithm to calculate TPL. A privacy budget allocation mechanism was also developed to convert a traditional differential privacy (DP) mechanism into TPL mechanism. In contrast to these works, our work focuses on the releasing mechanism rather than the privacy definition used to measure its efficacy. Specifically, we proposed an approach that can effectively reduce the total information leakage as data is released. The algorithmic property of our mechanism is orthogonal to the privacy notion used.

Some works focus on developing mechanisms that dynamically allocate privacy budgets over time. For example, [Kellaris et al. 2014] consider a setting where a subsequence of fixed length d is released at each time, and develops two privacy budget allocation mechanisms (i.e., Budget Distibution and Budget Absorption) that dynamically allocate privacy budget over time based on the dissimilarity between the previous released data and new data. Cao and Yoshikawa [2015] develop dynamic privacy budget allocation and approximation framework, where privacy budget decays exponentially over time and the previously released noisy data may be re-published if they are sufficiently close to the future data.

The rest of the paper is organized as follows. Section 2 presents background and preliminaries. Section 3 introduces the baseline approach and its issues. Our approach is presented and analyzed in Sections 4, 5, and 6. Section 7 presents Discussion. Experiments are presented in Section 8 and Section 9 concludes the paper. All proofs are given in the appendices.

Skip 2PRELIMINARIES Section

2 PRELIMINARIES

Consider a time-varying sequence \( \lbrace Z_t\rbrace _{t=1}^T \), where \( Z_t\in \mathbb {R} \) corresponds to a query \( \mathcal {Q}\in \mathcal {D}\rightarrow \mathbb {R} \) over a private dataset \( D_t \) at time \( t\in \mathbb {N} \), i.e., \( Z_t = \mathcal {Q}(D_t) \). The dataset \( D_t = \lbrace d_t^i\rbrace _{i=1}^N\in \mathcal {D} \) consists of data from N individuals \( (N\ge 1) \) where \( d_t^i \) denotes the data of the \( i^{th} \) individual at time step t and \( \mathcal {D} \) denotes the set of all possible datasets. Then \( d^i_{1:T} = \lbrace d_t^i\rbrace _{t=1}^T \) is the data of the \( i^{th} \) individual over T time steps and \( D = \lbrace d_{1:T}^i\rbrace _{i=1}^N \) includes sequences of N individuals over T time steps.

We assume \( \lbrace Z_t\rbrace _{t=1}^T \) can be modeled as a first-order autoregressive (AR(1)) process [Wei 2006], where the value at each time depends linearly on the value of the immediate preceding time step; we will see that the approach can be generalized to any (weakly) stationary process. The goal is to disclose/release this data in real time with privacy guarantees for each individual at all times. We denote by \( \lbrace X_t\rbrace _{t=1}^T \) the released sequence. Notationally, we will use X to denote a random variable with probability distribution \( \mathscr{F}_X(\cdot) \), x its realization and \( \hat{X}(y) \) the estimate of X given observation \( Y=y \); finally, \( X_{1:t} := \lbrace X_i\rbrace _{i=1}^t \).

2.1 First-order Autoregressive Process

AR(1) processes are commonly used for modeling a time series, among which Gaussian AR(1) process is one type that is widely used in various domains.

Theorem 1 (Gaussian AR(1) Process).

\( Z_{1:T} \) is a Gaussian AR(1) process [Wei 2006] if: (1) \( \begin{eqnarray} Z_{t} = \alpha + \rho Z_{t-1} + U_t,~ t\ge 1 \end{eqnarray} \) where \( U_t\overset{\text{i.i.d}}{\sim } \mathcal {N}(0,\sigma _u^2),~ Z_0\sim \mathcal {N}(\mu ,\sigma _z^2) \) and \( \sigma _u^2, \alpha , \rho \) are constants. If \( |\rho |\lt 1 \), then \( \lbrace Z_t\rbrace _{t=1}^T \) is a stationary Markov process with the following properties: (1) \( Z_t \sim \mathcal {N}(\mu ,\sigma ^2_z) \) with \( \mu =\frac{\alpha }{1-\rho } \) and \( \sigma _z^2 = \frac{\sigma _u^2}{1-\rho ^2} \); (2) its autocorrelation function is given by \( \text{Corr}(Z_tZ_{t-\tau }) = \text{Corr}(\tau) = \rho ^{|\tau |} \).

In this paper, Binomial AR(1) process is also studied and all results on Binomial AR(1) process are presented in Appendix A.

2.2 Differential Privacy

Theorem 2

(ε,δ- differential Privacy (DP)

[Dwork et al. 2006]) A randomized algorithm \( \mathcal {A}(\cdot) \) taking dataset D as input satisfies \( (\epsilon ,\delta) \)-differential privacy if for any D, \( \widehat{D} \) that are different in at most one individual’s data and for any set of possible output \( S \subseteq \text{range}(\mathcal {A}) \), we have \( \begin{align*} \text{Pr}(\mathcal {A}(D) \in S)\le \exp (\epsilon)\cdot \text{Pr}(\mathcal {A}(\widehat{D}) \in S) + \delta , \end{align*} \) where \( \epsilon \) and \( \delta \in [0,1] \) are privacy parameters and \( \epsilon \) bounds the privacy loss. Smaller \( \epsilon \) and \( \delta \) mean stronger privacy guarantee.

In this study we consider the setting where each individual’s data is of a sequential nature and the response to a query \( \mathcal {Q} \) over N individuals is released over T time steps as it is generated. Within this context, \( x_{1:T} = \mathcal {A}(z_{1:T}) = \mathcal {A}(\lbrace \mathcal {Q}(D_t)\rbrace _{t=1}^T) \) and a randomized algorithm \( \mathcal {A}(\cdot) \) is \( (\epsilon ,\delta) \)-differentially private if \( \mathscr{F}_{X_{1:T}|Z_{1:T}}(x_{1:T}|z_{1:T})\le \exp (\epsilon)\cdot \mathscr{F}_{X_{1:T}|Z_{1:T}}(x_{1:T}|\widehat{z}_{1:T})+ \delta \) holds for any possible \( x_{1:T} \) and any pairs of \( z_{1:T} \), \( \widehat{z}_{1:T} \) generated from D, \( \widehat{D} \), where D, \( \widehat{D} \) are datasets differing in at most one individual’s sequence.2 It suggests that the released sequence \( x_{1:T} \) should be relatively insensitive to the change of one individual’s sequential data, thereby preventing any meaningful inference on any individual from observing \( x_{1:T} \). Differential privacy has been widely adopted as the privacy notion and privacy-preserving technique for different applications such as machine learning [Wei et al. 2020; Abadi et al. 2016; Zhang et al. 2020; , 2018a; , 2018b; Khalili et al. 2021b; , 2021a], data mining [Friedman and Schuster 2010; Task and Clifton 2012], data market [Zheng 2020; Li et al. 2014; Khalili et al. 2019; , 2021c], etc.

Theorem 3

(Sensitivity of Query ℚ at Time t)

Consider a query \( \mathcal {Q}: \mathcal {D}\rightarrow \mathbb {R} \) taking a dataset as input, the sensitivity of \( \mathcal {Q} \) at t is defined as: \( \Delta \mathcal {Q}_t = \sup _{\forall {D_t}\sim {\widehat{D}_t}}|\mathcal {Q}({D_t})-\mathcal {Q}({\widehat{D}_t})| \), where \( {D_t},{\widehat{D}_t}\in \mathcal {D} \) are two datasets at t that are different in at most one individual’s data.

Since \( Z_t = \mathcal {Q}(D_t) \), \( \Delta \mathcal {Q}_t \) quantifies the maximum impact of an individual on \( Z_t \). In the rest of paper, unless explicitly stated, we consider scenarios where \( \Delta \mathcal {Q}_t \) does not change over time and use the notation \( \Delta \mathcal {Q}_t = \Delta \). For instance, if \( d_t^i \in \lbrace 0,1\rbrace \), \( \forall t \) and \( \mathcal {Q}(D_t) = \sum _{i=1}^Nd_t^i \) is the count query (e.g., daily count of patients), then \( \Delta = 1 \). It is worth noting that our method and analysis can be easily generalized to the case when \( \Delta \mathcal {Q}_t \) also changes over time.

2.3 Minimum Mean Squared Error Estimate

The minimum mean squared error (MMSE) estimate of a random variable X given observation \( Y=y \) is \( \hat{X}(y)=\text{argmin}_{h}~\mathbb {E}_X((X-h(Y))^2|Y=y) =\mathbb {E}(X|Y=y) \). If \( h(\cdot) \) is constrained to be linear, i.e., \( h(Y) = k_1Y+k_2 \), then the corresponding minimization leads to the linear MMSE (LMMSE) estimate and is given by \( \hat{X}(y) = \rho _{XY}\frac{\sigma _X}{\sigma _Y}(y-\mathbb {E}(Y)) + \mathbb {E}(X) \) with a mean squared error (MSE) \( = (1-\rho _{XY}^2)\sigma _X^2 \), where \( \rho _{XY} \) is the correlation coefficient of X and Y, \( \sigma _X^2 \), \( \sigma _Y^2 \) the variance of X, \( Y, \) respectively. Using these properties, we have the following result.

Proposition 2.1.

Consider a Gaussian AR(1) process \( Z_{1:T} \) defined by (1), the MMSE estimate of \( Z_{t+1} \) given \( Z_t=z_t \) is \( \hat{Z}_{t+1}(z_t) = \mu (1-\rho)+\rho z_{t} \), with MSE \( \sigma _z^2(1-\rho ^2) \). If we use a perturbed \( X_i = Z_i + N_i, ~ i\in \lbrace 1,\ldots ,t\rbrace \), to estimate \( Z_{t+1} \), where \( N_i\sim \mathcal {N}(0,\sigma ^2_n) \) is the added noise, then the MMSE estimate of \( Z_{t+1} \) given \( X_i = x_i \) is \( \begin{align*} \hat{Z}_{t+1}(x_i)=\mu \left(1-\rho ^{t+1-i}\frac{\sigma _z^2}{\sigma _z^2+\sigma ^2_n}\right) + \rho ^{t+1-i}\frac{\sigma _z^2}{\sigma _z^2+\sigma ^2_n}x_i~. \end{align*} \)

The (L)MMSE estimates for a Binomial AR(1) process are given in Appendix A.2. Note that for Gaussian AR(1) processes, both MMSE estimates \( \hat{Z}_{t+1}(z_t) \), \( \hat{Z}_{t+1}(x_i) \) are linear. For Binomial AR(1), the MMSE estimate \( \hat{Z}_{t+1}(z_t) \) is also linear, which may not hold for other AR(1) processes. However, due to the simple form of linear MMSE estimate and its applicability to more general random processes, we will solely focus on LMMSE estimates in this study.

Skip 3BASELINE APPROACH Section

3 BASELINE APPROACH

The baseline approach (Figure 1(a)) provides differential privacy for a sequence \( z_{1:T} \) by perturbing each \( z_t \) directly: \( x_t = z_t + perturbation \). The upper bound of the total privacy loss, \( \epsilon _T \), can be characterized as a log-likelihood ratio of the released output under two sequences, which can be decomposed as follows: \( \begin{align*} \log \frac{\mathscr{F}_{X_{1:T}|Z_{1:T}}(x_{1:T}|z_{1:T})}{\mathscr{F}_{X_{1:T}|Z_{1:T}}(x_{1:T}|\hat{z}_{1:T})} = \sum _{t=1}^T\log \frac{\mathscr{F}_{X_t|Z_t}(x_t|z_t)}{\mathscr{F}_{X_t|Z_t}(x_t|\hat{z}_t)}, \end{align*} \) where the term \( \log \frac{\mathscr{F}_{X_t|Z_t}(x_t|z_t)}{\mathscr{F}_{X_t|Z_t}(x_t|\hat{z}_t)} \) bounds the privacy loss at time t. As the total privacy loss accumulates over T time steps, balancing the privacy-accuracy tradeoff becomes more and more difficult as T increases. As long as the variance of perturbation is finite, as \( T\rightarrow \infty \), \( \epsilon _T \) inevitably approaches infinity.

We therefore propose a method that can (i) improve the privacy-accuracy tradeoff significantly, and (ii) bound the total privacy loss over an infinite horizon when the variance of perturbation is finite.

Skip 4THE PROPOSED APPROACH Section

4 THE PROPOSED APPROACH

In our proposed method, data point \( x_t \) at time step t is released based on the previously released data \( x_{t-1} \) and its true value \( z_t \) (shown in Figure 1(a)).

The idea behind our approach is based on two observations: (1) Since \( x_{t-1} \) is correlated with \( z_{t} \) through \( z_{t-1} \), we can use \( x_{t-1} \) to obtain an estimate3 of \( z_{t} \), denoted by \( \hat{Z}_t(x_{t-1}) \), and release (the perturbed version of) \( \hat{Z}_t(x_{t-1}) \) instead of \( z_t \). (2) Since differential privacy is immune to post-processing [Dwork et al. 2014], using \( x_{t-1} \) to estimate \( z_t \) does not introduce additional privacy loss. Thus, technically we can release an initial \( x_1 \) (the perturbed version of \( z_1 \)), followed by the sequence \( x_t=\hat{Z}_t(x_{t-1}), t\gt 1 \). However, doing so will lead to a fairly inaccurate released sequence compared to the original, for which the privacy loss does not accumulate over time but the estimation error does. To balance the competing needs of accuracy (having the released sequence resemble the true sequence) and privacy, one must calibrate the released version using the true values.

There are different ways to calibrate the released sequence. In this study, we shall examine the use of the convex combination \( (1-w_t)\hat{Z}_t(x_{t-1})+w_t z_t \), and the perturbed version of this as the released \( x_{t} \). Examples of other approaches to calibrating released sequences are discussed in Section 7.

The weight parameter \( w_t \) serves four purposes:

  • In addition to the perturbation \( N_t\sim \mathcal {N}(0,\sigma _n^2) \), \( w_t \) can also be tuned to better balance the privacy-accuracy tradeoff: larger \( w_t \) results in a more accurate but less private sequence. In contrast, \( \sigma _n^2 \) is the only means of controlling this tradeoff in the baseline method.

  • If MSE is the measure of accuracy, then \( w_t \) can also be used to balance the bias-variance tradeoff. For a deterministic sequence \( z_{1:T} \) with estimator \( X_t \) at t, \( Bias(X_t) = \mathbb {E}(X_t) - z_t \) and \( MSE(X_t) = \mathbb {E}((X_t-z_t)^2) \). The bias and variance can be controlled jointly by adjusting \( w_t \) and \( \sigma _n^2 \), which can result in smaller MSE as \( MSE = Variance + Bias^2 \). In contrast, \( x_{1:T} \) is always unbiased in the baseline method and \( MSE = Variance \) always holds.

  • If we keep \( w_t \) private, then the method can prevent certain attackers from knowing the detail of the perturbation mechanism, resulting in stronger protection (Section 7).

  • By adjusting \( w_t \), it is possible to release the sequence spanning an infinite horizon with bounded total privacy loss (Section 5).

4.1 Estimate of Zt with Learned Correlation

We can estimate the true value \( Z_{t} \) from \( x_{t-1} \) using the LMMSE estimate \( \hat{Z}_{t}(x_{t-1}) \) given in Section 2. However, it requires the knowledge of mean \( \mu \), variance \( \sigma ^2 \) and autocorrelation \( \rho \) of \( Z_{1:T} \), which may be unknown in reality and should be estimated. To avoid revealing more information about \( z_{1:T} \), this estimate is obtained using only the released \( x_{1:T} \), as shown in Algorithm 1, where both \( \hat{\mu } \) and \( {\hat{\sigma }}^2 \) are unbiased, and \( \hat{\rho } \) is adopted from Huitema and McKean [1991].4

4.2 Release xt with Estimate \( \hat{Z}_t(x_{t-1}) \) and True Value zt

Given the estimated parameters \( \hat{\mu }, {\hat{\sigma }}^2 \) and \( \hat{\rho } \), using results presented in Section 2, the LMMSE estimate \( \hat{Z}_{t}(x_{t-1}) \) can be approximated as \( \begin{eqnarray*} \hat{Z}_{t}(x_{t-1}) = \hat{\mu }_{t-1}\left(1-\hat{\rho }_{t-1}\frac{{\hat{\sigma }}^2_{t-1}}{{\hat{\sigma }}^2_{t-1}+\text{Var}(N_t)}\right)+\hat{\rho }_{t-1}\frac{{\hat{\sigma }}^2_{t-1}}{{\hat{\sigma }}^2_{t-1}+\text{Var(}N_t)}x_{t-1}~. \end{eqnarray*} \)

Take the convex combination of estimate \( \hat{Z}_t(x_{t-1}) \) and true value \( z_t \) with private weight \( w_t\in (0,1) \), and release: \( \begin{align*} x_t = (1-w_t)\hat{Z}_{t}(x_{t-1})+w_tz_{t} + {{\it perturbation}}~. \end{align*} \)

4.3 Privacy Mechanism

The perturbation term in the released data adds privacy protection, and existing literature provides methods on how to generate them. We shall adopt the Gaussian mechanism [Dwork et al. 2014] and bound the privacy loss in terms of perturbation.

Lemma 4.1 (Gaussian Mechanism).

Consider query \( \mathcal {Q}:\mathcal {D}\rightarrow \mathbb {R} \) with sensitivity \( \Delta \mathcal {Q} \), and the Gaussian mechanism \( \mathcal {G}(d) = \mathcal {Q}(d)+N \) which adds zero-mean Gaussian noise N with variance \( \sigma ^2 \) to the output. If \( \sigma \ge \frac{\Delta \mathcal {Q} \sqrt {2\log (1.25/\delta)}}{\epsilon } \) for \( \epsilon , \delta \in (0,1) \), then it satisfies \( (\epsilon ,\delta) \)-differential privacy.

We also propose a Binomial mechanism in Appendix A.3, which is a generalization (for arbitrary \( \Delta \mathcal {Q} \)) to the version (for the case \( \Delta \mathcal {Q} = 1 \)) first proposed in Dwork et al. [2006]. It is more suitable for discrete settings and doesn’t have a restriction on \( \epsilon \lt 1 \).

The complete procedure of our method is illustrated in Figure 3 and given in Algorithm 2, where \( n_t \) is a realization of Gaussian noise \( N_t \) (resp. Binomial noise defined in Appendix A.3) when adopting the Gaussian (resp. Binomial) mechanism. \( D_H \) in Figure 3 represents the history data that can be used for estimating parameters but won’t be revealed during this time horizon.

Fig. 3.

Fig. 3. Flowchart of the complete procedure: at each time step t, true data point \( z_t \) is first estimated based on previous released data \( x_{t-1} \) ; the estimation \( \hat{Z}_t(x_{t-1}) \) is then calibrated with true data \( z_t \) using convex combination, i.e., \( (1-w_t)\hat{Z}_t(x_{t-1})+w_tz_t, w_t\in (0,1) \) ; finally, the noisy version of the calibrated result (i.e., \( x_t \) ) is released.

Note that the Gaussian/Binomial mechanism only specifies the privacy parameters over one time step. In the next section we specify \( (\epsilon _T,\delta _T) \) over T steps.

Skip 5PRIVACY ANALYSIS Section

5 PRIVACY ANALYSIS

Next, we bound the total privacy loss when \( X_{1:T} \) is released using Algorithm 2. Since the total privacy loss is accumulated over T steps, various composition methods can be applied to calculate \( (\epsilon _T,\delta _T) \). We use the moments accountant method from Abadi et al. [2016] when \( N_t \) is Gaussian; the corresponding result is given in Theorem 5.1. In Appendix A.4, we use the composition theorem from Kairouz et al. [2017] when \( N_t \) is Binomial with the corresponding result given in Theorem A.5.

Theorem 5.1.

Let \( Z_t = \mathcal {Q}(D_t) \) and \( \Delta \) be the sensitivity of \( \mathcal {Q} \), \( \forall t \). Consider Algorithm 2 using zero-mean Gaussian noise with \( \text{Var}(N_t) = \sigma _{n}^2 \), \( \forall t \), that takes sequence \( z_{1:T} \) as input and outputs \( x_{1:T} \). The following holds.

(i) Given any \( \epsilon _T\ge \frac{\Delta ^2}{2\sigma _{n}^2}\sum _{t=1}^Tw_t^2 \), the algorithm satisfies \( (\epsilon _T,\delta _T) \)-differential privacy for \( \begin{align*} \delta _T = \exp \left(\left(\frac{\frac{\Delta ^2}{\sigma _{n}^2}\sum _{t=1}^Tw_t^2}{4}-\frac{\epsilon _T}{2}\right)\left(\frac{\epsilon _T}{\frac{\Delta ^2}{\sigma _{n}^2}\sum _{t=1}^Tw_t^2}-\frac{1}{2}\right)\right). \end{align*} \)

(ii) Given any \( \delta _T\in (0,1) \), the algorithm satisfies \( (\epsilon _T,\delta _T) \)-differential privacy for \( \begin{align*} \epsilon _T = 2\sqrt {\frac{\Delta ^2}{2\sigma _{n}^2}\sum _{t=1}^Tw_t^2\log (\frac{1}{\delta _T})}+\frac{\Delta ^2}{2\sigma _{n}^2}\sum _{t=1}^Tw_t^2. \end{align*} \)

Theorem 5.1 says that if a sequence of noisy data is released following Algorithm 2 and the noise has variance \( \sigma _n^2 \), then with probability \( 1-\delta _T \), the total amount of privacy loss incurred on each individual over T time steps is bounded by \( \epsilon _T \). Here \( \frac{\sigma _n}{\Delta } \) represents the degree of perturbation and \( w_t \) is the weight on the true value. Smaller perturbation and larger weight result in higher privacy loss. Because of the mapping between \( \sigma _n^2 \) and \( (\epsilon _T,\delta _T) \), we have the following result.

Corollary 5.2.

Let \( \lbrace w_t\rbrace _{t=1}^T \) be the weights used in generating \( x_{1:T} \) in Algorithm 2. To satisfy \( (\epsilon _T,\delta _T) \)-differential privacy, the variance of the Gaussian noise should be: \( \begin{align*} \sigma ^2_n \ge \frac{\Delta ^2\sum _{t=1}^Tw_t^2}{2\epsilon _T+4\ln \frac{1}{\delta _T} - 4\sqrt {(\ln \frac{1}{\delta _T})^2 + \epsilon _T\ln \frac{1}{\delta _T}}}. \end{align*} \)

To guarantee \( (\epsilon _T,\delta _T) \)-differential privacy, the noise magnitude will depend on both \( w_t \) and \( \Delta \). Larger sensitivity means larger impact of each individual on the released information and thus requires more perturbation for privacy protection; larger weights mean higher reliance on the true value in the released information, thus more perturbation is needed.

Note that Algorithm 2 reduces to the baseline approach when \( w_t = 1,~\forall t \). Theorems 5.1, A.5 and Corollary 5.2 also hold for the baseline method if we set \( w_t = 1,~\forall t \). When the noise variance is finite, using the baseline method we have \( \forall \delta _T \), \( \epsilon _T\rightarrow \infty \) as \( T\rightarrow \infty \). However, under the proposed method, it is possible that \( \lim _{T\rightarrow \infty }\epsilon _T \lt \infty \) by controlling \( w_t \), e.g., by taking \( w^2_t=ar^{t-1}, r\in (0,1) \) as a decreasing geometric sequence, we have \( \lim _{T\rightarrow \infty }\sum _{t=1}^Tw^2_t=\lim _{T\rightarrow \infty }\sum _{t=1}^Tar^{t-1} = \frac{a}{1-r} \), which leads to a bounded \( \epsilon _T \) even when \( T\rightarrow \infty \) (Theorem 5.1).

Skip 6ACCURACY ANALYSIS Section

6 ACCURACY ANALYSIS

In this section, we compare the accuracy of our method and the baseline method using the MSE measure, defined as \( \mathbb {E}_{X_{1:T}}(||x_{1:T}-z_{1:T} ||^2). \)

For simplicity of exposition, the analysis in this section is based on the assumption that the true values of parameters (\( \rho , \mu , \sigma ^2 \)) of the underlying process are known. Additional error introduced by estimating parameters in Algorithm 1 is examined numerically in Section 8. In addition, we will only present the case of Gaussian AR(1) process and \( \text{Var}(N_t) = \sigma _{n}^2, \forall t \).

Theorem 6.1.

Let the sequence \( z_{1:T} \) be generated by the Gaussian AR(1) process \( Z_{1:T} \) with \( Z_t\sim \mathcal {N}(\mu ,\sigma _z^2) \) and \( \text{Corr}(Z_tZ_{t-\tau }) = \rho ^{|\tau |} \), \( \forall t \). Let \( x_{1:T} \) be the sequence released by Algorithm 2. Then \( \mathbb {E}_{X_{1:T}}(||x_{1:T}-z_{1:T} ||^2) \) is given by \( \begin{align*} \underbrace{\sigma _z^2\left(1- \rho ^2\frac{\sigma _z^2}{\sigma _z^2+\sigma ^2_n}\right)\sum _{t=1}^T(1-w_t)^2}_{\text{estimation error}}+ \underbrace{T\sigma _n^2}_{\text{perturbation error}}. \end{align*} \)

Theorem 6.1 suggests that the total error consists of two parts: (i) estimation error and (ii) perturbation error. For the former, a sequence with stronger autocorrelation (larger \( \rho \)) enables more accurate estimate, resulting in lower estimation error. Further, higher weight on the true value \( z_t \) (larger \( w_t \)), or less perturbation (smaller \( \sigma _n^2 \)), also lowers the estimation error.

Theorem 6.2 below further compares the privacy-accuracy tradeoff of the two methods, where MSE is compared under the same privacy parameters \( (\epsilon _T,\delta _T) \).

Theorem 6.2.

Let sequential data \( z_{1:T} \) be generated by the Gaussian AR(1) process \( Z_{1:T} \) with \( Z_t\sim \mathcal {N}(\mu ,\sigma _z^2) \) and \( \text{Corr}(Z_tZ_{t-\tau }) = \rho ^{|\tau |} \), \( \forall t \). Let \( x^{A}_{1:T} \), \( x^{B}_{1:T} \) be the sequences released by Algorithm 2 and the baseline method, respectively. Let \( (\sigma _n^2)^A \), \( (\sigma _n^2)^B \) be the corresponding noise variance. Suppose both outputs satisfy \( (\epsilon _T,\delta _T) \)-differential privacy, then (2) \( \begin{eqnarray} \frac{T}{ (\sigma _n^2)^B} = \frac{\sum _{t=1}^{T}w_t^2}{ (\sigma _n^2)^A} = \frac{2\epsilon _T+4\ln \frac{1}{\delta _T} - 4\sqrt {(\ln \frac{1}{\delta _T})^2 + \epsilon _T\ln \frac{1}{\delta _T}}}{\Delta ^2}~. \end{eqnarray} \) Furthermore, \( \exists \) \( \lbrace w_t\rbrace _{t=1}^T, w_t\in (0,1) \) and \( (\sigma _{n}^2)^A \), that satisfy Equation (2) and with which \( x^{A}_{1:T} \) is more accurate than \( x^{B}_{1:T} \).

Moreover, if a constant weight \( w_t = w, \forall t \) is used, then \( x^{A}_{1:T} \) is more accurate than \( x^{B}_{1:T} \) if (3) \( \begin{eqnarray} w\gt \frac{1-(\sigma _n^2)^B/\sigma _z^2}{1+(\sigma _n^2)^B/\sigma _z^2}~. \end{eqnarray} \)

As mentioned earlier, when \( w_t = 1,\forall t \), Algorithm 2 reduces to the baseline method, and \( x^A_{1:T} \) and \( x^B_{1:T} \) become equivalent. Theorem 6.2 shows that our method can strictly improve the privacy-accuracy tradeoff by controlling \( w_t\in (0,1) \). It also provides the guidance on how to select a constant weight \( w_t = w, \forall t \), to guarantee this improvement from Equation (3): (i) If \( (\sigma _n^2)^B\gt \sigma _z^2 \), i.e., the privacy requirement is high and large perturbation is needed, then our method can always outperform the baseline regardless of the choice of \( w\in (0,1) \). In particular, if choosing \( w\rightarrow 0 \), our method will have large estimation error, but privacy can be provided with insignificant perturbation; the overall error is dominated by the estimation error, which is still smaller than the perturbation error in the baseline. (ii) If \( (\sigma _n^2)^B\lt \sigma _z^2 \), then w should be sufficiently large to maintain accuracy.

Skip 7DISCUSSION Section

7 DISCUSSION

Generalization: The proposed method is not limited to AR(1) processes; it can be applied to any (weakly) stationary random process. This is because the LMMSE estimate only depends on the mean, variance and correlation of the random process. The methodologies used in Sections 5 and 6 are also not limited to AR(1) processes. For example, the bound in Theorem 5.1 is not limited to Gaussian AR(1). The error bound derived in Theorem 6.1 only depends on the MSE of \( \hat{Z}_t(x_{t-1}) \) at each time step (i.e., term 3 in Appendix B.4). That is, this bound can be generalized to sequences following other random processes as long as we can quantify their MSE. In Section 8, the real-world datasets used in the experiments do not necessarily follow AR(1), but our method is shown to achieve better performance.

Robustness against certain attacks: Differential privacy is a strong privacy guarantee and a worst-case measure, as it bounds privacy loss over all possible outputs and inputs. In practice, how much information about \( z_{1:T} \) can really be inferred by an attacker depends on how strong it is assumed to be. An attacker is able to infer more information with higher confidence if they know the exact perturbation mechanism used in generating \( x_{1:T} \), i.e., \( \text{Pr}(X_t|Z_t) \). Specifically, real data sequence \( z_{1:T} \) and observations \( x_{1:T} \) form a Hidden Markov Model (HMM); if an attacker knows the transition probability \( P(Z_{t+1}|Z_t) \) and emission probability \( P(X_t|Z_t) \), then they can infer hidden states \( z_{1:T} \) based on observations \( x_{1:T} \) using dynamic programming such as the Viterbi algorithm [Forney 1973].

Therefore, if an attacker knows the noise distribution \( \mathcal {N}(0,\sigma _n^2) \), then they will know \( \text{Pr}(X_t|Z_t) \) automatically with the baseline method, i.e., \( X_t|Z_t \sim \mathcal {N}(Z_t,\sigma _n^2) \). However, with our method, \( X_t|Z_t \sim \mathcal {N}(w_tZ_t + (1-w_t)\hat{Z}_t(x_{t-1}),\sigma _n^2) \); thus if \( w_t \) is private and unknown to the attacker, then \( \text{Pr}(X_t|Z_t) \) cannot be readily inferred even when they know the noise distribution. As a result, in practice our method can prevent this class of attackers from knowing the details of the perturbation mechanism, thus can be stronger.

Impact of estimating parameters from a noisy sequence: The analysis in Section 6 shows that when the true parameters of the underlying process are known, our algorithm can always outperform the baseline method. However, these may be unknown in reality and need to be estimated from the released sequence using Algorithm 1, which leads to additional estimation error. Nevertheless, this can still outperform the baseline method. Consider the extreme case where \( (\sigma _n^2)^A\rightarrow +\infty \). The LMMSE estimate from the noisy data \( \hat{Z}_{t}(x_{t-1})\rightarrow \mathbb {E}(Z_{t}) \approx \hat{\mu }_{t-1} \). Since the added noise is zero-mean, with enough released data \( \hat{\mu }_{t-1} \) can attain sufficient accuracy. Then \( x_t \) determined by both \( \hat{\mu }_{t-1} \) and true \( z_{t} \) before adding noise becomes a filtered version of the true sequence, and its accuracy after adding noise will still be higher than the baseline method under the same privacy measure; this point is further validated by experiments in Section 8.

Other approaches to calibrating released sequence: We have used the convex combination of estimate \( \hat{Z}_t(x_{t-1}) \) and true data \( z_t \) to calibrate the released data. This method is effective and easy to use and analyze. In particular, the weight in the convex combination provides an additional degree of freedom and serves four purposes (Section 4). There are also other approaches to calibrating the released sequence. For example, we can leverage all released points to estimate new data, and use a sequence of estimates to calibrate, i.e., \( \sum _{i=1}^{t-1} w_i\hat{Z}_t(x_{i}) + w_tz_t \). One could also use a non-linear combination to calibrate, e.g., \( w_tz_t + (1-w_t)\sqrt {z_t \hat{Z}_t(x_{t-1})} \).

Skip 8EXPERIMENTS Section

8 EXPERIMENTS

In this section, we compare the privacy-accuracy tradeoff of our method with other methods using real-world datasets. Unless explicitly mentioned, fixed weights, \( w_t = w \), \( \forall t \), are used in the proposed method.

Methods: For comparison, in addition to the baseline method, we also consider the following.

  • Baseline-Laplace: Laplace noise \( n_t\sim Lap(0,\frac{ T\Delta }{\epsilon _T}) \) is added to \( z_t \) independently at each time step.

  • FAST without sampling [Fan and Xiong 2014]:5 Laplace noise \( n_t\sim Lap(0,\frac{ T\Delta }{\epsilon _T}) \) is first added to \( z_t \), then a posterior estimate of each \( z_t \) using the Kalman filter is released. Since it assumes the time series follows a random process \( Z_{t+1} = Z_{t}+U_t \) with \( U_t\sim \mathcal {N}(0,\sigma _u^2) \), to use the Kalman filter it requires \( \sigma _{u}^2 \) to be known in advance. Moreover, it also needs to use a Gaussian noise \( \tilde{n}_t\sim \mathcal {N}(0,\sigma _{app}^2) \) to approximate the added Laplace noise \( n_t \). In our experiments, \( \sigma _{app}^2 \) is chosen based on the guidelines provided in [Fan and Xiong 2014] and \( \sigma _u^2 \) that gives the best performance is selected using exhaustive search.

  • DFT [Rastogi and Nath 2010]: Discrete Fourier Transform is applied to the entire sequence first, then among T Fourier coefficients \( DFT(z_{1:T})_j = \sum _{i=1}^T\exp (\frac{2\pi \sqrt {-1}}{T}ji)x_i \), \( j\in [T] \), it selects the top d and perturbs each of them using Laplace noise \( \frac{\sqrt {dT}\Delta }{\epsilon _T} \). Lastly, it pads \( T-d \) 0’s to this perturbed coefficients vector and applies Inverse Discrete Fourier Transform. In our experiments, d that gives the best performance is selected from \( \lbrace 1,\ldots ,T\rbrace \) using exhaustive search.

  • BA and BD [Kellaris et al. 2014]: Two privacy budget allocation mechanisms, Budget Distribution (BD) & Budget Absorption (BA), are used to dynamically allocate privacy budget over time based on the dissimilarity between the previously released data and the new data. The new private data is released at each time step only when the data is sufficiently different from the previously released data; otherwise, the previous data is recycled and released again. The idea is to improve accuracy by allocating more privacy budgets to the most important data points.

  • Binary tree mechanism [Chan et al. 2011]: binary tree mechanism at each time outputs the approximate sum of the data points seen so far, i.e., \( B_t = \sum _{i=1}^tz_i+\text{noise} \), while preserving event-level privacy.6 The idea is to first internally group the data arrived so far based on the binary representation of current time t, the partial sum (p-sum) of data within each group can be computed and perturbed. Finally, sum over all these noisy p-sum’s gives the result. To compare with our method, we generate sequence \( b_{1:T} \) based on the summations released by binary tree mechanism \( B_{1:T} \), i.e., \( \forall t, b_t=B_t-B_{t-1} \). Then the performance of sequence \( b_{1:T} \) is compared with the sequence released by our method \( x_{1:T} \).

Real-world Datasets: We use the following datasets in our experiments.

  • Ride-sharing counts [Fanaee-T and Gama 2013]: this is generated using historical logs from the Capital Bikeshare system in 2011. It includes the counts of rented bikes aggregated on both an hourly and daily basis. Because each data point is a count over a dataset, query sensivity \( \Delta =1 \).

  • NY traffic volume counts in 2011 [DOT 2011]: this is collected by the Department of Transportation (DOT). It contains the counts of traffic in various roadways from 12AM to 1PM on an hourly basis each day. We aggregate the counts from all roadways and concatenate sequences from different days in chronological order. Because each data point is a count over a dataset, query sensivity \( \Delta =1 \).

  • Federal Test Procedure (FTP)drive cycle [EPA 2008]: this dataset includes a speed profile for vehicles and simulates urban driving patterns. It can be used for emission certification and fuel economy testing of vehicles in the United States. Because each data point is the speed of one vehicle, query sensitivity \( \Delta \) is the range of the vehicle’s speed. In this setting, \( D_t \) only includes one data point and the definition of differential privacy (Definition 2.2) is reduced to that of local differential privacy [Kasiviswanathan et al. 2011].

Accuracy metric: We use relative error (RE) defined as the normalized MSE to measure the accuracy of \( x_{1:T} \): \( \begin{align*} RE(z_{1:T},x_{1:T}) = \frac{1}{T}\frac{||z_{1:T}-x_{1:T}||_2}{\max _{1\le t\le T} |z_t|}~. \end{align*} \)

8.1 Comparison with other Methods

The comparison results are shown in Figure 4, where we use \( \delta _T = 10^{-7} \) in baseline-Normal and the proposed method; \( \Delta = 1 \) as each data point \( z_t \) is a count over a dataset. The left plot compares the relative error achieved by different methods under the same \( \epsilon _T \). However, the baseline-Laplace, FAST and DFT methods satisfy \( (\epsilon _T,0) \)-differential privacy while the baseline-Normal and proposed methods satisfy \( (\epsilon _T,10^{-7}) \)-differential privacy. Although \( \delta _T = 10^{-7} \) appears small, the total privacy loss \( \epsilon _T \) under these methods are calculated using different composition methods. Comparing different methods solely based on \( \epsilon _T \) may not be appropriate as the improvement in \( \epsilon _T \) may come from the composition strategy rather than the algorithm itself.

Fig. 4.

Fig. 4. Comparison of different methods.

To address this issue, we add the right plot in Figure 4, where the noises in baseline-Laplace and baseline-Normal are chosen such that the error achieved by baseline-Normal is no less than that under baseline-Laplace, i.e., the black curve is slightly over the green curve in the plot. This would guarantee that baseline-Normal provides stronger privacy than baseline-Laplace. By further controlling the proposed method to have the same privacy as baseline-Normal (noise variances in two methods satisfy Equation (2)), and FAST and DFT to have the same privacy as baseline-Laplace, we can guarantee that the proposed method is at least as private as FAST and DFT. In the plot, the x-axis denotes the variance of added noise in baseline-Laplace and the noise parameters of the other methods are selected accordingly. It shows that the proposed method outperforms FAST; the improvement is more significant when the privacy requirement is high. While generally DFT performs better than the proposed method, it is an offline method which requires the entire sequence to be known a priori. However, as perturbation increases (more private), the proposed method can achieve similar performance as DFT.

The DFT method can also be adapted online. One way to do this is to perform DFT over a subsequence of length \( T_{delay} \ll T \) (data released with delay \( T_{delay} \)). We examine the performance of such a method on the Traffic dataset by comparing it with DFT and baseline-Laplace. Figure 5 shows that when \( T_{delay} = 0 \) (data released in real-time, DFT applied to one data point each time and on one coefficient), the performance is similar to baseline-Laplace; as \( T_{delay} \) increases, its accuracy increases at the expense of increased delay. Because the performance of the proposed algorithm falls between baseline-Laplace and offline DFT (Figure 4), there exists a delay \( 0\lt T_{delay}\lt T \) such that the sequence released using DFT with delay \( T_{delay} \) and the proposed method have similar performance.

Fig. 5.

Fig. 5. Comparison with Online DFT.

Figure 6 shows the private traffic counts generated using various methods. For each method, we repeat the experiment 10 times and obtain 10 sample paths \( \lbrace x_{1:T}\rbrace ^{10}_{k=1} \). The curves in the plot show the average \( \frac{1}{10}\sum _{k=1}^{10}x_{1:T}^k \) while the shaded area indicates their variance whose upper and lower bound at each t are \( \max _k x_t^k \) and \( \min _k x_t^k \), respectively.

Fig. 6.

Fig. 6. Sequences aggregated from 10 runs of experiments using different methods under the same \( \epsilon _T \) (upper plot). In the lower plot, noise variance is selected in each method such that the proposed method and baseline-Normal are at least as private as FAST and baseline-Laplace.

We also compare our proposed method with BA and BD proposed in Kellaris et al. [2014]. Unlike our model, where a single query is released at every time step, BA and BD are designed to release a vector of length d each time. Moreover, BA and BD adopt \( (\epsilon ,0) \)-differential privacy. In order to compare with our method, we set \( d=1 \) and use baseline-Laplace and baseline-Normal as two baselines. Specifically, we choose noises for different methods such that: (1) our proposed method and baseline-Normal have the same privacy guarantee; (2) BA, BD, and baseline-Laplace have the same privacy guarantee; and (3) baseline-Normal is at least as private as baseline-Laplace. The results are shown in Figure 7, where the y-axis indicates the averaged relative error of 10 independent runs of experiment and x-axis is the privacy loss per time step under baseline-Laplace. As illustrated, our method outperforms others. It is worth noting that BA and BD may not even outperform baseline-Laplace. This is because in both BA and BD, half of the privacy budget is assigned to measure the dissimilarity between previously released data and new data; thus only half of the privacy budget is left for releasing the sequence. Moreover, as mentioned, BA and BD are meant for releasing a vector, especially when d is large; the error of the released sequence can be large when d is small (Theorems 6 and 7 in Kellaris et al. [2014]). It further suggests that in settings where only a single query is released (\( d=1 \)), BA and BD may not be suitable.

Fig. 7.

Fig. 7. Comparison with BA and BD [Kellaris et al. 2014].

We then compare our method with the binary tree (BT) mechanism proposed in Chan et al. [2011]. Figure 8 compares the performance of different algorithms on Traffic volume counts data, similar results are observed for other datasets. Since the BT mechanism adopts \( (\epsilon ,0) \)-differential privacy, we use the same strategy above and take baseline-Laplace and baseline-Normal as two baselines, and choose noises such that the proposed method is at least as private as the BT mechanism. It shows that ours is significantly better than BT, and the performance of BT is even worse than baseline-Laplace. This is mainly because the BT mechanism focuses on a different setting: (1) it is proposed for releasing the sum of data points seen so far, i.e., \( \sum _{i=1}^tz_i \), which means that the data point \( z_t \) needs to be repeatedly queried \( \forall t^{\prime }\ge t \), and (2) the BT mechanism aims at controlling event-level privacy (which our comparison is on user-level privacy), which generally requires more perturbation depending on how events are defined.

Fig. 8.

Fig. 8. Comparison with Binary Tree Mechanism.

8.2 Impact of Parameters ρ and wt

As mentioned earlier, the baseline is a special case (\( w_t=1 \), \( \forall t \)) of our method, which can always outperform the former with better tuned weights. The achievable improvement depends on the correlation of the sequence. We show this in Figure 9, the error of various synthetic sequences using different weights under the same privacy \( \epsilon _T \). Each sequence follows Gaussian AR(1) with \( Z_t\sim \mathcal {N}(0,1) \) but the correlation \( \rho \) varies from 0.1 to 0.9. It shows that (i) in all cases, one can find weights for our method to outperform the baseline; sequences with high \( \rho \) have the highest accuracy under the same \( \epsilon _T \); (ii) with weak (resp. strong) privacy as shown on the right (resp. left), the smallest weights that can give improvement are close to 1 (resp. 0) and the achievable improvement is small (resp. large) as compared to the baseline. As released data depends less (resp. more) on estimates when weights are large (resp. small), the correlation within the sequence does not (resp. does) affect performance significantly. In the right (resp. left) plot with weak (resp. strong) privacy, curves with lowest error are similar under different \( \rho \) (resp. decreases in \( \rho \)).

Fig. 9.

Fig. 9. Impact of correlation on performance.

We also examine the impact of estimating parameters from a noisy sequence; the result is shown in Figure 10, where Gaussian AR(1) sequences are generated. Red curves represent the relative error achieved using the proposed method where \( \hat{\mu }_t \), \( \hat{\sigma }^2_{t} \) and \( \hat{\rho }_t \) at each time are estimated from the previous released sequence; blue curves represent the case where we use true parameters \( \mu \), \( \sigma ^2 \), \( \rho \) to estimate \( z_t \) using \( x_{t-1} \). As expected, estimating parameters from a noisy sequence degrades the performance. However, even with this impact the proposed method continues to outperform the baseline significantly.

Fig. 10.

Fig. 10. Impact of estimation from noisy sequence: \( Z_{1:T} \) satisfies \( Z_{t+1} = \rho Z_t+U_t \) with \( U_t\sim \mathcal {N}(0,10) \) , \( Z_0 = 0 \) and weak ( \( \rho =0.1 \) ) or strong ( \( \rho =0.8 \) ) autocorrelation.

As mentioned in Section 5, the proposed method can attain a bounded total privacy loss (\( \lim _{T\rightarrow \infty }\epsilon _T\lt \infty \)) by taking weights \( w_t^2 = ar^{t-1}, r\in (0,1) \) as a decreasing geometric sequence. In Figure 11, we examine the performance of our algorithm on traffic count dataset when weights \( w_t^2 = r^{t-1}, r\in (0,1) \). Specifically, each figure shows the relative error of the proposed method (red curves) and baseline (black curves) as functions of \( r\in (0.97,1) \) under a certain privacy loss \( \epsilon _T \). Because r doesn’t impact the baseline method, the error of baseline remains the same (the oscillation in the plot comes from the randomness). The results show that for any privacy loss \( \epsilon _T \), always there exists a lower bound \( r_o\in (0,1) \) such that \( \forall r\ge r_o \), the proposed algorithm with weights \( w_t^2 = r^{t-1} \) outperforms the baseline method. Moreover, as privacy guarantee gets weaker (i.e., \( \epsilon _T \) increases), the lower bound \( r_o \) that leads to the improvement also increases. This is consistent with Theorem 6.2.

Fig. 11.

Fig. 11. Performance of the proposed method under a geometric sequence of weights \( w_t^2 = r^{t-1}, r\in (0.97,1) \) . It shows that for any privacy loss \( \epsilon _T \) , always there exists a lower bound \( r_o\in (0,1) \) such that \( \forall r\ge r_o \) , the proposed algorithm with weights \( w_t^2 = r^{t-1} \) outperforms the baseline method.

8.3 Queries Outside of Count/stationary Queries

As discussed in Section 7, the proposed method is not limited to AR(1) processes but is applicable to any (weakly) stationary random process. In fact, the empirical results further show that the proposed method works well even for non-stationary sequences. In particular, real-world datasets we considered such as ride sharing counts (daily), NY traffic volume counts, and FTP drive cycle are non-stationary, i.e., the mean/variance of data changes over time, the correlation between values at any two time steps depend not only on their time difference, but also on the particular time step. Next, we further demonstrate this on synthetic non-stationary data sequences (Figure 12).

Fig. 12.

Fig. 12. Sample paths of synthetic non-stationary data sequence: \( Z_{1:T} \) satisfies \( Z_{t+1} = \rho Z_t+U_t+\eta \cdot t \) with \( U_t\sim \mathcal {N}(0,10) \) , \( Z_0 = 0 \) and \( \rho =0.8 \) . \( \eta \in [0,0.05] \) controls the degree of non-stationarity.

Specifically, the synthetic data is generated based on the following: (4) \( \begin{eqnarray} Z_t = \rho Z_{t-1} + U_t + \eta \cdot t, ~\forall t\ge 1~. \end{eqnarray} \) We adopt Gaussian distributed \( U_t\sim \mathcal {N}(0,\sigma _u^2) = \mathcal {N}(0,10) \) and \( \rho =0.8 \). When \( \eta =0 \), Equation (4) reduces to Gaussian AR(1) process in (1). Term \( \eta \cdot t \) in Equation (4) is added for interrupting stationarity: as \( \eta \ge 0 \) increases, the interruption is more severe. Figure 12 illustrates the sample paths of \( Z_{1:T} \) under different values of \( \eta \), where each \( Z_t \) is Gaussian distributed with mean \( \mu \cdot t \) and variance \( \sigma _z^2 = \frac{\sigma _u^2}{1-\rho ^2}\approx 27.8 \). Note that for these sequences, the sensitivity \( \Delta _t \) is same over time and we set \( \Delta _t = 1 \) in the experiment. To examine the impact of non-stationarity, we adjust the accuracy metric and define relative error as \( \begin{align*} RE(z_{1:T},x_{1:T}) = \frac{1}{T}\frac{||z_{1:T}-x_{1:T}||_2}{3\sigma _z}~. \end{align*} \) Figures 13 compares the accuracy of our method with baseline-Normal under the same privacy guarantee. We observe that our method always outperforms the baseline significantly, and the impact of non-stationarity is minor; it is more significant when the privacy guarantee is more restrictive: when \( \epsilon _T=10^{-3}\times T \), the performance of our method deteriorates as the data becomes more non-stationary.

Fig. 13.

Fig. 13. Impact of non-stationarity: \( Z_{1:T} \) satisfies \( Z_{t+1} = \rho Z_t+U_t+\eta \cdot t \) with \( U_t\sim \mathcal {N}(0,10) \) , \( Z_0 = 0 \) and \( \rho =0.8 \) . \( \eta \in [0,0.05] \) controls the degree of non-stationarity. In each plot, the accuracy of the proposed method and baseline-Normal are compared under the same privacy guarantee. Results show that the impact of non-stationarity is minor; it is more significant when the privacy guarantee is more restrictive: when \( \epsilon _T=10^{-3}\times T \) , the performance of our method deteriorates as the data becomes more non-stationary.

The proposed method is not limited to count queries and is more broadly applicable. For example, this method can be used in intelligent transportation systems to enable private vehicle-to-vehicle communication. In our studies [Huang et al. 2020; Zhang et al. 2019], a predictive cruise controller is designed for a follower vehicle using a private speed profile transmitted from its leader vehicle. Specifically, instead of broadcasting the real speed profile (FTP drive cycle), the leader vehicle generates a differentially private speed profile using the proposed method. A follower vehicle then designs an optimal speed planner based on the received information. Within this application context, query \( \mathcal {Q}(v) \) represents the vehicle’s speed information, and sensitivity \( \Delta \) is the range of the vehicle’s speed. Figure 14 shows the private speed profiles generated using the proposed method and baseline-Normal. A private optimal speed planner is designed in Huang et al. [2020]; Zhang et al. [2019] using these private profiles. The results show that the controller performance deteriorates significantly under the baseline method. In contrast, the controller designed with the proposed method can attain an accuracy that is sufficient for predictive control purposes. We refer an interested reader to Huang et al. [2020], Zhang et al. [2019] for more details and the performance of the private controller.

Fig. 14.

Fig. 14. Drive cycles under different levels of privacy: the privacy guarantee in the upper plot is stronger than that of the lower plot.

Skip 9CONCLUSION Section

9 CONCLUSION

This paper presented a method for releasing a differentially private data sequence in real time. It estimates the unreleased data from those already released based on their correlation, which is learned on the fly during the release process. This estimate along with the actual data is then released as a convex combination with added perturbation. This is shown to achieve higher accuracy with lower privacy loss compared to various existing approaches.

APPENDICES

A RESULTS OF BINOMIAL AR(1) PROCESS

Skip A.1Definition Section

A.1 Definition

Theorem 4 (Binomial AR(1) Process).

Let \( \pi \in (0,1) \) and \( \rho \in [\max (-\frac{\pi }{1-\pi },-\frac{1-\pi }{\pi }),1] \). Define \( \beta = \pi (1-\rho) \), \( \alpha = \beta + \rho \), and fix \( n \in \mathbb {N} \). Then \( Z_{1:T} \) is a binomial AR(1) process [McKenzie 1985] if: (5) \( \begin{eqnarray} Z_t = \alpha \circ Z_{t-1} + \beta \circ (n-Z_{t-1}), ~ t\ge 1 \end{eqnarray} \) where \( Z_0 \sim Binomial(n,\pi) \) and “\( \circ \)” is called the thinning operator defined as \( a \circ Z_{t-1} = \sum _{i=1}^{Z_{t-1}} Y_{i,t-1} \), where \( Y_{i,t-1}, i=1, \ldots , Z_{t-1} \) are i.i.d Bernoulli random variables with \( \text{Pr}(Y_{i,t-1}=1)=a \), and all thinnings are independent of each other. Binomial AR(1) is also a stationary Markov process with the following properties: (1) \( Z_t \sim Binomial(n,\pi) \); (2) its autocorrelation is \( \text{Corr}(Z_tZ_{t-\tau }) = \text{Corr}(\tau) = \rho ^{|\tau |} \).

Binomial AR(1) is typically used for modeling integer-valued counts sequences. Consider n independent entities, each of which can be either in state “1” or state “0”. Then \( Z_t \) can be interpreted as the number of entities in state “1” at time t. Equation (5) implies that this “1”-entity count \( (Z_t) \) can be given by the number of “1”-entities in the previous time instant that didn’t change state \( (\alpha \circ Z_{t-1}) \) plus the number of “0”-entities in the previous time instant that changed to state “1” \( (\beta \circ (n-Z_{t-1})) \); here \( \alpha \), \( \beta \) can be interpreted as the respective transition probabilities. Binomial AR(1) has been used to model many real-world scenarios such as counts of computer log-ins and log-outs [Weiß 2009], daily counts of occupied rooms in a hotel, etc.

Skip A.2MMSE Estimates Section

A.2 MMSE Estimates

Proposition A.1.

Consider a Binomial AR(1) process \( Z_{1:T} \) defined by (5), the MMSE estimate of \( Z_{t+1} \) given \( Z_t=z_t \) is \( \hat{Z}_{t+1}(z_t) = \rho z_t + n\pi (1-\rho) \). If we use a perturbed \( X_i = Z_i + N_i, ~ i\in \lbrace 1,\ldots ,t\rbrace \) with \( \text{Var}(N_t) = \frac{m}{2} \) to estimate \( Z_{t+1} \), then the LMMSE estimate of \( Z_{t+1} \) given \( X_i=x_i \) is \( \begin{eqnarray*} \hat{Z}_{t+1}(x_i)&=& n\pi \left(1-\rho ^{t+1-i}\frac{n\pi (1-\pi)}{n\pi (1-\pi) + \frac{m}{2}}\right)+ \rho ^{t+1-i}\frac{n\pi (1-\pi)}{n\pi (1-\pi) + \frac{m}{2}} x_i \end{eqnarray*} \)

Skip A.3Binomial Mechanism Section

A.3 Binomial Mechanism

Theorem 5 (Binomial Noise).

We call random variable N the binomial noise if it is zero mean and follows the shifted binomial distribution: \( \begin{align*} N+m\sim {\it Binomial} \left(2m,\frac{1}{2}\right), \end{align*} \) whose probability mass function (PMF) is \( \begin{align*} \text{Pr}(N=k) = \binom{2m}{k+m}\frac{1}{2^{2m}}, ~ k \in \lbrace -m,\ldots ,m-1,m\rbrace , \end{align*} \) with a variance \( \frac{m}{2} \).

Lemma A.2 (Binomial Mechanism).

Consider a query \( \mathcal {Q}:\mathcal {D}\rightarrow \mathbb {Z} \) that takes data \( d\in \mathcal {D} \) as input and outputs an integer. The Binomial mechanism \( \mathcal {B}(d) = \mathcal {Q}(d)+N \) adds binomial noise N with variance \( \frac{m}{2} \) to the output. If \( 1\le \Delta \mathcal {Q}+\frac{2m+1}{\exp (\frac{\epsilon }{\Delta \mathcal {Q}})+1}\le m+1 \) for \( \epsilon \gt 0 \), then the following holds:

(i) \( \forall \) \( \epsilon \gt 0 \), it satisfies \( (\epsilon ,\delta) \)-differential privacy with: \( \begin{align*} \delta = \exp \left(-\frac{1}{m}\left(m-\Delta \mathcal {Q}+1-\frac{2m+1}{\exp (\frac{\epsilon }{\Delta \mathcal {Q}})+1}\right)^2\right). \end{align*} \)

(ii) \( \forall \) \( \delta \in (0,1) \), it satisfies \( (\epsilon ,\delta) \)-differential privacy with: \( \begin{align*} \epsilon = \Delta \mathcal {Q}\log \left(\frac{2m+1}{m-\Delta \mathcal {Q}+1-\sqrt {m\log \frac{1}{\delta }}}-1\right). \end{align*} \)

Note that this is a generalization (for arbitrary sensitivity \( \Delta \mathcal {Q} \)) to the version (for the case \( \Delta \mathcal {Q} = 1 \)) first proposed in Dwork et al. [2006]. This is an approximation of the Gaussian mechanism; it has a much looser bound compared to the latter and more noise is needed to ensure a same level of privacy, which is consistent with the conclusion in Dwork et al. [2006]. However, the Gaussian mechanism only works when \( \epsilon \lt 1 \), while our Binomial mechanism does not have this restriction and is more suitable for a discrete setting.

Skip A.4Privacy Analysis using Binomial Noise Section

A.4 Privacy Analysis using Binomial Noise

Theorem A.3.

Let \( Z_t = \mathcal {Q}(D_t) \) and \( \Delta \) be the sensitivity of \( \mathcal {Q} \) \( \forall t \), consider Algorithm 2 using Binomial noise with \( \text{Var}(N_t) = \frac{m}{2} \), \( \forall t \) that takes sequence \( z_{1:T} \) as input and outputs \( x_{1:T} \), \( \forall \) \( \tilde{\delta }\in [0,1] \), if \( 1\le w_t\Delta +\frac{2m+1}{\exp (\frac{\epsilon }{w_t\Delta })+1}\le m+1, ~\forall t \), then the algorithm is \( (\tilde{\epsilon }_{\tilde{\delta }},1-(1-\tilde{\delta })\prod _{t=1}^T(1-\delta _t)) \)-differentially private for: \( \begin{eqnarray*} &\tilde{\epsilon }_{\tilde{\delta }}=\nonumber\\ & \min \left\lbrace \sum _{t=1}^T\epsilon _t,\sum _{t=1}^T\frac{(e^{\epsilon _t}-1)\epsilon _t}{e^{\epsilon _t}+1} +\sqrt {\sum _{t=1}^T2\epsilon ^2_t\log \left(e+\frac{\sqrt {\sum _{t=1}^T\epsilon _t^2}}{\tilde{\delta }}\right)}, \sum _{t=1}^T\frac{(e^{\epsilon _t}-1)\epsilon _t}{e^{\epsilon _t}+1}+\sqrt {\sum _{t=1}^T2\epsilon ^2_t\log \left(\frac{1}{\tilde{\delta }}\right)}\right\rbrace \nonumber \end{eqnarray*} \) with any \( \epsilon _t\gt 0 \) and corresponding \( \begin{align*} \delta _t = \exp \left(-\frac{1}{m}(m-w_t\Delta +1-\frac{2m+1}{\exp (\frac{\epsilon _t}{w_t\Delta })+1})^2\right), \end{align*} \) or with any \( \delta _t\in (0,1) \) and corresponding \( \begin{align*} \epsilon _t = w_t\Delta \log \left(\frac{2m+1}{m-w_t\Delta +1-\sqrt {m\log \frac{1}{\delta _t}}}-1\right). \end{align*} \)

Skip A.5Proofs for Results of Binomial AR(1) Process Section

A.5 Proofs for Results of Binomial AR(1) Process

A.5.1 Proof of Proposition A.2.

The MMSE estimate of \( Z_{t+1} \) given \( Z_t = z_t \) is \( \mathbb {E}(Z_{t+1}|Z_t = z_t) \). Since the thinning is performed independently, given \( Z_t = z_t \), the probability generating function satisfies the following: \( \begin{align*} G(s) = &\,\mathbb {E}_{Z_{t+1}|Z_t = z_t}(s^{Z_{t+1}}|Z_t=z_t) = \mathbb {E}(s^{\alpha \circ Z_t}|Z_t=z_t)\mathbb {E}(s^{\beta \circ (n- Z_t)}|Z_t=z_t)\\ =&\, (1-\beta +\beta s)^n \left(\frac{1-\alpha +\alpha s}{1-\beta + \beta s}\right)^{z_t}\\ G^{\prime }(s) =&\, n\beta \left(1-\beta +\beta s\right)^{n-1}\left(\frac{1-\alpha +\alpha s}{1-\beta + \beta s}\right)^{z_t}\\ & +\, (1-\beta +\beta s)^n z_t \left(\frac{1-\alpha +\alpha s}{1-\beta + \beta s}\right)^{z_t-1}\frac{\alpha }{\beta }\frac{\frac{1}{\beta }-\frac{1}{\alpha }}{\left(\frac{1}{\beta }-1+s\right)^2}\\ G^{\prime \prime }(s) = &\,n\beta ^2(n-1)(1-\beta +\beta s)^{n-2}\left(\frac{1-\alpha +\alpha s}{1-\beta + \beta s}\right)^{z_t} \\ &+\, 2n\beta (1-\beta +\beta s)^{n-1}z_t\left(\frac{1-\alpha +\alpha s}{1-\beta + \beta s}\right)^{z_t-1}\frac{\alpha }{\beta }\frac{\frac{1}{\beta }-\frac{1}{\alpha }}{(\frac{1}{\beta }-1+s)^2}+(1-\beta +\beta s)^nz_t\\ &\times \, \left((z_1-1)\left(\frac{1-\alpha +\alpha s}{1-\beta + \beta s}\right)^{z_t-2}\left(\frac{\alpha }{\beta }\frac{\frac{1}{\beta }-\frac{1}{\alpha }}{(\frac{1}{\beta }-1+s)^2}\right)^2 + \left(\frac{1-\alpha +\alpha s}{1-\beta + \beta s}\right)^{z_t-1}\frac{\alpha }{\beta }2\frac{\frac{1}{\alpha }-\frac{1}{\beta }}{(\frac{1}{\beta }-1+s)^3}\right) \end{align*} \)

Since \( \beta =\pi (1-\rho) \), \( \alpha =\beta +\rho \), \( \mathbb {E}(Z_{t+1}|Z_t=z_t) = \lim _{s\rightarrow 1}G^{\prime }(s) \) and \( \text{Var}(Z_{t+1}|Z_t=z_t) = \lim _{s\rightarrow 1} G^{\prime \prime }(s)+G^{\prime }(s)-(G^{\prime }(s))^2 \) gives: \( \begin{eqnarray*} \mathbb {E}(Z_{t+1}|Z_t=z_t)&=& \rho z_t + n\pi (1-\rho) ~;\\ \text{Var}(Z_{t+1}|Z_t=z_t) &=& \rho (1-\rho)(1-2\pi)z_t + n\beta (1-\beta). \end{eqnarray*} \) The corresponding MSE is: \( \begin{eqnarray*} \mathbb {E}((Z_{t+1}-\mathbb {E}(Z_{t+1}|Z_t = z_t))^2|Z_t = z_t) = \text{Var}(Z_{t+1}|Z_t = z_t) \end{eqnarray*} \)

A.5.2 Proof of Lemma A.4.

Consider any \( d, \hat{d}\in \mathcal {D} \), and with them the binomial mechanism outputs the same results b. Let \( \bar{b}=b-\mathcal {Q}(d) \) \( \begin{eqnarray*} \frac{\text{Pr}(b= \mathcal {Q}(d)+noise)}{\text{Pr}(b=\mathcal {Q}(\hat{d})+\hat{noise})} = \frac{\mathscr{F}_N(b-\mathcal {Q}(d))}{\mathscr{F}_N(b-\mathcal {Q}(\hat{d}))}\nonumber =\frac{\binom{2m}{m+b-\mathcal {Q}(d)}}{\binom{2m}{m+b-\mathcal {Q}(\hat{d})}}= \frac{\binom{2m}{m+\bar{b}}}{\binom{2m}{m+\bar{b}+\Delta \mathcal {Q}}} =\prod _{i=1}^{\Delta \mathcal {Q}}\frac{m+\bar{b}+i}{m-\bar{b}+1-i} \end{eqnarray*} \) A sufficient condition for (6) being bounded by \( \exp (\epsilon) \) is: (6) \( \begin{eqnarray} \forall i\in \lbrace 1,2,\ldots ,\Delta \mathcal {Q}\rbrace , ~~\frac{m+\bar{b}+i}{m-\bar{b}+1-i} \le \exp \left(\frac{\epsilon }{\Delta \mathcal {Q}}\right) ~~\stackrel{\tilde{b}=\bar{b}+i}{\Longleftrightarrow }~~\frac{m+\tilde{b}}{m-\tilde{b}+1} \le \exp \left(\frac{\epsilon }{\Delta \mathcal {Q}}\right) \end{eqnarray} \) from (6), we have: \( \begin{eqnarray*} \bar{b}\le \underset{i\in [\Delta \mathcal {Q}]}{\min }~ m+1-\frac{2m+1}{\exp (\frac{\epsilon }{\Delta \mathcal {Q}})+1}-i = m+1-\frac{2m+1}{\exp (\frac{\epsilon }{\Delta \mathcal {Q}})+1}-\Delta \mathcal {Q}\nonumber \end{eqnarray*} \) Let \( \bar{B} \) be the random variable of shifted \( Binomial(2m,\frac{1}{2}) \) with zero mean and realization \( \bar{b} \). According to Chernoff bound, \( \forall t \in [0,\sqrt {2m}] \), there is \( \text{Pr}(\bar{B}\ge t\frac{\sqrt {2m}}{2})\le e^{-t^2/2} \).

Then if \( 1\le \Delta \mathcal {Q}+\frac{2m+1}{\exp (\frac{\epsilon }{\Delta \mathcal {Q}})+1} \le m+1 \), there is: \( \begin{eqnarray*} \text{Pr}\left(\bar{B}\ge m+1-\frac{2m+1}{\exp (\frac{\epsilon }{\Delta \mathcal {Q}})+1}-\Delta \mathcal {Q} \right)\le \exp \left(-\frac{1}{m}(m-\Delta \mathcal {Q}+1-\frac{2m+1}{\exp (\frac{\epsilon }{\Delta \mathcal {Q}})+1})^2\right) = \delta \end{eqnarray*} \) Similarly, given \( \delta \in [0,1] \), the corresponding \( \epsilon \) is: \( \begin{eqnarray*} \epsilon = \Delta \mathcal {Q}\log \left(\frac{2m+1}{m-\Delta \mathcal {Q}+1-\sqrt {m\log \frac{1}{\delta }}}-1\right) \end{eqnarray*} \)

A.5.3 Proof of Theorem A.5.

The data of each individual here spans over T time steps, the total privacy loss is the accumulation of privacy loss from T time steps: \( \begin{eqnarray*} \frac{\mathscr{F}_{X_{1:T}|Z_{1:T}}(x_{1:T}|z_{1:T})}{\mathscr{F}_{X_{1:T}|Z_{1:T}}(x_{1:T}|\hat{z}_{1:T})} = \frac{\mathscr{F}_{X_{1}|Z_{1}}(x_{1}|z_{1})}{\mathscr{F}_{X_{1}|Z_{1}}(x_{1}|\hat{z}_{1})} \cdot \prod _{t=2}^T \frac{\mathscr{F}_{X_{t}|Z_t,X_{1:t-1}}(x_{t}|z_{t},x_{1:t-1})}{\mathscr{F}_{X_{t}|Z_{t},X_{1:t-1}}(x_{t}|\hat{z}_{t},x_{1:t-1})} \end{eqnarray*} \) If \( x_t \) is released under \( (\epsilon _t,\delta _t) \)-differential privacy at time t, then the total privacy loss can be calculated using the advanced composition theorem below:

Theorem A.4 (Advanced Composition Theorem for Differential Privacy [Kairouz et al. 2017]).

For any \( \epsilon _k\gt 0 \), \( \delta _k\in [0,1] \) for \( k\in \lbrace 1,2,\ldots ,T\rbrace \), and \( \tilde{\delta }\in [0,1] \), the class of \( (\epsilon _k,\delta _k) \)-differentially private mechanisms satisfy \( (\tilde{\epsilon }_{\tilde{\delta }},1-(1-\tilde{\delta })\prod _{k=1}^T(1-\delta _k)) \)-differential privacy under T-fold adaptive composition, for \( \begin{align*} \tilde{\epsilon }_{\tilde{\delta }}=&\, \min \left\lbrace \sum _{k=1}^T\frac{(e^{\epsilon _k}-1)\epsilon _k}{e^{\epsilon _k}+1}+\sqrt {\sum _{k=1}^T2\epsilon ^2_k\log \left(e+\frac{\sqrt {\sum _{k=1}^T\epsilon _k^2}}{\tilde{\delta }}\right)},\right. \\ &\times \, \left. \sum _{k=1}^T\epsilon _k,\sum _{k=1}^T\frac{(e^{\epsilon _k}-1)\epsilon _k}{e^{\epsilon _k}+1}+\sqrt {\sum _{k=1}^T2\epsilon ^2_k\log \left(\frac{1}{\tilde{\delta }}\right)}\right\rbrace \end{align*} \)

First calculate the \( (\epsilon _t,\delta _t) \) at each stage by Lemma A.4. Since \( N_t+m \sim Binomial(2m,\frac{1}{2}) \), for \( t\le 2 \), \( X_t = Z_t + N_t \) so that the sensitivity \( \Delta \mathcal {Q}_t = \Delta \); for \( t\gt 2 \), \( X_t = [(1-w_{t})(\hat{\mu }_{t-1}(1-r_t) + r_tX_{t-1})+w_{t}Z_{t}+N_t] \) and the sensitivity is \( \Delta \mathcal {Q}_t = w_t\Delta \). Let \( w_t = 1 \) for \( t\le 2 \). Then \( \forall \epsilon _t\gt 0 \), \( \begin{align*} \delta _t = \exp \left(-\frac{1}{m}\left(m-w_t\Delta +1-\frac{2m+1}{\exp (\frac{\epsilon _t}{w_t\Delta })+1}\right)^2\right) \end{align*} \) or \( \forall \delta _t \in (0,1) \), \( \begin{align*} \epsilon _t = w_t\Delta \log \left(\frac{2m+1}{m-w_t\Delta +1-\sqrt {m\log \frac{1}{\delta _t}}}-1\right) \end{align*} \)

Apply Theorem A.6 directly, Theorem A.5 is proved.

B PROOFS

Skip B.1Proof of Proposition 2.4 Section

B.1 Proof of Proposition 2.4

Finding the MMSE estimate of \( Z_{t+1} \) given \( Z_t =z_t \) is equivalent to finding the mapping \( \begin{eqnarray*} f^*= \underset{f}{\text{argmin}}~\mathbb {E}((Z_{t+1}-f(Z_t))^2|Z_t = z_t) = \underset{f}{\text{argmin}}~ \int _{-\infty }^{\infty }p(z_{t+1}|z_t)(z_{t+1}-f(z_t))^2dz_{t+1} \end{eqnarray*} \) Differentiating with respect to f and equating the result to zero gives: \( \begin{eqnarray*} \int _{-\infty }^{\infty }p(z_{t+1}|z_t)f^*(z_t)dz_{t+1} = f^*(z_t) = \int _{-\infty }^{\infty }p(z_{t+1}|z_t)z_{t+1}dz_{t+1} = \mathbb {E}(Z_{t+1}|Z_t = z_t) \end{eqnarray*} \) Therefore, the MMSE estimate of \( Z_{t+1} \) given \( Z_t = z_t \) is \( \mathbb {E}(Z_{t+1}|Z_t = z_t) \).

Since \( (Z_t,Z_{t+1}) \) is jointly Gaussian: \( \begin{eqnarray*} \begin{pmatrix} Z_t\\ Z_{t+1} \end{pmatrix} \sim \mathcal {N}\left(\begin{bmatrix} \mu \\ \mu \end{bmatrix},\begin{bmatrix} \sigma _z^2 & \rho \sigma _z^2\\ \rho \sigma _z^2 & \sigma _z^2 \end{bmatrix}\right),\nonumber \end{eqnarray*} \) it implies that \( Z_{t+1}|Z_t \sim \mathcal {N}(\mu (1-\rho)+\rho Z_{t},\sigma _z^2(1-\rho ^2)) \), combining with the above result, the MMSE estimate of \( Z_{t+1} \) given \( Z_t = z_t \) is \( \mathbb {E}(Z_{t+1}|Z_t = z_t) = \mu (1-\rho)+\rho z_{t} \). The corresponding MSE is: \( \begin{eqnarray*} \mathbb {E}((Z_{t+1}-\mathbb {E}(Z_{t+1}|Z_t = z_t))^2|Z_t = z_t) = \text{Var}(Z_{t+1}|Z_t = z_t) = \sigma _z^2(1-\rho ^2) \end{eqnarray*} \)

Since \( Z_i\sim \mathcal {N}(\mu ,\sigma _z^2) \), \( N_i\sim \mathcal {N}(0,\sigma ^2_n) \), there is \( X_i = Z_i + N_i \sim \mathcal {N}(\mu ,\sigma _z^2+\sigma ^2_n) \) and \( \text{Corr}(X_iZ_{t+1}) =\rho ^{t+1-i}\frac{\sigma _z}{\sqrt {\sigma _z^2+\sigma ^2_n}} \).

\( (X_i,Z_{t+1}) \) is jointly Gaussian: \( \begin{eqnarray*} \begin{pmatrix} X_i\\ Z_{t+1} \end{pmatrix} \sim \mathcal {N}\left(\begin{bmatrix} \mu \\ \mu \end{bmatrix},\begin{bmatrix} \sigma _z^2+\sigma ^2_n & \rho ^{t+1-i}\sigma _z^2~\\ \rho ^{t+1-i}\sigma _z^2 & \sigma _z^2 \end{bmatrix}\right),\nonumber \end{eqnarray*} \) it implies the MMSE estimate of \( Z_{t+1} \) given \( X_i = x_i \): \( \begin{eqnarray*} \mathbb {E}(Z_{t+1}|X_i=x_i)=\mu \left(1-\rho ^{t+1-i}\frac{\sigma _z^2}{\sigma _z^2+\sigma ^2_n}\right)+ \rho ^{t+1-i}\frac{\sigma _z^2}{\sigma _z^2+\sigma ^2_n}x_i \end{eqnarray*} \)

The corresponding MSE is: \( \begin{align*} \sigma _z^2\left(1- (\rho ^{t+1-i})^2\frac{\sigma _z^2}{\sigma _z^2+\sigma ^2_n}\right) \end{align*} \)

Skip B.2Proof of Theorem 5.1 Section

B.2 Proof of Theorem 5.1

According to Abadi et al. [2016], for a mechanism \( \mathscr{M} \) outputs o, with inputs d and \( \hat{d} \), let a random variable \( c(o;\mathscr{M},d,\hat{d})=\log \frac{\text{Pr}(\mathscr{M}(d)=o)}{\text{Pr}(\mathscr{M}(\hat{d})=o)} \) denote the privacy loss at o, and \( \begin{align*} \alpha _{\mathscr{M}}(\lambda) = \underset{d,\hat{d}}{\max }\log \mathbb {E}_{o\sim \mathscr{M}(d)}\lbrace \exp (\lambda c(o;\mathscr{M},d,\hat{d}))\rbrace \end{align*} \) There is: \( \begin{align*} c(x_{1:T};\mathscr{M},z_{1:T},\hat{z}_{1:T}) =& \log \frac{\mathscr{F}_{X_{1:T}|Z_{1:T}}(x_{1:T}|z_{1:T})}{\mathscr{F}_{X_{1:T}|Z_{1:T}}(x_{1:T}|z_{1:T})}\\ [4pt] =& \sum _{t=2}^T\log \frac{\mathscr{F}_{X_t|Z_t,X_{1:t-1}}(x_t|z_t,x_{1:t-1})}{\mathscr{F}_{X_t|Z_t,X_{1:t-1}}(x_t|\hat{z}_t,x_{1:t-1})} + \log \frac{\mathscr{F}_{X_1|Z_1}(x_1|z_1)}{\mathscr{F}_{X_1|Z_1}(x_1|\hat{z}_1)}\\ [4pt] =&\sum _{t=1}^T c(x_t;\mathscr{M}_t,z_t,\hat{z}_t) \end{align*} \) and for any pair of sequences \( z_{1:T}, \hat{z}_{1:T} \), the following holds (7) \( \begin{align} \log \mathbb {E}_{X_{1:T}\sim \mathscr{M}(Z_{1:T})}\lbrace \exp (\lambda c(x_{1:T};\mathscr{M},z_{1:T},\hat{z}_{1:T}))\rbrace =& \,\log \mathbb {E}_{X_{1:T}\sim \mathscr{M}(Z_{1:T})}\left\lbrace \exp (\lambda \sum _{t=1}^T c(x_t;\mathscr{M}_t,z_t,\hat{z}_t)\right\rbrace \nonumber\\ [4pt] \le & \,\sum _{t=1}^T\log \mathbb {E}_{X_{t}\sim \mathscr{M}(Z_{t})}\lbrace \exp (\lambda c(x_t;\mathscr{M}_t,z_t,\hat{z}_t)\rbrace \end{align} \)

Therefore, \( \alpha _{\mathscr{M}}(\lambda) \le \sum _{t=1}^T \alpha _{\mathscr{M}_t}(\lambda) \) also holds.

Consider \( \alpha _{\mathscr{M}_t}(\lambda) \) first.

For \( t\le 2-T_0 \), \( X_t = Z_t + N_t \) with \( N_t \sim \mathcal {N}(0,\sigma _{n}^2) \) \( \begin{eqnarray*} c(x_t;\mathscr{M}_t,z_t,\hat{z}_t) &=& \log \frac{\mathscr{F}_{X_t|Z_t,X_{1:t-1}}(x_t|z_t,x_{1:t-1})}{\mathscr{F}_{X_t|Z_t,X_{1:t-1}}(x_t|\hat{z}_t,x_{1:t-1})} = \log \frac{\mathscr{F}_{N_t}(n_t)}{\mathscr{F}_{N_t}(\hat{n}_t)} \le \frac{1}{2\sigma _{n}^2}\Delta (2n_t+\Delta)~. \end{eqnarray*} \) \( \begin{align*} \alpha _{\mathscr{M}_t}(\lambda)\, &= \log \mathbb {E}_{N_t\sim \mathcal {N}(0,\sigma _{n}^2)}\left\lbrace \exp \left(\lambda \frac{1}{2\sigma _{n}^2}\Delta (2n_t+\Delta) \right)\right\rbrace \\ [8pt]&= \log \int _{-\infty }^{\infty }\frac{1}{\sqrt {2\pi }\sigma _{n}}\exp \left(-\frac{1}{2\sigma _{n}^2}(n_t-\lambda \Delta)^2\right)\cdot \exp \left(\frac{1}{2\sigma _{n}^2}(\lambda ^2+\lambda)\Delta ^2\right)dn_t \\ [8pt]&=\frac{\lambda (\lambda +1)\Delta ^2}{2\sigma _{n}^2}~. \end{align*} \)

For \( t\gt 2 \), \( \begin{align*} X_t = (1-w_{t})(\hat{\mu }_{t-1}(1-r_t) + r_tX_{t-1})+w_{t}Z_{t}+N_t \end{align*} \) with \( N_t\sim \mathcal {N}(0,\sigma ^2_{n}) \). \( \begin{eqnarray*} c(x_t;\mathscr{M}_t,z_t,\hat{z}_t) =\log \frac{\mathscr{F}_{X_t|Z_t,X_{1:t-1}}(x_t|z_t,x_{1:t-1})}{\mathscr{F}_{X_t|Z_t,X_{1:t-1}}(x_t|\hat{z}_t,x_{1:t-1})} = \log \frac{\mathscr{F}_{N_t}(n_t)}{\mathscr{F}_{N_t}(\hat{n}_t)} \le \frac{1}{2\sigma _{n}^2}w_t\Delta (2n_t+w_t\Delta)~. \end{eqnarray*} \) \( \begin{align*} \alpha _{\mathscr{M}_t}(\lambda)\, &= \log \mathbb {E}\left\lbrace \exp \left(\lambda \frac{1}{2\sigma _{n}^2}w_t\Delta (2n_t+w_t\Delta) \right)\right\rbrace \\ &= \log \int _{-\infty }^{\infty }\frac{1}{\sqrt {2\pi }\sigma _{n}} \exp \left(-\frac{1}{2\sigma _{n}^2}(n_t-\lambda w_t\Delta)^2\right)\cdot \exp \left(\frac{1}{2\sigma _{n}^2}(\lambda ^2+\lambda)w_t^2\Delta ^2\right)dn_t\\ &=\frac{\lambda (\lambda +1)w_t^2\Delta ^2}{2\sigma _{n}^2}~. \end{align*} \)

If let \( w_t = 1 \) for \( t\le 2 \), there is: \( \begin{align*} \alpha _{\mathscr{M}}(\lambda)\le \lambda (\lambda +1)\frac{\Delta ^2}{2\sigma _{n}^2}\sum _{t=1}^Tw_t^2 \end{align*} \) Use the tail bound [Theorem 2, Abadi et al. [2016]] , for any \( \epsilon _T\ge \frac{\Delta ^2}{2\sigma _{n}^2}\sum _{t=1}^Tw_t^2 \), the algorithm is \( (\epsilon _T,\delta _T) \)-differentially private for \( \begin{eqnarray*} \delta _T &=& \underset{\lambda :\lambda \ge 0}{\min }~h(\lambda)\nonumber =\underset{\lambda :\lambda \ge 0}{\min }\exp \left(\lambda (\lambda +1)\frac{\Delta ^2}{2\sigma _{n}^2}\sum _{t=1}^Tw_t^2-\lambda \epsilon _T\right) \end{eqnarray*} \)

To find \( \lambda ^* = \underset{\lambda :\lambda \ge 0}{\text{argmin}}~h(\lambda) \), take derivative of \( h(\lambda) \) and assign 0 gives the solution \( \bar{\lambda } = \frac{\epsilon _T}{\frac{\Delta ^2}{\sigma _{n}^2}\sum _{t=1}^Tw_t^2}-\frac{1}{2}\ge 0 \), and \( h^{\prime \prime }(\bar{\lambda })\gt 0 \), implies \( \lambda ^*=\bar{\lambda } \). Plug into (8) gives: (8) \( \begin{eqnarray} \delta _T = \exp \left(\left(\frac{\frac{\Delta ^2}{\sigma _{n}^2}\sum _{t=1}^Tw_t^2}{4}-\frac{\epsilon _T}{2}\right)\left(\frac{\epsilon _T}{\frac{\Delta ^2}{\sigma _{n}^2}\sum _{t=1}^Tw_t^2}-\frac{1}{2}\right)\right) \end{eqnarray} \)

Similarly, for any \( \delta _T\in [0,1] \), the algorithm is \( (\epsilon _T,\delta _T) \)-differentially private for (9) \( \begin{eqnarray} \epsilon _T &=& \underset{\lambda :\lambda \ge 0}{\min }~ h_1(\lambda) = \underset{\lambda :\lambda \ge 0}{\min }~(\lambda +1)\frac{\Delta ^2}{2\sigma _{n}^2}\sum _{t=1}^Tw_t^2 + \frac{1}{\lambda }\log \left(\frac{1}{\delta _T}\right)~~~~~ \end{eqnarray} \) with \( \lambda ^* = \underset{\lambda :\lambda \ge 0}{\text{argmin}}~h_1(\lambda) = \sqrt {\frac{\log \frac{1}{\delta _T}}{\frac{\Delta ^2}{2\sigma _{n}^2}\sum _{t=1}^Tw_t^2}} \). Plug into (9) gives: (10) \( \begin{eqnarray} \epsilon _T = 2\sqrt {\frac{\Delta ^2}{2\sigma _{n}^2}\sum _{t=1}^Tw_t^2\log \left(\frac{1}{\delta _T}\right)}+\frac{\Delta ^2}{2\sigma _{n}^2}\sum _{t=1}^Tw_t^2 \end{eqnarray} \)

Skip B.3Proof of Corollary 5.2 Section

B.3 Proof of Corollary 5.2

Let \( \phi = \frac{\Delta ^2\sum _{t=1}^Tw_t^2}{\sigma ^2_n} \), then according to Theorem 5.1, \( \begin{eqnarray*} \ln \delta _T = \left(\frac{\phi }{4}-\frac{\epsilon _T}{2}\right)\left(\frac{\epsilon _T}{\phi }-\frac{1}{2}\right) \end{eqnarray*} \) reorganized gives: \( \begin{eqnarray*} \phi ^2 + (8\ln \delta _T-4\epsilon _T)\phi + 4\epsilon _T^2 = 0 \end{eqnarray*} \) \( \begin{align*} \phi = 2\epsilon _T-4\ln \delta _T \pm 4\sqrt {(\ln \delta _T)^2 - \epsilon _T\ln \delta _T} \end{align*} \) Since \( \epsilon _T \ge \frac{\phi }{2} \) must hold, only one case is possible. \( \begin{align*} \phi = \frac{\Delta ^2\sum _{t=1}^Tw_t^2}{\sigma ^2_n} = 2\epsilon _T-4\ln \delta _T - 4\sqrt {(\ln \delta _T)^2 - \epsilon _T\ln \delta _T} \end{align*} \) Therefore, \( \begin{align*} \sigma ^2_n = \frac{\Delta ^2\sum _{t=1}^Tw_t^2}{2\epsilon _T+4\ln \frac{1}{\delta _T} - 4\sqrt {(\ln \frac{1}{\delta _T})^2 + \epsilon _T\ln \frac{1}{\delta _T}}} \end{align*} \)

Skip B.4Proof of Theorem 6.1 Section

B.4 Proof of Theorem 6.1

(11) \( \begin{align} \mathbb {E}_{X_{1:T}}(||x_{1:T}-z_{1:T} ||^2) \,&= \mathbb {E}_{X_{1:T}}\left(\sum _{t=1}^T(x_{t}-z_{t})^2\right)= \mathbb {E}_{X_{1:T-1}}\left\lbrace \sum _{t=1}^{T-1}(x_{t}-z_{t})^2 + \underbrace{\mathbb {E}_{X_T|X_{1:T-1}}[ (x_{T}-z_{T})^2 ] }_{{\bf term 1}} \right\rbrace \nonumber\\ \end{align} \)

Replacing \( x_T = (1-w_T)\hat{z}_T(x_{T-1}) + w_Tz_T+n_T \) into term 1 gives: \( \begin{eqnarray*} {\bf term 1} = \mathbb {E}_{X_T|X_{1:T-1}}\left[ ((1-w_t)(\hat{z}_T(x_{T-1})-z_T) +n_T)^2 \right] =(1-w_T)^2(\hat{z}_T(x_{T-1})-z_T)^2 +\sigma _n^2 \end{eqnarray*} \) Plug into Equation (11): \( \begin{align*} (11)\,& = \mathbb {E}_{X_{1:T-1}}\left\lbrace \sum _{t=1}^{T-1}(x_{t}-z_{t})^2 +(1-w_T)^2(\hat{z}_T(x_{T-1})-z_T)^2 +\sigma _n^2 \right\rbrace \\ &= \mathbb {E}_{X_{1:T-2}}\left\lbrace \sum _{t=1}^{T-2}(x_{t}-z_{t})^2 + \sigma _n^2+ {\bf term 2} \right\rbrace \end{align*} \) with \( \begin{align*} {\bf term 2} &=\mathbb {E}_{X_{T-1}|X_{1:T-2}}\lbrace (x_{T-1}-z_{T-1})^2 + (1-w_T)^2(\hat{z}_T(x_{T-1})-z_T)^2 \rbrace \\ &=(1-w_{T-1})^2(\hat{z}_{T-1}(x_{T-2})-z_{T-1})^2 +\sigma _n^2+ (1-w_T)^2\underbrace{\mathbb {E}_{X_{T-1}}\lbrace (\hat{z}_T(x_{T-1})-z_T)^2\rbrace }_{{\bf term 3}} \end{align*} \) Since \( \hat{z}_T(x_{T-1}) \) is the LMMSE estimator of \( Z_T \) given \( X_{T-1} \), term 3 is just the corresponding MSE. For a Gaussian AR(1) process \( Z_{1:T} \) with \( Z_t\sim \mathcal {N}(\mu ,\sigma _z^2) \) and \( \text{Corr}(Z_tZ_{t-1}) = \rho \). there is: \( \begin{eqnarray*} {\bf term 3} = \sigma _z^2\left(1- \rho ^2\frac{\sigma _z^2}{\sigma _z^2+\sigma ^2_n}\right) \end{eqnarray*} \) Therefore, \( \begin{align*} (11)=&\, \mathbb {E}_{X_{1:T-2}}\left\lbrace \sum _{t=1}^{T-2}(x_{t}-z_{t})^2 + (1-w_{T-1})^2(\hat{z}_{T-1}(x_{T-2})-z_{T-1})^2 \right\rbrace \\ & +\, (1-w_T)^2\sigma _z^2\left(1- \rho ^2\frac{\sigma _z^2}{\sigma _z^2+\sigma ^2_n}\right)+ 2\sigma _n^2 \\ = &\, \cdots \\ = &\, \mathbb {E}_{X_{1}}\lbrace (x_{1}-z_{1})^2 + (1-w_{2})^2(\hat{z}_{2}(x_{1})-z_{2})^2 \rbrace + \sigma _z^2\left(1- \rho ^2\frac{\sigma _z^2}{\sigma _z^2+\sigma ^2_n}\right)\sum _{t=3}^T(1-w_t)^2+ (T-1)\sigma _n^2\\ =& \,\sigma _z^2\left(1- \rho ^2\frac{\sigma _z^2}{\sigma _z^2+\sigma ^2_n}\right)\sum _{t=2}^T(1-w_t)^2+ T\sigma _n^2 \end{align*} \) Since \( w_1 = 1 \), \( \begin{eqnarray*} \mathbb {E}_{X_{1:T}}(||x_{1:T}-z_{1:T} ||^2)= \sigma _z^2\left(1- \rho ^2\frac{\sigma _z^2}{\sigma _z^2+\sigma ^2_n}\right)\sum _{t=1}^T(1-w_t)^2+ T\sigma _n^2 \end{eqnarray*} \)

Skip B.5Proof of Theorem 6.2 Section

B.5 Proof of Theorem 6.2

Since both satisfy \( (\epsilon _T,\delta _T) \)-differential privacy, according to Corollary 5.2, \( (\sigma _n^2)^A \), \( (\sigma _n^2)^B \) should satisfy: \( \begin{eqnarray*} \frac{T}{ (\sigma _n^2)^B} &=& \frac{\sum _{t=1}^{T}w_t^2}{ (\sigma _n^2)^A}= \frac{2\epsilon _T+4\ln \frac{1}{\delta _T} - 4\sqrt {(\ln \frac{1}{\delta _T})^2 + \epsilon _T\ln \frac{1}{\delta _T}}}{\Delta ^2} \nonumber \end{eqnarray*} \) By Theorem 6.1, \( \begin{eqnarray*} && \mathbb {E}_{X^A_{1:T}}(||x_{1:T}-z_{1:T} ||^2)= \sigma _z^2\left(1- \rho ^2\frac{\sigma _z^2}{\sigma _z^2+(\sigma ^2_n)^A}\right)\sum _{t=1}^T(1-w_t)^2+T(\sigma _n^2)^A \\ && \mathbb {E}_{X^B_{1:T}}(||x_{1:T}-z_{1:T} ||^2)= T(\sigma _n^2)^B =T (\sigma _n^2)^A\frac{T}{\sum _{t=1}^{T}w_t^2} \end{eqnarray*} \) If \( \mathbb {E}_{X^A_{1:T}}(||x_{1:T}-z_{1:T} ||^2) \lt \mathbb {E}_{X^B_{1:T}}(||x_{1:T}-z_{1:T} ||^2) \), then \( \exists t \) s.t. \( w_t\ne 1 \) and \( \begin{eqnarray*} \sigma _z^2\left(1- \rho ^2\frac{\sigma _z^2}{\sigma _z^2+(\sigma ^2_n)^A}\right)\sum _{t=1}^T(1-w_t)^2\lt T(\sigma _n^2)^A\left(\frac{T}{\sum _{t=1}^{T}w_t^2}-1\right) \end{eqnarray*} \)

Reorganized it implies: (12) \( \begin{eqnarray} \frac{(\sigma _n^2)^A/\sigma _z^2}{1-\frac{\rho ^2}{1+(\sigma _n^2)^A/\sigma _z^2}} \gt \frac{\sum _{t=1}^T(1-w_t)^2}{T(\frac{T}{\sum _{t=1}^{T}w_t^2}-1)} \end{eqnarray} \)

Therefore, if \( \exists \) \( \lbrace w_t\rbrace _{t=1}^T, w_t\in (0,1) \) and \( (\sigma _{n}^2)^A \) satisfy both (12) and (12), then \( x^A_{1:T} \) will be more accurate than \( x^B_{1:T} \).

Consider the case when \( w_t = w\in (0,1),\forall t \).

Then the right hand side of (12) is reduced to \( h_1(w) = \frac{(1-w)^2}{\frac{1}{w^2}-1} \), since \( \begin{eqnarray*} \lim _{w\rightarrow 1}h_1(w) &=& 0; ~~~~ \lim _{w\rightarrow 0}h_1(w) = 0 \\ h_1^{\prime }(w) &=&\frac{-2w(w^2+w-1)}{(1+w)^2}\\ \end{eqnarray*} \) \( \exists \) only one \( \bar{w} \) over \( (0,1) \) such that \( \bar{w}^2+\bar{w}-1 = 0 \). Therefore, \( h_1(w) \) is strictly increasing from 0 to \( h_1(\bar{w})\gt 0 \) over \( (0,\bar{w}) \) and strictly decreasing over from \( h_1(\bar{w})\gt 0 \) to 0 over \( (\bar{w},1) \).

Let \( \xi = (\sigma _n^2)^A/\sigma _z^2\ge 0 \), then the left hand side of (12) can be re-written as \( h_2(\xi) = \frac{\xi }{1-\frac{\rho ^2}{1+\xi }} \), we have: \( \begin{align*} h_2^{\prime }(\xi) = \frac{\xi ^2 +2\xi (1-\rho ^2)+(1-\rho ^2)}{(1+\xi -\rho ^2)^2} \end{align*} \) Since \( h_2(0) = 0 \) and \( h_2^{\prime }(\xi)\gt 0 \) over \( \xi \in [0,\infty) \), \( h_2(\xi) \) is strictly increasing from 0 to \( +\infty \) over \( \xi \in [0,\infty) \). For all pairs of \( (w,(\sigma _n^2)^A) \) satisfying (12), w and \( (\sigma _n^2)^A \) is bijective and we can write \( \xi = h_3(w) \) for some strictly increasing function \( h_3 \).

Since both \( h_2 \) and \( h_3 \) are strictly increasing functions, \( h_2(h_3(w)) \) is strictly increasing from 0 over \( w\in (0,1) \). Therefore, \( \exists w \in (0,1) \), such that \( h_2(h_3(w)) \gt h_1(w) \) and \( x^A_{1:T} \) released by our method is more accurate than \( x^B_{1:T} \).

Moreover, if \( w\gt \frac{1-(\sigma _n^2)^B/\sigma _z^2}{1+(\sigma _n^2)^B/\sigma _z^2} \), then re-organized it implies \( \begin{align*} w^2\frac{(\sigma _n^2)^B}{\sigma _z^2} \gt h_1(w). \end{align*} \) Since \( h_2(h_3(w)) = \frac{ w^2\frac{(\sigma _n^2)^B}{\sigma _z^2}}{1-\frac{\rho ^2}{1+ w^2\frac{(\sigma _n^2)^B}{\sigma _z^2}}}\gt w^2\frac{(\sigma _n^2)^B}{\sigma _z^2} \), it further implies \( h_2(h_3(w)) \gt h_1(w) \).

Therefore, if \( \begin{align*} w\gt \frac{1-(\sigma _n^2)^B/\sigma _z^2}{1+(\sigma _n^2)^B/\sigma _z^2}, \end{align*} \) then \( x^A_{1:T} \) will be more accurate than \( x^B_{1:T} \).

C ADDITIONAL EXPERIMENTS

Results on ride-sharing counts datasets are shown in Figures 1519. Specifically, Figures 1619 show the private sequences aggregated from 10 runs of experiments using different methods, and Figure 15 shows the comparison of various methods under a different privacy guarantee.

Fig. 15.

Fig. 15. Comparison of different methods.

Fig. 16.

Fig. 16. Private sequences using different methods under the same \( \epsilon _T = 2*10^{-2}T \) (left) and \( \epsilon _T = 1.3*10^{-2}T \) (right).

Fig. 17.

Fig. 17. Private sequences using different methods where parameters are selected such that error achieved by baseline-Normal is slightly larger than baseline-Laplace: \( \sigma _n^2\text{(Laplace)} = 5*10^3 \) (left) and \( \sigma _n^2\text{(Laplace)} = 1.18*10^4 \) (right).

Fig. 18.

Fig. 18. Private sequences using different methods under the same \( \epsilon _T = 2*10^{-3}T \) (left) and \( \epsilon _T = 10^{-4}T \) (right).

Fig. 19.

Fig. 19. Private sequences using different methods where parameters are selected such that error achieved by baseline-Normal is slightly larger than baseline-Laplace: \( \sigma _n^2\text{(Laplace)} = 5*10^5 \) (left) and \( \sigma _n^2\text{(Laplace)} = 2*10^8 \) (right).

Footnotes

  1. 1 User-level privacy is much stronger than event-level privacy because it guarantees the privacy of all events.

    Footnote
  2. 2 If we express D in matrix form, i.e., \( D\in \mathbb {R}^{N\times T} \) with \( D_{it} = d_t^i \), then D and \( \widehat{D} \) are different in at most one row, i.e., the hamming distance between two matrices is at most 1.

    Footnote
  3. 3 This estimate can be obtained with or without the knowledge of the statistics of the AR(1) process; in the absence of such knowledge one can employ a separate procedure to first estimate the statistics as detailed later in this section.

    Footnote
  4. 4 The extra term \( \frac{1}{t-1} \) is used for correcting the negative bias if there is prior knowledge of the positive autocorrelation \( \rho ^{true}\gt 0 \).

    Footnote
  5. 5 FAST samples \( k\lt T \) points and allocates privacy budget \( \epsilon _T \) to the sampled points. It adds Laplace noise \( Lap(0,\frac{ K\Delta }{\epsilon _T}) \) to each sampled point and outputs the corresponding a posterior estimate, while for non-sampled points it outputs prior estimates. A similar sampling procedure can be added to our proposed method where we set \( w_t=0 \) for non-sampled points.

    Footnote
  6. 6 Event-level privacy only guarantees the privacy at one time step; it requires sequence pairs \( z_{1:T} \), \( \widehat{z}_{1:T} \) defined in differential privacy (Definition 2.2) to be different in at most one data point. In contrast, our work considers user-level privacy where privacy at all times is ensured.

    Footnote

REFERENCES

  1. Abadi Martin, Chu Andy, Goodfellow Ian, McMahan H. Brendan, Mironov Ilya, Talwar Kunal, and Zhang Li. 2016. Deep learning with differential privacy. In Proceedings of the 2016 ACM SIGSAC Conference on Computer and Communications Security. 308318.Google ScholarGoogle ScholarDigital LibraryDigital Library
  2. Cao Yang and Yoshikawa Masatoshi. 2015. Differentially private real-time data release over infinite trajectory streams. In 2015 16th IEEE International Conference on Mobile Data Management, Vol. 2. IEEE, 6873.Google ScholarGoogle ScholarDigital LibraryDigital Library
  3. Cao Yang, Yoshikawa Masatoshi, Xiao Yonghui, and Xiong Li. 2019. Quantifying differential privacy in continuous data release under temporal correlations. IEEE Transactions on Knowledge and Data Engineering 31, 7 (2019), 12811295. Google ScholarGoogle ScholarCross RefCross Ref
  4. Chan T.-H. Hubert, Shi Elaine, and Song Dawn. 2011. Private and continual release of statistics. ACM Transactions on Information and System Security (TISSEC) 14, 3 (2011), 124.Google ScholarGoogle ScholarDigital LibraryDigital Library
  5. Chen Rui, Acs Gergely, and Castelluccia Claude. 2012. Differentially private sequential data publication via variable-length n-grams. In Proceedings of the 2012 ACM Conference on Computer and Communications Security. 638649.Google ScholarGoogle ScholarDigital LibraryDigital Library
  6. Chen Rui, Fung Benjamin, and Desai Bipin C.. 2011. Differentially private trajectory data publication. arXiv preprint arXiv:1112.2020 (2011).Google ScholarGoogle Scholar
  7. DOT. 2011. Traffic volume counts, NYC OpenData. https://data.cityofnewyork.us/Transportation/Traffic-Volume-Counts-2011-2012-/wng2-85mv.Google ScholarGoogle Scholar
  8. Dwork Cynthia. 2006. Differential privacy. In 33rd International Colloquium on Automata, Languages and Programming, part II (ICALP’06). 4052, 112. https://www.microsoft.com/en-us/research/publication/differential-privacy/.Google ScholarGoogle Scholar
  9. Dwork Cynthia, Kenthapadi Krishnaram, McSherry Frank, Mironov Ilya, and Naor Moni. 2006. Our data, ourselves: Privacy via distributed noise generation. In Annual International Conference on the Theory and Applications of Cryptographic Techniques. Springer, 486503.Google ScholarGoogle ScholarDigital LibraryDigital Library
  10. Dwork Cynthia, Naor Moni, Pitassi Toniann, and Rothblum Guy N.. 2010. Differential privacy under continual observation. In Proceedings of the Forty-second ACM Symposium on Theory of Computing. 715724.Google ScholarGoogle ScholarDigital LibraryDigital Library
  11. Dwork Cynthia, Roth Aaron, et al. 2014. The algorithmic foundations of differential privacy. Foundations and Trends® in Theoretical Computer Science 9, 3–4 (2014), 211407.Google ScholarGoogle ScholarDigital LibraryDigital Library
  12. EPA. 2008. EPA Federal Test Procedure (FTP). https://www.epa.gov/emission-standards-reference-guide/epa-federal-test-procedure-ftp.Google ScholarGoogle Scholar
  13. Fan Liyue and Xiong Li. 2014. An adaptive approach to real-time aggregate monitoring with differential privacy. IEEE Transactions on Knowledge and Data Engineering 26, 9 (2014), 20942106.Google ScholarGoogle ScholarCross RefCross Ref
  14. Fanaee-T Hadi and Gama Joao. 2013. Event labeling combining ensemble detectors and background knowledge. Progress in Artificial Intelligence (2013), 115. Google ScholarGoogle ScholarCross RefCross Ref
  15. Fioretto Ferdinando and Van Hentenryck Pascal. 2019. OptStream: Releasing time series privately. Journal of Artificial Intelligence Research 65 (2019), 423456.Google ScholarGoogle ScholarDigital LibraryDigital Library
  16. Forney G. D.. 1973. The Viterbi algorithm. Proc. IEEE 61, 3 (1973), 268278. Google ScholarGoogle ScholarCross RefCross Ref
  17. Friedman Arik and Schuster Assaf. 2010. Data mining with differential privacy. In Proceedings of the 16th ACM SIGKDD International Conference on Knowledge Discovery and Data Mining. 493502.Google ScholarGoogle ScholarDigital LibraryDigital Library
  18. Ghane Soheila, Kulik Lars, and Ramamohanarao Kotagiri. 2019. TGM: A generative mechanism for publishing trajectories with differential privacy. IEEE Internet of Things Journal 7, 4 (2019), 26112621.Google ScholarGoogle ScholarCross RefCross Ref
  19. Gursoy Mehmet Emre, Liu Ling, Truex Stacey, and Yu Lei. 2018. Differentially private and utility preserving publication of trajectory data. IEEE Transactions on Mobile Computing 18, 10 (2018), 23152329.Google ScholarGoogle ScholarCross RefCross Ref
  20. Hua Jingyu, Gao Yue, and Zhong Sheng. 2015. Differentially private publication of general time-serial trajectory data. In 2015 IEEE Conference on Computer Communications (INFOCOM). 549557. Google ScholarGoogle ScholarCross RefCross Ref
  21. Huang Chunan, Zhang Xueru, Salehi Rasoul, Ersal Tulga, and Stefanopoulou Anna G.. 2020. A robust energy and emissions conscious cruise controller for connected vehicles with privacy considerations. In 2020 American Control Conference (ACC). 48814886. Google ScholarGoogle ScholarCross RefCross Ref
  22. Huitema Bradley E. and McKean Joseph W.. 1991. Autocorrelation estimation and inference with small samples.Psychological Bulletin 110, 2 (1991), 291.Google ScholarGoogle ScholarCross RefCross Ref
  23. Kairouz Peter, Oh Sewoong, and Viswanath Pramod. 2017. The composition theorem for differential privacy. IEEE Transactions on Information Theory 63, 6 (2017), 40374049.Google ScholarGoogle ScholarDigital LibraryDigital Library
  24. Kasiviswanathan Shiva Prasad, Lee Homin K., Nissim Kobbi, Raskhodnikova Sofya, and Smith Adam. 2011. What can we learn privately?SIAM J. Comput. 40, 3 (2011), 793826.Google ScholarGoogle ScholarDigital LibraryDigital Library
  25. Kellaris Georgios and Papadopoulos Stavros. 2013. Practical differential privacy via grouping and smoothing. In Proceedings of the VLDB Endowment, Vol. 6. VLDB Endowment, 301312.Google ScholarGoogle ScholarDigital LibraryDigital Library
  26. Kellaris Georgios, Papadopoulos Stavros, Xiao Xiaokui, and Papadias Dimitris. 2014. Differentially private event sequences over infinite streams. Proceedings of the VLDB Endowment 7, 12 (2014), 11551166.Google ScholarGoogle ScholarDigital LibraryDigital Library
  27. Khalili Mohammad Mahdi, Zhang Xueru, and Abroshan Mahed. 2021b. Fair sequential selection using supervised learning models. Advances in Neural Information Processing Systems 34 (2021).Google ScholarGoogle Scholar
  28. Khalili Mohammad Mahdi, Zhang Xueru, Abroshan Mahed, and Sojoudi Somayeh. 2021a. Improving fairness and privacy in selection problems. In Proceedings of the AAAI Conference on Artificial Intelligence, Vol. 35. 80928100.Google ScholarGoogle ScholarCross RefCross Ref
  29. Khalili Mohammad Mahdi, Zhang Xueru, and Liu Mingyan. 2019. Contract design for purchasing private data using a biased differentially private algorithm. In Proceedings of the 14th Workshop on the Economics of Networks, Systems and Computation. 16.Google ScholarGoogle ScholarDigital LibraryDigital Library
  30. Khalili Mohammad Mahdi, Zhang Xueru, and Liu Mingyan. 2021c. Designing contracts for trading private and heterogeneous data using a biased differentially private algorithm. IEEE Access 9 (2021), 7073270745.Google ScholarGoogle ScholarCross RefCross Ref
  31. Li Chao, Li Daniel Yang, Miklau Gerome, and Suciu Dan. 2014. A theory of pricing private data. ACM Transactions on Database Systems (TODS) 39, 4 (2014), 128.Google ScholarGoogle ScholarDigital LibraryDigital Library
  32. Liu Changchang, Chakraborty Supriyo, and Mittal Prateek. 2016. Dependence makes you vulnerable: Differential privacy under dependent tuples. In NDSS, Vol. 16. 2124.Google ScholarGoogle Scholar
  33. McKenzie Ed. 1985. Some simple models for discrete variate time series. JAWRA Journal of the American Water Resources Association 21, 4 (1985), 645650.Google ScholarGoogle ScholarCross RefCross Ref
  34. Perrier Victor, Asghar Hassan Jameel, and Kaafar Dali. 2018. Private continual release of real-valued data streams. arXiv preprint arXiv:1811.03197 (2018).Google ScholarGoogle Scholar
  35. Rastogi Vibhor and Nath Suman. 2010. Differentially private aggregation of distributed time-series with transformation and encryption. In Proceedings of the 2010 ACM SIGMOD International Conference on Management of data. ACM, 735746.Google ScholarGoogle ScholarDigital LibraryDigital Library
  36. Song Shuang, Wang Yizhen, and Chaudhuri Kamalika. 2017. Pufferfish privacy mechanisms for correlated data. In Proceedings of the 2017 ACM International Conference on Management of Data. 12911306.Google ScholarGoogle ScholarDigital LibraryDigital Library
  37. Task Christine and Clifton Chris. 2012. A guide to differential privacy theory in social network analysis. In 2012 IEEE/ACM International Conference on Advances in Social Networks Analysis and Mining. IEEE, 411417.Google ScholarGoogle ScholarDigital LibraryDigital Library
  38. Wang Hao, Xu Zhengquan, Xiong Lizhi, and Wang Tao. 2017. Conducting correlated Laplace mechanism for differential privacy. In Cloud Computing and Security, Sun Xingming, Chao Han-Chieh, You Xingang, and Bertino Elisa (Eds.). Springer International Publishing, Cham, 7285.Google ScholarGoogle ScholarCross RefCross Ref
  39. Wei Kang, Li Jun, Ding Ming, Ma Chuan, Yang Howard H., Farokhi Farhad, Jin Shi, Quek Tony Q. S., and Poor H. Vincent. 2020. Federated learning with differential privacy: Algorithms and performance analysis. IEEE Transactions on Information Forensics and Security 15 (2020), 34543469.Google ScholarGoogle ScholarDigital LibraryDigital Library
  40. Wei William W. S.. 2006. Time series analysis. In The Oxford Handbook of Quantitative Methods in Psychology: Vol. 2.Google ScholarGoogle Scholar
  41. Weiß Christian H.. 2009. Monitoring correlated processes with binomial marginals. Journal of Applied Statistics 36, 4 (2009), 399414.Google ScholarGoogle ScholarCross RefCross Ref
  42. Xiao Yonghui and Xiong Li. 2015. Protecting locations with differential privacy under temporal correlations. In Proceedings of the 22nd ACM SIGSAC Conference on Computer and Communications Security. ACM, 12981309.Google ScholarGoogle ScholarDigital LibraryDigital Library
  43. Yang Bin, Sato Issei, and Nakagawa Hiroshi. 2015. Bayesian differential privacy on correlated data. In Proceedings of the 2015 ACM SIGMOD International Conference on Management of Data. 747762.Google ScholarGoogle ScholarDigital LibraryDigital Library
  44. Zhang Xueru, Huang Chunan, Liu Mingyan, Stefanopoulou Anna, and Ersal Tulga. 2019. Predictive cruise control with private vehicle-to-vehicle communication for improving fuel consumption and emissions. IEEE Communications Magazine 57, 10 (2019), 9197. Google ScholarGoogle ScholarCross RefCross Ref
  45. Zhang Xueru, Khalili Mohammad Mahdi, and Liu Mingyan. 2018a. Improving the privacy and accuracy of ADMM-based distributed algorithms. In International Conference on Machine Learning. PMLR, 57965805.Google ScholarGoogle Scholar
  46. Zhang Xueru, Khalili Mohammad Mahdi, and Liu Mingyan. 2018b. Recycled ADMM: Improve privacy and accuracy with less computation in distributed algorithms. In 2018 56th Annual Allerton Conference on Communication, Control, and Computing (Allerton). IEEE, 959965.Google ScholarGoogle ScholarDigital LibraryDigital Library
  47. Zhang Xueru, Khalili Mohammad Mahdi, and Liu Mingyan. 2020. Recycled ADMM: Improving the privacy and accuracy of distributed algorithms. IEEE Transactions on Information Forensics and Security 15 (2020), 17231734. Google ScholarGoogle ScholarDigital LibraryDigital Library
  48. Zheng Xiao. 2020. Data trading with differential privacy in data market. In Proceedings of 2020 the 6th International Conference on Computing and Data Engineering. 112115.Google ScholarGoogle ScholarDigital LibraryDigital Library
  49. Zhu Tianqing, Xiong Ping, Li Gang, and Zhou Wanlei. 2015. Correlated differential privacy: Hiding information in non-IID data set. IEEE Transactions on Information Forensics and Security 10, 2 (2015), 229242. Google ScholarGoogle ScholarDigital LibraryDigital Library

Index Terms

  1. Differentially Private Real-Time Release of Sequential Data

    Recommendations

    Comments

    Login options

    Check if you have access through your login credentials or your institution to get full access on this article.

    Sign in

    Full Access

    • Published in

      cover image ACM Transactions on Privacy and Security
      ACM Transactions on Privacy and Security  Volume 26, Issue 1
      February 2023
      342 pages
      ISSN:2471-2566
      EISSN:2471-2574
      DOI:10.1145/3561959
      Issue’s Table of Contents

      Permission to make digital or hard copies of all or part of this work for personal or classroom use is granted without fee provided that copies are not made or distributed for profit or commercial advantage and that copies bear this notice and the full citation on the first page. Copyrights for components of this work owned by others than ACM must be honored. Abstracting with credit is permitted. To copy otherwise, or republish, to post on servers or to redistribute to lists, requires prior specific permission and/or a fee. Request permissions from [email protected].

      Publisher

      Association for Computing Machinery

      New York, NY, United States

      Publication History

      • Published: 7 November 2022
      • Online AM: 23 June 2022
      • Accepted: 16 May 2022
      • Revised: 3 March 2022
      • Received: 13 October 2021
      Published in tops Volume 26, Issue 1

      Permissions

      Request permissions about this article.

      Request Permissions

      Check for updates

      Qualifiers

      • research-article
      • Refereed
    • Article Metrics

      • Downloads (Last 12 months)1,023
      • Downloads (Last 6 weeks)132

      Other Metrics

    PDF Format

    View or Download as a PDF file.

    PDF

    eReader

    View online with eReader.

    eReader

    HTML Format

    View this article in HTML Format .

    View HTML Format