Skip to main content
Log in

Decentralized bilevel optimization

  • Original Paper
  • Published:
Optimization Letters Aims and scope Submit manuscript

Abstract

Bilevel optimization has been successfully applied to many important machine learning problems. Algorithms for solving bilevel optimization have been studied under various settings. In this paper, we study the nonconvex-strongly-convex bilevel optimization under a decentralized setting. We design decentralized algorithms for both deterministic and stochastic bilevel optimization problems. Moreover, we analyze the convergence rates of the proposed algorithms in difference scenarios including the case where data heterogeneity is observed across agents. Numerical experiments on both synthetic and real data demonstrate that the proposed methods are efficient.

This is a preview of subscription content, log in via an institution to check access.

Access this article

Price excludes VAT (USA)
Tax calculation will be finalised during checkout.

Instant access to the full article PDF.

Algorithm 1
Algorithm 2
Algorithm 3
Algorithm 4
Algorithm 5
Fig. 1
Fig. 2

Similar content being viewed by others

Data availability

All data sets used in this paper are publicly available.

Notes

  1. For simplicity we use constant stepsize \(\eta _x\) in the outer loop. Similar results can be obtained for diminishing stepsizes.

  2. http://qwone.com/~jason/20Newsgroups/.

References

  1. Yang, S., Zhang, X., Wang, M.: Decentralized gossip-based stochastic bilevel optimization over communication networks. arXiv:2206.10870 (2022)

  2. Gao, H., Gu, B., Thai, M.T.: Stochastic bilevel distributed optimization over a network. arXiv:2206.15025 (2022)

  3. Terashita, N., Hara, S.: Personalized decentralized bilevel optimization over stochastic and directed networks. arXiv:2210.02129 (2022)

  4. Chen, X., Huang, M., Ma, S., Balasubramanian, K.: Decentralized stochastic bilevel optimization with improved per-iteration complexity. In: International Conference on Machine Learning, pp. 4641–4671. PMLR (2023)

  5. Jiao, Y., Yang, K., Wu, T., Song, D., Jian, C.: Asynchronous distributed bilevel optimization. arXiv:2212.10048 (2022)

  6. Snell, J., Swersky, K., Zemel, R.: Prototypical networks for few-shot learning. Adv. Neural Inf. Process. Syst. 30 (2017)

  7. Bertinetto, L., Henriques, J.F., Torr, P.H., Vedaldi, A.: Meta-learning with differentiable closed-form solvers. arXiv:1805.08136 (2018)

  8. Rajeswaran, A., Finn, C., Kakade, S.M., Levine, S.: Meta-learning with implicit gradients. Adv. Neural Inf. Process. Syst. 32 (2019)

  9. Pedregosa, F.: Hyperparameter optimization with approximate gradient. In: International Conference on Machine Learning, pp. 737–746. PMLR (2016)

  10. Franceschi, L., Frasconi, P., Salzo, S., Grazzi, R., Pontil, M.: Bilevel programming for hyperparameter optimization and meta-learning. In: International Conference on Machine Learning, pp. 1568–1577. PMLR (2018)

  11. Hong, M., Wai, H.-T., Wang, Z., Yang, Z.: A two-timescale framework for bilevel optimization: complexity analysis and application to actor-critic. arXiv:2007.05170 (2020)

  12. Ghadimi, S., Wang, M.: Approximation methods for bilevel programming. arXiv:1802.02246 (2018)

  13. Ji, K., Yang, J., Liang, Y.: Bilevel optimization: convergence analysis and enhanced design. In: International Conference on Machine Learning, pp. 4882–4892. PMLR (2021)

  14. Chen, T., Sun, Y., Yin, W.: Closing the gap: tighter analysis of alternating stochastic gradient methods for bilevel problems. Adv. Neural Inf. Process. Syst. 34, 25294–25307 (2021)

    Google Scholar 

  15. Lian, X., Zhang, C., Zhang, H., Hsieh, C.-J., Zhang, W., Liu, J.: Can decentralized algorithms outperform centralized algorithms? A case study for decentralized parallel stochastic gradient descent. Adv. Neural Inf. Process. Syst. 30, 5336–5346 (2017)

    Google Scholar 

  16. Tang, H., Lian, X., Yan, M., Zhang, C., Liu, J.: \(d^2\): decentralized training over decentralized data. In: International Conference on Machine Learning, pp. 4848–4856. PMLR (2018)

  17. Stackelberg, H.V.: Theory of the Market Economy (1952)

  18. Bracken, J., McGill, J.T.: Mathematical programs with optimization problems in the constraints. Oper. Res. 21(1), 37–44 (1973)

    Article  MathSciNet  Google Scholar 

  19. Bennett, K., Ji, X., Hu, J., Kunapuli, G., Pang, J.: Model selection via bilevel programming. In: Proceedings of the International Joint Conference on Neural Networks (IJCNN’06) Vancouver, BC Canada, pp. 1922–1929 (2006)

  20. Kunapuli, G., Bennett, K., Hu, J., Pang, J.-S.: Bilevel model selection for support vector machines. In: CRM Proceedings and Lecture Notes, vol. 45, pp. 129–158 (2008)

  21. Kunapuli, G., Bennett, K.P., Hu, J., Pang, J.-S.: Classification model selection via bilevel programming. Optim. Methods Softw. 23(4), 475–489 (2008)

    Article  MathSciNet  Google Scholar 

  22. Domke, J.: Generic methods for optimization-based modeling. In: Artificial Intelligence and Statistics, pp. 318–326. PMLR (2012)

  23. Gould, S., Fernando, B., Cherian, A., Anderson, P., Cruz, R.S., Guo, E.: On differentiating parameterized argmin and argmax problems with application to bi-level optimization. arXiv:1607.05447 (2016)

  24. Grazzi, R., Franceschi, L., Pontil, M., Salzo, S.: On the iteration complexity of hypergradient computation. In: International Conference on Machine Learning, pp. 3748–3758. PMLR (2020)

  25. Maclaurin, D., Duvenaud, D., Adams, R.: Gradient-based hyperparameter optimization through reversible learning. In: International Conference on Machine Learning, pp. 2113–2122. PMLR (2015)

  26. Chen, T., Sun, Y., Xiao, Q., Yin, W.: A single-timescale method for stochastic bilevel optimization. In: International Conference on Artificial Intelligence and Statistics, pp. 2466–2488. PMLR (2022)

  27. Guo, Z., Hu, Q., Zhang, L., Yang, T.: Randomized stochastic variance-reduced methods for multi-task stochastic bilevel optimization. arXiv:2105.02266 (2021)

  28. Khanduri, P., Zeng, S., Hong, M., Wai, H.-T., Wang, Z., Yang, Z.: A near-optimal algorithm for stochastic bilevel optimization via double-momentum. Adv. Neural Inf. Process. Syst. 34, 30271–30283 (2021)

    Google Scholar 

  29. Yang, J., Ji, K., Liang, Y.: Provably faster algorithms for bilevel optimization. Adv. Neural Inf. Process. Syst. 34, 13670–13682 (2021)

    Google Scholar 

  30. Gan, S., Lian, X., Wang, R., Chang, J., Liu, C., Shi, H., Zhang, S., Li, X., Sun, T., Jiang, J., et al.: Bagua: scaling up distributed learning with system relaxations. arXiv:2107.01499 (2021)

  31. Yuan, B., He, Y., Davis, J., Zhang, T., Dao, T., Chen, B., Liang, P.S., Re, C., Zhang, C.: Decentralized training of foundation models in heterogeneous environments. Adv. Neural Inf. Process. Syst. 35, 25464–25477 (2022)

    Google Scholar 

  32. Xu, J., Zhu, S., Soh, Y.C., Xie, L.: Augmented distributed gradient methods for multi-agent optimization under uncoordinated constant stepsizes. In: 2015 54th IEEE Conference on Decision and Control (CDC), pp. 2055–2060. IEEE (2015)

  33. Di Lorenzo, P., Scutari, G.: Next: in-network nonconvex optimization. IEEE Trans. Signal Inf. Process. Netw. 2(2), 120–136 (2016)

    MathSciNet  Google Scholar 

  34. Qu, G., Li, N.: Harnessing smoothness to accelerate distributed optimization. IEEE Trans. Control Netw. Syst. 5(3), 1245–1260 (2017)

    Article  MathSciNet  Google Scholar 

  35. Nedic, A., Olshevsky, A., Shi, W.: Achieving geometric convergence for distributed optimization over time-varying graphs. SIAM J. Optim. 27(4), 2597–2633 (2017)

    Article  MathSciNet  Google Scholar 

  36. Xin, R., Kar, S., Khan, U.A.: Decentralized stochastic optimization and machine learning: a unified variance-reduction framework for robust performance and fast convergence. IEEE Signal Process. Mag. 37(3), 102–113 (2020)

    Article  Google Scholar 

  37. Altae-Tran, H., Ramsundar, B., Pappu, A.S., Pande, V.: Low data drug discovery with one-shot learning. ACS Cent. Sci. 3(4), 283–293 (2017)

    Article  Google Scholar 

  38. Zhang, X.S., Tang, F., Dodge, H.H., Zhou, J., Wang, F.: Metapred: meta-learning for clinical risk prediction with limited patient electronic health records. In: Proceedings of the 25th ACM SIGKDD International Conference on Knowledge Discovery and Data Mining, pp. 2487–2495 (2019)

  39. Kayaalp, M., Vlaski, S., Sayed, A.H.: Dif-MAML: decentralized multi-agent meta-learning. IEEE Open J. Signal Process. 3, 71–93 (2022)

    Article  Google Scholar 

  40. Tarzanagh, D.A., Li, M., Thrampoulidis, C., Oymak, S.: Fednest: Federated bilevel, minimax, and compositional optimization. arXiv:2205.02215 (2022)

  41. Li, J., Huang, F., Huang, H.: Local stochastic bilevel optimization with momentum-based variance reduction. arXiv: 2205.01608 (2022)

  42. Xian, W., Huang, F., Zhang, Y., Huang, H.: A faster decentralized algorithm for nonconvex minimax problems. Adv. Neural Inf. Process. Syst. 34, 25865–25877 (2021)

    Google Scholar 

  43. Luo, L., Ye, H.: Decentralized stochastic variance reduced extragradient method. arXiv:2202.00509 (2022)

  44. Sharma, P., Panda, R., Joshi, G., Varshney, P.K.: Federated minimax optimization: improved convergence analyses and algorithms. arXiv:2203.04850 (2022)

  45. Lu, S., Cui, X., Squillante, M.S., Kingsbury, B., Horesh, L.: Decentralized bilevel optimization for personalized client learning. In: ICASSP 2022-2022 IEEE International Conference on Acoustics, Speech and Signal Processing (ICASSP), pp. 5543–5547. IEEE (2022)

  46. Pu, S., Nedić, A.: Distributed stochastic gradient tracking methods. Math. Program. 187(1), 409–457 (2021)

    Article  MathSciNet  Google Scholar 

  47. Mota, J.F., Xavier, J.M., Aguiar, P.M., Püschel, M.: D-ADMM: a communication-efficient distributed algorithm for separable optimization. IEEE Trans. Signal process. 61(10), 2718–2723 (2013)

    Article  MathSciNet  Google Scholar 

  48. Chang, T.-H., Hong, M., Wang, X.: Multi-agent distributed optimization via inexact consensus ADMM. IEEE Trans. Signal Process. 63(2), 482–497 (2014)

    Article  MathSciNet  Google Scholar 

  49. Shi, W., Ling, Q., Yuan, K., Wu, G., Yin, W.: On the linear convergence of the ADMM in decentralized consensus optimization. IEEE Trans. Signal Process. 62(7), 1750–1761 (2014)

    Article  MathSciNet  Google Scholar 

  50. Aybat, N.S., Wang, Z., Lin, T., Ma, S.: Distributed linearized alternating direction method of multipliers for composite convex consensus optimization. IEEE Trans. Autom. Control 63(1), 5–20 (2017)

    Article  MathSciNet  Google Scholar 

  51. Makhdoumi, A., Ozdaglar, A.: Convergence rate of distributed ADMM over networks. IEEE Trans. Autom. Control 62(10), 5082–5095 (2017)

    Article  MathSciNet  Google Scholar 

  52. Koloskova, A., Lin, T., Stich, S.U.: An improved analysis of gradient tracking for decentralized machine learning. Adv. Neural Inf. Process. Syst. 34, 11422–11435 (2021)

    Google Scholar 

  53. Shaban, A., Cheng, C.-A., Hatch, N., Boots, B.: Truncated back-propagation for bilevel optimization. In: The 22nd International Conference on Artificial Intelligence and Statistics, pp. 1723–1732. PMLR (2019)

  54. LeCun, Y., Bottou, L., Bengio, Y., Haffner, P.: Gradient-based learning applied to document recognition. Proc. IEEE 86(11), 2278–2324 (1998)

    Article  Google Scholar 

  55. Olshevsky, A., Paschalidis, I.C., Pu, S.: A non-asymptotic analysis of network independence for distributed stochastic gradient descent. arXiv:1906.02702 (2019)

Download references

Acknowledgements

We would like to thank the Guest Editor and two anonymous reviewers whose insightful comments have helped improve the presentation of this paper. The research of Shiqian Ma was supported in part by NSF Grants DMS-2243650, CCF-2308597, CCF-2311275 and ECCS-2326591, UC Davis CeDAR (Center for Data Science and Artificial Intelligence Research) Innovative Data Science Seed Funding Program, and a startup fund from Rice University.

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to Shiqian Ma.

Ethics declarations

Conflict of interest

The authors declare that they have no conflict of interest.

Additional information

Publisher's Note

Springer Nature remains neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Here we provide a brief history of this paper. This paper appeared on arxiv on 06/12/2022 (arxiv ID 2206.05670). To the best of our knowledge, this is the first paper discussing decentralized algorithms for bilevel optimization. All other papers on the same topic appeared later than ours, including [1] which appeared on arxiv on 06/22/2022 (arxiv ID 2206.10870), Gao et al. [2] which appeared on arxiv on 06/30/2022 (arxiv ID 2206.15025), Terashita and Hara [3] which appeared on 10/05/2022 (arxiv ID 2210.02129), Chen et al. [4] which appeared on 10/23/2022 (arxiv ID 2210.12839), and [5] which appeared on arxiv on 12/20/2022 (arxiv ID 2212.10048). This paper was first submitted to NeurIPS 2022 and was rejected, although the reviewers did not raise any questions about the novelty and correctness.

Appendices

Appendix 1: Details about experiments and other results

In this section we provide details about our experiments as well as results about training and test loss. For each experiment, we set our network topology as a special ring network, where \(W= (w_{i,j})\) and the only nonzero elements are given by:

$$\begin{aligned} w_{i,i} = a,\ w_{i,i+1}=w_{i,i-1} = \frac{1-a}{2},\ \text { for some }a\in (0,1). \end{aligned}$$

Here we overload the notation and set \(w_{n,n+1}= w_{n,1}, w_{1,0}=w_{1,n}\). Note that a is the unique parameter that determines the weight matrix and will be specified in each experiment.

1.1 Synthetic data

1.1.1 Logistic regression on synthetic data

In this experiment, on node i we have:

$$\begin{aligned}&f_i(\lambda ,\tau ^*(\lambda )) = \sum _{(x_e,y_e)\in {\mathcal {D}}_i'}\psi (y_ex_e^{{\textsf{T}}}\tau ^*(\lambda )),\\&g_i(\lambda ,\tau ) = \sum _{(x_e,y_e)\in {\mathcal {D}}_i}\psi (y_ex_e^{{\textsf{T}}}\tau ) + \frac{1}{2}\tau ^{{\textsf{T}}}\text {diag}(e^{\lambda })\tau , \end{aligned}$$

where \(e^{\lambda }\) is element-wise, \(\text {diag}(v)\) denotes the diagonal matrix generated by vector v, and \(\psi (x) = \log (1+e^{-x})\). \({\mathcal {D}}_i'\) and \({\mathcal {D}}_i\) represent validation set and training set on node i. Following the setup in [24], we first randomly generate \(\tau ^*\in {\mathbb {R}}^{p}\) and the noise vector \(\epsilon \in {\mathbb {R}}^{p}\). For the data point \((x_e, y_e)\) on node i, each element of \(x_e\) is sampled from the normal distribution with mean 0, variance \(i^2\). \(y_e\) is then set by \(y_e = \text {sign}(x_e^{{\textsf{T}}}\tau ^* + m\epsilon )\), where \(\text {sign}\) denotes the sign function and \(m=0.1\) denotes the noise rate. In the experiment we choose \(p = q = 50,\) and the number of inner-loop and outer-loop iterations as 10 and 100 respectively. N,  the number of iterations of the JHIP oracle 1 is 20. The stepsizes are \(\eta _x=\eta _y=\gamma = 0.01.\) The number of agents n is chosen as 20,  and the weight parameter \(a=0.4\) (Fig. 3).

Fig. 3
figure 3

Logistic regression on synthetic data

1.2 Real-world data

1.2.1 Logistic regression on 20 Newsgroup dataset

In this experiment, on node i we have:

$$\begin{aligned}&f_i(\lambda , \tau ^*(\lambda ))= \frac{1}{|{\mathcal {D}}^{(i)}_{val}|}\sum _{(x_e,y_e)\in {\mathcal {D}}^{(i)}_{val}}L(x_e^{{\textsf{T}}}\tau ^*, y_e), \\&g_i(\lambda , \tau )= \frac{1}{|{\mathcal {D}}_{tr}^{(i)}|}\sum _{(x_e,y_e)\in {\mathcal {D}}^{(i)}_{tr}}L(x_e^{{\textsf{T}}}\tau , y_e) + \frac{1}{cp} \sum _{i=1}^{c}\sum _{j=1}^{p}e^{\lambda _j}\tau _{ij}^2, \end{aligned}$$

where \(c=20\) denotes the number of topics, \(p=101631\) is the feature dimension, L is the cross entropy loss, \({\mathcal {D}}_{val}\) and \({\mathcal {D}}_{tr}\) are the validation and training data sets, respectively. Our codes can be seen as decentralized versions of the one provided in [13].

We first set inner and outer stepsizes \(\eta _x=\eta _y=100\) (the same as the ones used in [13]), and then compare its performance with different stepsizes. We set the number of inner-loop iterations \(T = 10,\) the number of outer-loop iterations \(K=30,\) the number of agents \(n=20,\) and the weight parameter \(a=0.33\). At the end of jth outer-loop iteration we use the average \(\overline{\tau _j} = \frac{1}{n}\sum _{i=1}^{n}\tau _{i,j}\) as the model parameter and then do the classification on the test set to get the test accuracy (Fig. 4).

Fig. 4
figure 4

Logistic regression on 20 Newsgroup dataset

1.2.2 Data hyper-cleaning on MNIST

In this experiment, on node i we have:

$$\begin{aligned} f_i(\lambda , \tau )&= \frac{1}{\vert {\mathcal {D}}^{(i)}_{val}\vert }\sum _{(x_e,y_e)\in {\mathcal {D}}^{(i)}_{val}}L(x_e^{{\textsf{T}}}\tau , y_e), \\ g_i(\lambda , \tau )&= \frac{1}{\vert {\mathcal {D}}^{(i)}_{tr}\vert }\sum _{(x_e,y_e)\in {\mathcal {D}}^{(i)}_{tr}}\sigma (\lambda _e)L(x_e^{{\textsf{T}}}\tau , y_e) + C_r\Vert \tau \Vert ^2, \end{aligned}$$

where L is the cross-entropy loss and \(\sigma (x)= (1+e^{-x})^{-1}\) is the sigmoid function. The number of inner-loop iterations T and outer-loop iterations K are set as 10 and 30, respectively. The number of agents \(n=20\) and the weight parameter \(a=0.5\). Following [13, 53] the regularization parameter \(C_r\) is set as 0.001. We first choose stepsizes similar to those in [13] and then set larger stepsizes. In each iteration we evaluate the norm of the hypergradient at the average of the hyperparameters \({\bar{\lambda }}\), and plot the logarithm (base 10) of the norm of the hypergradient versus iteration number in Fig. 2 (Figs. 5, 6).

Fig. 5
figure 5

Data hyper-cleaning on MNIST

Fig. 6
figure 6

Data hyper-cleaning on MNIST

Appendix 2: Convergence analysis

In this section we provide the proofs of convergence results. For convenience, we first list the notation below.

$$\begin{aligned}&W := (w_{ij}) \text { is symmetric doubly stochastic, and } \rho := \max \left( |\lambda _2|, |\lambda _n|\right) < 1 \\&X_k := \left( x_{1,k},x_{2,k},\ldots ,x_{n,k}\right) ,\ {{\bar{x}}}_k := \frac{1}{n}\sum _{i=1}^{n}x_{i,k},\\&\partial \Phi (X_k) := \left( {\hat{\nabla }} f_1(x_{1,k}, y_{1,k}^{(T)}),\ldots ,{\hat{\nabla }} f_n(x_{n,k}, y_{n,k}^{(T)})\right) ,\\&\partial \Phi (X_k;\phi ) := \left( {\hat{\nabla }} f_1(x_{1,k}, y_{1,k}^{(T)};\phi _{1,k}),\ldots ,{\hat{\nabla }} f_n(x_{n,k}, y_{n,k}^{(T)};\phi _{n,k})\right) \\&\overline{\partial \Phi (X_k)} := \frac{1}{n}\sum _{i=1}^{n}{\hat{\nabla }}f_{i}(x_{i,k}, y_{i,k}^{(T)}),\ \overline{\partial \Phi (X_k;\phi )} := \frac{1}{n}\sum _{i=1}^{n}{\hat{\nabla }}f_{i}(x_{i,k}, y_{i,k}^{(T)}; \phi _{i,k}), \\&q_{i,k} := x_{i,k} - {{\bar{x}}}_k,\ r_{i,k} := u_{i,k} - {{\bar{u}}}_k, \\&Q_k := \left( q_{1,k}, q_{2,k},\ldots ,q_{n,k}\right) ,\ R_k := \left( r_{1,k},r_{2,k},\ldots ,r_{n,k}\right) \in {\mathbb {R}}^{p\times n}, \\&S_K := \sum _{k=1}^{K}\Vert Q_k\Vert ^2,\ T_K := \sum _{j=0}^{K}\Vert \nabla \Phi ({{\bar{x}}}_j)\Vert ^2,\ E_K := \sum _{j=1}^{K}\sum _{i=1}^{n}\Vert x_{i,j} - x_{i,j-1}\Vert ^2, \\&A_K := \sum _{j=0}^{K}\sum _{i=1}^{n}\Vert y_{i,j}^{(T)} - y_i^*(x_{i,j})\Vert ^2, B_K := \sum _{j=0}^{K}\sum _{i=1}^{n}\Vert v_{i,j}^* - v_{i,j}^{(0)}\Vert ^2, \\&v_{i,j}^* = \left( \nabla _y^2g_i(x_{i,j},y_i^*(x_{i,j}))\right) ^{-1}\nabla _y f_i(x_{i,j},y_i^*(x_{i,j})), \\&\delta _y := (1 - \eta _y\mu )^2,\ \delta _{\kappa } := \left( \frac{\sqrt{\kappa } - 1}{\sqrt{\kappa } + 1}\right) ^2. \end{aligned}$$

We first introduce a few lemmas that are useful in the proofs.

Lemma 1

For any \(p, q, r\in {\mathbb {N}}_+\) and matrix \(A\in {\mathbb {R}}^{p\times q}, B\in {\mathbb {R}}^{q\times r}\), we have:

$$\begin{aligned} \Vert AB\Vert \le \min \left( \Vert A\Vert _2\cdot \Vert B\Vert , \Vert A\Vert \cdot \Vert B^{{\textsf{T}}}\Vert _2\right) . \end{aligned}$$

Lemma 2

For any matrix \(A = (a_1,a_2,\ldots ,a_q)\in {\mathbb {R}}^{p\times q}\), we have:

$$\begin{aligned} \Vert a_j\Vert ^2\le \Vert A\Vert _2^2 \le \Vert A\Vert ^2 = \sum _{i=1}^{q}\Vert a_i\Vert ^2,\ \forall {j}\in \{1,2,\ldots ,q\}. \end{aligned}$$

For one-step gradient descent, we have the following result (see, e.g., Lemma 10 in [34] and Lemma 3 in [46]).

Lemma 3

Suppose f(x) is \(\mu\)-strongly convex and \(L-smooth\). For any x and \(\eta <\frac{2}{\mu + L}\), define \(x^+ = x - \eta \nabla f(x),\ x^*=\mathop {\mathrm {arg\,min}}\limits f(x)\). Then we have

$$\begin{aligned} \Vert x^+ - x^*\Vert \le (1-\eta \mu )\Vert x-x^*\Vert . \end{aligned}$$

The following lemma is a common result in decentralized optimization (e.g., [15, Lemma 4]).

Lemma 4

Suppose Assumption 2.2 holds. We have for any integer \(k\ge 0\),

$$\begin{aligned} \left\| W^{k} - \frac{{\textbf{1}}_n{\textbf{1}}^{\top }_n}{n}\right\| _2\le \rho ^k. \end{aligned}$$

Proof

Assume \(1=\lambda _1>\lambda _2\ge \cdots \ge \lambda _n>-1\) are eigenvalues of W. Since \(W^{k}{\textbf{1}}_n{\textbf{1}}^{\top }_n = {\textbf{1}}_n{\textbf{1}}^{\top }_nW^k\), we know \(W^k\) and \({\textbf{1}}_n{\textbf{1}}^{\top }_n\) are simultaneously diagonalizable. Hence there exists an orthogonal matrix P such that

$$\begin{aligned} W^k = P\text {diag}(\lambda _i^k)P^{-1},\quad \frac{{\textbf{1}}_n{\textbf{1}}^{\top }_n}{n}=P\text {diag}(1,0,0,\ldots ,0)P^{-1}, \end{aligned}$$

and thus:

$$\begin{aligned} \left\| W^{k} - \frac{{\textbf{1}}_n{\textbf{1}}^{\top }_n}{n}\right\| _2 = \left\| P(\text {diag}(\lambda _i^{k}) - \text {diag}(1,0,0,\ldots ,0) )P^{-1}\right\| _2\le \max \left( |\lambda _2|^k, |\lambda _n|^k\right) . \end{aligned}$$

By definition of rho, the proof is complete. \(\square\)

The following three lemmas are adopted from Lemma 2.2 in [12]:

Lemma 5

(Hypergradient) Define \(\Phi _i(x):= f_i(x, y^*(x))\), where \(y^*(x) = {\mathrm{arg~min}}_{y\in {\mathbb {R}}^{q}}g(x,y)\). Under Assumption 2.1 we have:

$$\begin{aligned} \nabla \Phi _i(x) = \nabla _x f_i(x,y^*(x)) - \nabla _{xy} g(x,y^*(x))\left( \nabla _y^2g(x,y^*(x))\right) ^{-1}\nabla _y f_i(x,y^*(x)). \end{aligned}$$

Moreover, \(\nabla \Phi _i\) is Lipschitz continuous:

$$\begin{aligned} \Vert \nabla \Phi _i(x_1) - \nabla \Phi _i(x_2)\Vert \le L_{\Phi }\Vert x_1 - x_2\Vert , \end{aligned}$$

with the Lipschitz constant given by:

$$\begin{aligned} L_{\Phi } = L + \frac{2L^2 + L_{g,2}L_{f,0}^2 }{\mu } + \frac{LL_{f,0}L_{g,2}+L^3 + L_{g,2}L_{f,0}L }{\mu ^2} + \frac{L_{g,2}L^2L_{f,0}}{\mu ^3} = \Theta (\kappa ^3). \end{aligned}$$

Remark

if Assumption 2.3 does not hold, then this hypergradient is completely different from the local hypergradient:

$$\begin{aligned} \nabla f_i(x, y_i^*(x)) = \nabla _x f_i(x,y_i^*(x)) - \nabla _{xy} g_i(x,y_i^*(x))\left( \nabla _y^2g_i(x,y_i^*(x))\right) ^{-1}\nabla _y f_i(x,y_i^*(x)), \end{aligned}$$
(11)

where \(y_i^*(x)={\mathrm{arg~min}}_{y\in {\mathbb {R}}^{q}}g_i(x,y)\).

Lemma 6

Define:

$$\begin{aligned} {\bar{\nabla }} f_i(x, y) = \nabla _x f_i(x,y) - \nabla _{xy} g(x,y) \left( \nabla _y^2g(x,y)\right) ^{-1}\nabla _y f_i(x,y). \end{aligned}$$

Under the Assumption 2.1 we have:

$$\begin{aligned} \Vert {\bar{\nabla }} f_i(x, y) - {\bar{\nabla }} f_i({\tilde{x}}, {\tilde{y}}) \Vert \le L_{f}\Vert (x,y) - ({\tilde{x}}, {\tilde{y}})\Vert , \end{aligned}$$

where the Lipschitz constant is given by:

$$\begin{aligned} L_{f} = L + \frac{L^2}{\mu } + L_{f,0}\left( \frac{L_{g,2}}{\mu } + \frac{L_{g,2}L}{\mu ^2} \right) = \Theta (\kappa ). \end{aligned}$$

Lemma 7

Suppose Assumption 2.1 holds. We have:

$$\begin{aligned} \Vert y_i^*(x_1) - y_i^*(x_2)\Vert \le \kappa \Vert x_1 - x_2\Vert ,\quad \forall i\in \{1,2,\ldots ,n\}. \end{aligned}$$

These lemmas reveal some nice properties of functions in bilevel optimization under Assumption 2.1. We will make use of these lemmas in our theoretical analysis.

Lemma 8

Suppose Assumption 2.1 holds. If the iterates satisfy:

$$\begin{aligned} {{\bar{x}}}_{k+1} = {{\bar{x}}}_k - \eta _x\overline{\partial \Phi (X_k)},\quad \text {where } 0<\eta _x\le \frac{1}{L_{\Phi }}, \end{aligned}$$

then we have the following inequality holds:

$$\begin{aligned}\frac{1}{K+1}\sum _{k=0}^{K}\Vert \nabla \Phi ({{\bar{x}}}_k)\Vert ^2 &\le \frac{2}{\eta _x(K+1)}(\Phi ({{\bar{x}}}_0) - \inf _x\Phi (x)) \nonumber \\&\quad+ \frac{1}{K+1}\sum _{k=0}^{K}\Vert \overline{\partial \Phi (X_k)} - \nabla \Phi ({{\bar{x}}}_k)\Vert ^2. \end{aligned}$$
(12)

Proof

Since \(\Phi (x)\) is \(L_{\Phi }\)-smooth, we have:

$$\begin{aligned}&\Phi ({{\bar{x}}}_{k+1}) - \Phi ({{\bar{x}}}_k) \\&\quad \le \nabla \Phi ({{\bar{x}}}_k)^{{\textsf{T}}}(-\eta _x\overline{\partial \Phi (X_k)}) + \frac{L_{\Phi }\eta _x^2}{2}\Vert \overline{\partial \Phi (X_k)}\Vert ^2 \\&\quad =\frac{L_{\Phi }\eta _x^2}{2}\Vert \overline{\partial \Phi (X_k)}\Vert ^2 - \eta _x \nabla \Phi ({{\bar{x}}}_k)^{{\textsf{T}}}\overline{\partial \Phi (X_k)} \\&\quad =\frac{L_{\Phi }\eta _x^2}{2}\Vert \overline{\partial \Phi (X_k)} - \nabla \Phi ({{\bar{x}}}_k)\Vert ^2 + \left( \frac{L_{\Phi }\eta _x^2}{2} - \eta _x\right) \Vert \nabla \Phi ({{\bar{x}}}_k)\Vert ^2 \\&\qquad +\,(L_{\Phi }\eta _x^2- \eta _x) \nabla \Phi (\bar{x}_k)^{{\textsf{T}}}(\overline{\partial \Phi (X_k)} - \nabla \Phi ({{\bar{x}}}_k)) \\&\quad \le \frac{L_{\Phi }\eta _x^2}{2}\Vert \overline{\partial \Phi (X_k)} - \nabla \Phi ({{\bar{x}}}_k)\Vert ^2 + \left( \frac{L_{\Phi }\eta _x^2}{2} - \eta _x\right) \Vert \nabla \Phi ({{\bar{x}}}_k)\Vert ^2 \\&\qquad +\, (\eta _x - L_{\Phi }\eta _x^2)\left( \frac{1}{2}\Vert \overline{\partial \Phi (X_k)} - \nabla \Phi ({{\bar{x}}}_k)\Vert ^2 + \frac{1}{2}\Vert \nabla \Phi ({{\bar{x}}}_k)\Vert ^2 \right) \\&\quad = \frac{\eta _x}{2} \Vert \overline{\partial \Phi (X_k)} - \nabla \Phi ({{\bar{x}}}_k)\Vert ^2 -\frac{\eta _x}{2}\Vert \nabla \Phi ({{\bar{x}}}_k)\Vert ^2, \end{aligned}$$

where the second inequality is due to Young’s inequality and \(\eta _x\le \frac{1}{L_{\Phi }}\). Therefore, we have:

$$\begin{aligned} \begin{aligned} \Vert \nabla \Phi ({{\bar{x}}}_k)\Vert ^2&\le \frac{2}{\eta _x}(\Phi ({{\bar{x}}}_k) - \Phi ({{\bar{x}}}_{k+1})) + \Vert \overline{\partial \Phi (X_k)} - \nabla \Phi ({{\bar{x}}}_k)\Vert ^2. \end{aligned} \end{aligned}$$
(13)

Summing (13) over \(k=0,\ldots ,K\), yields:

$$\begin{aligned} \sum _{k=0}^{K}\Vert \nabla \Phi ({{\bar{x}}}_k)\Vert ^2 \le \frac{2}{\eta _x}(\Phi ({{\bar{x}}}_0) - \Phi ({{\bar{x}}}_{k+1})) + \sum _{k=0}^{K}\Vert \overline{\partial \Phi (X_k)} - \nabla \Phi (\bar{x}_k)\Vert ^2, \end{aligned}$$

which completes the proof. \(\square\)

We have the following lemma which provides an upper bound for \(E_K\):

Lemma 9

In each iteration, if we have \({{\bar{x}}}_{k+1} = {{\bar{x}}}_k - \eta _x\overline{\partial \Phi (X_k)}\), then the following inequality holds:

$$\begin{aligned} E_K \le 8S_K + 4n\eta _x^2\sum _{j=0}^{K-1}\Vert \overline{\partial \Phi (X_{j})} - \nabla \Phi ({{\bar{x}}}_j)\Vert ^2 + 4n\eta _x^2T_{K-1}. \end{aligned}$$

Proof

By the definition of \(E_K\), we have:

$$\begin{aligned} E_K&= \sum _{j=1}^{K}\sum _{i=1}^{n}\Vert x_{i,j} - x_{i,j-1}\Vert ^2 = \sum _{j=1}^{K}\sum _{i=1}^{n}\Vert x_{i,j} - {{\bar{x}}}_j + {{\bar{x}}}_j - \bar{x}_{j-1} + {{\bar{x}}}_{j-1} - x_{i,j-1}\Vert ^2 \\&= \sum _{j=1}^{K}\sum _{i=1}^{n}\Vert q_{i,j} -\eta _x(\overline{\partial \Phi (X_{j-1})} - \nabla \Phi ({{\bar{x}}}_{j-1})) - \eta _x\nabla \Phi ({{\bar{x}}}_{j-1}) - q_{i,j-1}\Vert ^2 \\&\le 4\sum _{j=1}^{K}\sum _{i=1}^{n}(\Vert q_{i,j}\Vert ^2 + \eta _x^2\Vert \overline{\partial \Phi (X_{j-1})} - \nabla \Phi (\bar{x}_{j-1}))\Vert ^2 \\&\quad +\, \eta _x^2\Vert \nabla \Phi ({{\bar{x}}}_{j-1})\Vert ^2 + \Vert q_{i,j-1}\Vert ^2) \\&\le 4\sum _{j=1}^{K}(\Vert Q_j\Vert ^2 + \Vert Q_{j-1}\Vert ^2 +\, n\eta _x^2\Vert \overline{\partial \Phi (X_{j-1})} - \nabla \Phi ({{\bar{x}}}_{j-1})\Vert ^2 + n\eta _x^2\Vert \nabla \Phi (\bar{x}_{j-1})\Vert ^2) \\&\le 8S_K + 4n\eta _x^2\sum _{j=0}^{K-1}(\Vert \overline{\partial \Phi (X_{j})} - \nabla \Phi ({{\bar{x}}}_j)\Vert ^2 + \Vert \nabla \Phi ({{\bar{x}}}_j)\Vert ^2) \\&= 8S_K + 4n\eta _x^2\sum _{j=0}^{K-1}\Vert \overline{\partial \Phi (X_{j})} - \nabla \Phi ({{\bar{x}}}_j)\Vert ^2 + 4n\eta _x^2T_{K-1}, \end{aligned}$$

where the second inequality is by the definition of \(Q_j\), the third inequality is by the definition of \(S_K\) and \(Q_0 = 0\), the last equality is by the definition of \(T_{K-1}.\) \(\square\)

Next we give bounds for \(A_K\) and \(B_K\).

Lemma 10

Suppose Assumptions 2.1 and 2.3 hold. If \(\eta _y, T\) and N in Algorithm 3 and 4 satisfy:

$$\begin{aligned} 0<\eta _y< \frac{2}{\mu + L},\quad \delta _y^T< \frac{1}{3},\quad \delta _{\kappa }^N <\frac{1}{8\kappa }, \end{aligned}$$
(14)

then the following inequalities hold:

$$\begin{aligned} A_K\le 3\delta _y^T(c_1 + 2\kappa ^2E_K),\quad B_K\le 2c_2 + 2d_1A_{K-1} + 2d_2E_K, \end{aligned}$$

where the constants are defined as follows:

$$\begin{aligned}&c_1 = \sum _{i=1}^{n}\Vert y_{i,0}^{(0)} - y_i^*(x_{i,0})\Vert ^2,\ c_2 = \sum _{i=1}^{n}\Vert v_{i,0}^* - v_{i,0}^{(0)}\Vert ^2, \nonumber \\&d_1 = 4(1+\sqrt{\kappa })^2\left( \kappa + \frac{L_{g,2} L_{f,0}}{\mu ^2}\right) ^2=\Theta (\kappa ^3), \nonumber \\&d_2 = 2\left( \kappa ^2 + \frac{2L_{f,0}\kappa }{\mu } + \frac{2L_{f,0}\kappa ^2}{\mu }\right) ^2 = \Theta (\kappa ^4). \end{aligned}$$
(15)

Proof

For each term in \(A_K\) we have

$$\begin{aligned} \begin{aligned}\Vert y_{i,j}^{(T)} - y_i^*(x_{i,j})\Vert ^2 &= \Vert y_{i,j}^{(T-1)} - \eta _y\nabla _y g(x_{i,j}, y_{i,j}^{(T-1)}) - y_i^*(x_{i,j}) \Vert ^2 \\& \le (1 - \eta _y\mu )^2\Vert y_{i,j}^{(T-1)} - y_i^*(x_{i,j})\Vert ^2\le \delta _y^T \Vert y_{i,j}^{(0)} - y_i^*(x_{i,j})\Vert ^2, \end{aligned} \end{aligned}$$
(16)

where the first inequality uses Lemma 3. We further have:

$$\begin{aligned} \Vert y_{i,j}^{(0)} - y_i^*(x_{i,j})\Vert ^2 &= \Vert y_{i,j-1}^{(T)} - y_i^*(x_{i,j-1}) + y_i^*(x_{i,j-1}) - y_i^*(x_{i,j})\Vert ^2 \\& \le 2(\Vert y_{i,j-1}^{(T)} - y_i^*(x_{i,j-1})\Vert ^2 + \Vert y_i^*(x_{i,j-1}) - y_i^*(x_{i,j})\Vert ^2) \\& \le 2\delta _y^T \Vert y_{i,j-1}^{(0)} - y_i^*(x_{i,j-1})\Vert ^2 + 2\kappa ^2\Vert x_{i,j-1} - x_{i,j}\Vert ^2 \\& < \frac{2}{3}\Vert y_{i,j-1}^{(0)} - y_i^*(x_{i,j-1})\Vert ^2 + 2\kappa ^2\Vert x_{i,j-1} - x_{i,j}\Vert ^2, \end{aligned}$$

where the second inequality is by (16) and Lemma 7, and the last inequality is by the condition (14). Taking summation on both sides, we get

$$\begin{aligned}&\sum _{j=1}^{K}\sum _{i=1}^{n}\Vert y_{i,j}^{(0)} - y_i^*(x_{i,j})\Vert ^2 \\&\quad \le \frac{2}{3}\sum _{j=1}^{K}\sum _{i=1}^{n}\Vert y_{i,j-1}^{(0)} - y_i^*(x_{i,j-1})\Vert ^2 + 2\kappa ^2\sum _{j=1}^{K}\sum _{i=1}^{n}\Vert x_{i,j} - x_{i,j-1}\Vert ^2 \\&\quad \le \frac{2}{3} \sum _{j=0}^{K}\sum _{i=1}^{n}\Vert y_{i,j}^{(0)} - y_i^*(x_{i,j})\Vert ^2 + 2\kappa ^2E_K \\&\quad \le \frac{2}{3} c_1 + \frac{2}{3}\sum _{j=1}^{K}\sum _{i=1}^{n}\Vert y_{i,j}^{(0)} - y_i^*(x_{i,j})\Vert ^2 + 2\kappa ^2E_K, \end{aligned}$$

which directly implies:

$$\begin{aligned} \sum _{j=1}^{K}\sum _{i=1}^{n}\Vert y_{i,j}^{(0)} - y_i^*(x_{i,j})\Vert ^2 \le 2c_1 + 6\kappa ^2E_K. \end{aligned}$$
(17)

Combining (16) and (17) leads to:

$$\begin{aligned} A_K =&\sum _{j=0}^{K}\sum _{i=1}^{n}\Vert y_{i,j}^{T} - y_i^*(x_{i,j})\Vert ^2\le \delta _y^T \sum _{j=0}^{K}\sum _{i=1}^{n}\Vert y_{i,j}^{(0)} - y_i^*(x_{i,j})\Vert ^2 \\ \le&\delta _y^T (c_1 + 2c_1 + 6\kappa ^2E_K) = 3\delta _y^T(c_1 + 2\kappa ^2E_K). \end{aligned}$$

We then consider the bound for \(B_K\). Recall that:

$$\begin{aligned} v_{i,k}^* = \left( \nabla _y^2g_i(x_{i,k},y_i^*(x_{i,k}))\right) ^{-1}\nabla _y f_i(x_{i,k},y_i^*(x_{i,k})), \end{aligned}$$

which is the solution of the linear system \(\nabla _y^2g_i(x_{i,k},y_i^*(x_{i,k}))v = \nabla _y f_i(x_{i,k},y_i^*(x_{i,k}))\) in the AID-based approach in Algorithm 2. Note that \(v_{i,k}^*\) is a function of \(x_{i,k}\), and it is \((\kappa ^2 + \frac{2L_{f,0}L}{\mu ^2} + \frac{2L_{f,0}L\kappa }{\mu ^2})\)-Lipschitz continuous with respect to \(x_{i,k}\) [13]. For each term in \(B_K\), we have:

$$\begin{aligned}&\Vert v_{i,j}^* - v_{i,j}^{(0)}\Vert ^2 \\&\quad \le 2(\Vert v_{i,j-1}^* - v_{i,j-1}^{(N)}\Vert ^2 + \Vert v_{i,j}^* - v_{i,j-1}^*\Vert ^2 ) \\&\quad \le 4(1+\sqrt{\kappa })^2\left( \kappa + \frac{L_{g, 2} L_{f,0}}{\mu ^2}\right) ^2\Vert y_{i,j-1}^{(T)} - y_i^*(x_{i,j-1})\Vert ^2 \\&\qquad + 4\kappa \left( \frac{\sqrt{\kappa } - 1}{\sqrt{\kappa } + 1}\right) ^{2N}\Vert v_{i,j-1}^* - v_{i,j-1}^{(0)}\Vert ^2 + 2\left( \kappa ^2 + \frac{2L_{f,0}L(1+\kappa )}{\mu ^2}\right) ^2 \Vert x_{i,j} - x_{i,j-1}\Vert ^2, \end{aligned}$$

where the second inequality follows [13, Lemma 4]. Taking summation over ij, we get

$$\begin{aligned} \sum _{j=1}^{K}\sum _{i=1}^{n}\Vert v_{i,j}^* - v_{i,j}^{(0)}\Vert ^2\le d_1 A_{K-1} + 4\kappa \delta _\kappa ^N B_{K-1} + d_2E_K\le d_1 A_{K-1} + \frac{1}{2}B_{K} + d_2E_K, \end{aligned}$$
(18)

where the last inequality holds since we pick N such that \(4\kappa \delta _\kappa ^N <\frac{1}{2}\). Therefore, we can get:

$$\begin{aligned} B_K\le c_2 + d_1 A_{K-1} + \frac{1}{2}B_{K} + d_2E_K\quad \Rightarrow \quad B_K\le 2c_2 + 2d_1A_{K-1} + 2d_2E_K, \end{aligned}$$

which completes the proof. \(\square\)

The following lemmas give bounds on \(\sum \Vert \overline{\partial \Phi (X_k)} - \nabla \Phi ({{\bar{x}}}_k)\Vert ^2\) in (13).We first consider the case when the Assumption 2.3 holds. In this case, the outer loop computes the hypergradient via AID based approach. Therefore, we borrow [13, Lemma 3] and restate it as follows.

Lemma 11

[13, Lemma 3] Suppose Assumptions 2.1 and 2.3 hold, then we have:

$$\begin{aligned}&\Vert {\hat{\nabla }}f_i(x_{i,j}, y_{i,j}^{(T)}) - \nabla f_i(x_{i,j}, y_i^*(x_{i,j})) \Vert ^2\\&\hspace{6em}\le \Gamma \Vert y_i^*(x_{i,j}) - y_{i, j}^{(T)}\Vert ^2 + 6L^2\kappa \left( \frac{\sqrt{\kappa } -1 }{\sqrt{\kappa } + 1}\right) ^{2N}\Vert v_{i,j}^* - v_{i,j}^{(0)}\Vert ^2 \end{aligned}$$

where the constant \(\Gamma\) is

$$\begin{aligned} \Gamma = 3L^2 + \frac{3L_{g,2}^2L_{f,0}}{\mu ^2} + 6L^2(1+\sqrt{\kappa })^2\left( \kappa + \frac{L_{g,2}L_{f,0}}{\mu ^2}\right) ^2=\Theta (\kappa ^3). \end{aligned}$$

Next, we bound \(\sum \Vert \overline{\partial \Phi (X_k)} - \nabla \Phi ({{\bar{x}}}_k)\Vert ^2\) under Assumption 2.3.

Lemma 12

Suppose Assumptions 2.1 and 2.3 hold. We have:

$$\begin{aligned} \begin{aligned}&\sum _{k=0}^{K}\Vert \overline{\partial \Phi (X_k)} - \nabla \Phi (\bar{x}_k)\Vert ^2 \le \frac{2L_{\Phi }^2}{n}S_K + \frac{2\Gamma }{n}A_K + \frac{12L^2\kappa }{n}\delta _\kappa ^NB_K. \end{aligned} \end{aligned}$$
(19)

Proof

Under Assumption 2.3 we know \(g_i = g\), and thus from (5) and (6) we have

$$\begin{aligned} \nabla \Phi _i({{\bar{x}}}_k) = \nabla f_i({{\bar{x}}}_k, y^*({{\bar{x}}}_k)). \end{aligned}$$

Therefore, we have

$$\begin{aligned}&\Vert \overline{\partial \Phi (X_k)} - \nabla \Phi ({{\bar{x}}}_k)\Vert ^2 = \frac{1}{n^2}\left\| \sum _{i=1}^{n}\left( {\hat{\nabla }}f_i(x_{i,k}, y_{i,k}^{(T)}) - \nabla f_i({{\bar{x}}}_k, y^*(\bar{x}_k))\right) \right\| ^2 \\&\quad \le \frac{1}{n}\sum _{i=1}^{n}\Vert {\hat{\nabla }}f_i(x_{i,k}, y_{i,k}^{(T)}) - \nabla f_i({{\bar{x}}}_k, y^*({{\bar{x}}}_k))\Vert ^2 \\&\quad \le \frac{2}{n}\sum _{i=1}^n( \Vert {\hat{\nabla }}f_i(x_{i,k}, y_{i,k}^{(T)}) - \nabla f_i(x_{i,k}, y_i^*(x_{i,k}) )\Vert ^2 \\&\qquad +\, \Vert \nabla f_i(x_{i,k}, y_i^*(x_{i,k})) - \nabla f_i({{\bar{x}}}_k, y^*({{\bar{x}}}_k)) \Vert ^2 ) \\&\quad \le \frac{2}{n}\sum _{i=1}^n (\Gamma \Vert y_i^*(x_{i,k}) - y_{i, k}^{(T)}\Vert ^2 + 6L^2\kappa \delta _\kappa ^N\Vert v_{i,k}^* - v_{i,k}^{(0)}\Vert ^2 + L_{\Phi }^2\Vert x_{i,k} - {{\bar{x}}}_k\Vert ^2) \\&\quad \le \frac{2\Gamma }{n}\sum _{i=1}^n\Vert y_i^*(x_{i,k}) - y_{i, k}^{(T)}\Vert ^2 + \frac{12L^2\kappa }{n}\delta _\kappa ^N\sum _{i=1}^{n}\Vert v_{i,k}^* - v_{i,k}^{(0)}\Vert ^2 + \frac{2L_{\Phi }^2}{n}\Vert Q_k\Vert ^2 , \end{aligned}$$

where the first inequality follows from the convexity of \(\Vert \cdot \Vert ^2\), the third inequality follows from Lemma 11 and Assumption 2.3, the last inequality is by Lemma 5:

$$\begin{aligned} \Vert \nabla f_i(x_{i,k}, y^*(x_{i,k})) - \nabla f_i({{\bar{x}}}_k, y^*(\bar{x}_k)) \Vert ^2=\Vert \nabla \Phi _i(x_{i,k}) - \nabla \Phi _i({{\bar{x}}}_k)\Vert ^2 \le L_{\Phi }^2\Vert q_{i,k}\Vert ^2. \end{aligned}$$

Taking summation on both sides, we get:

$$\begin{aligned} \sum _{k=0}^{K}\Vert \overline{\partial \Phi (X_k)} - \nabla \Phi (\bar{x}_k)\Vert ^2 \le \frac{2L_{\Phi }^2}{n}S_K + \frac{2\Gamma }{n}A_K + \frac{12L^2\kappa }{n}\delta _\kappa ^NB_K. \end{aligned}$$

\(\square\)

We now consider the case when Assumption 2.3 does not hold. In this case, our target in the lower level problem is

$$\begin{aligned} y^*({{\bar{x}}}_k) = \mathop {\mathrm {arg\,min}}\limits _{y}\frac{1}{n}\sum _{i=1}^{n}g_i({{\bar{x}}}_k, y). \end{aligned}$$
(20)

However, the update in our decentralized algorithm (e.g. line 8 of Algorithm 3) aims at solving

$$\begin{aligned} {\tilde{y}}_k^* := \mathop {\mathrm {arg\,min}}\limits _{y}\frac{1}{n}\sum _{i=1}^{n}g_i(x_{i,k}, y), \end{aligned}$$
(21)

which is completely different from our target (20). To resolve this problem, we introduce the following lemma to characterize the difference:

Lemma 13

The following inequality holds:

$$\begin{aligned} \Vert {\tilde{y}}_k^* - y^*({{\bar{x}}}_k)\Vert \le \frac{\kappa }{n}\sum _{i=1}^{n}\Vert x_{i,k} - {{\bar{x}}}_k\Vert \le \frac{\kappa }{\sqrt{n}} \Vert Q_k\Vert . \end{aligned}$$

Proof

By optimality conditions of (20) and (21), we have:

$$\begin{aligned} \frac{1}{n}\sum _{i=1}^{n}\nabla _yg_i(x_{i,k}, {\tilde{y}}_k^*) = 0,\quad \frac{1}{n}\sum _{i=1}^{n}\nabla _yg_i({{\bar{x}}}_k, y^*(\bar{x}_k)) = 0. \end{aligned}$$

Combining with the strongly convexity and the smoothness of \(g_i\) yields:

$$\begin{aligned}&\left\| \frac{1}{n}\sum _{i=1}^{n}\nabla _yg_i({{\bar{x}}}_k, {\tilde{y}}_k^*)\right\| \\&\quad = \left\| \frac{1}{n}\sum _{i=1}^{n}\nabla _yg_i({{\bar{x}}}_k, {\tilde{y}}_k^*) - \frac{1}{n}\sum _{i=1}^{n}\nabla _yg_i({{\bar{x}}}_k, y^*({{\bar{x}}}_k))\right\| \ge \mu \Vert {\tilde{y}}_k^* - y^*({{\bar{x}}}_k)\Vert , \\&\left\| \frac{1}{n}\sum _{i=1}^{n}\nabla _yg_i({{\bar{x}}}_k, {\tilde{y}}_k^*)\right\| \\&\quad = \left\| \frac{1}{n}\sum _{i=1}^{n}\nabla _yg_i({{\bar{x}}}_k, {\tilde{y}}_k^*) - \frac{1}{n}\sum _{i=1}^{n}\nabla _yg_i(x_{i,k}, {\tilde{y}}_k^*) \right\| \le \frac{L}{n}\sum _{i=1}^{n}\Vert x_{i,k} - {{\bar{x}}}_k\Vert . \end{aligned}$$

Therefore, we obtain the following inequality:

$$\begin{aligned} \Vert {\tilde{y}}_k^* - y^*({{\bar{x}}}_k)\Vert \le \frac{\kappa }{n} \sum _{i=1}^{n}\Vert x_{i,k} - {{\bar{x}}}_k\Vert =\frac{\kappa }{n} \sum _{i=1}^{n}\Vert q_{i,k}\Vert \le \frac{\kappa }{\sqrt{n}} \Vert Q_k\Vert , \end{aligned}$$

where the last inequality is by Cauchy–Schwarz inequality. \(\square\)

Notice that in the inner loop of Algorithms 3, 4 and 5, i.e., Lines 4–11 of Algorithms 3 and 4, and Lines 4–10 of Algorithm 5, \(y_{i,k}^{(T)}\) converges to \({\tilde{y}}_k^*\) and the rates are characterized by [34, 36, 46, 55] (e.g., Corollary 4.7. in [55], Theorem 10 in [35] and Theorem 1 in [46]). We include all the convergence rates here.

Lemma 14

Suppose Assumption 2.3 does not hold. We have:

  • In Algorithm 3 and 4 there exists a constant \(\eta _y\) such that

    $$\begin{aligned} \frac{1}{n}\sum _{i=1}^{n}\Vert y_{i,k}^{(T)} - {\tilde{y}}_k^*\Vert ^2\le C_1 \alpha _1^T. \end{aligned}$$
  • In Algorithm 5 there exists \(\eta _y^{(t)} = {\mathcal {O}}(\frac{1}{t})\) such that

    $$\begin{aligned} \frac{1}{n}\sum _{i=1}^{n}{\mathbb {E}}\left[ \Vert y_{i,k}^{(T)} - {\tilde{y}}_k^*\Vert ^2\right] \le \frac{C_2}{T}. \end{aligned}$$

Here \(C_1, C_2\) are positive constants and \(\alpha _1\in (0,1)\).

Besides, the JHIP oracle (Algorithm 1) also performs standard decentralized optimization with gradient tracking in deterministic case (Algorithms 3, 4) and stochastic case (Algorithm 5). We have:

Lemma 15

In Algorithm 1, we have:

  • For deterministic case, there exists a constant \(\gamma\) such that if \(\gamma _t \equiv \gamma\) then

    $$\begin{aligned} \Vert Z_i^{(t)}-Z^*\Vert ^2 \le C_3\alpha _2^t.\quad \text {(See }[34]\text {).} \end{aligned}$$
  • For stochastic case and there exists a diminishing stepsize sequence \(\gamma _t = {\mathcal {O}}(\frac{1}{t})\), such that

    $$\begin{aligned} \frac{1}{n}\sum _{i=1}^{n}{\mathbb {E}}\left[ \Vert Z_i^{(t)}-Z^*\Vert ^2\right] \le \frac{C_4}{t}.\quad \text {(See }[36]\text {)}. \end{aligned}$$

Here \(C_3, C_4\) are positive constant, and \(\alpha _2\in (0,1)\). Here the optimal solution is denoted by \((Z^*)^{{\textsf{T}}}= \left( \sum _{i=1}^{n}J_i\right) \left( \sum _{i=1}^{n}H_i\right) ^{-1}\).

For simplicity we define:

$$\begin{aligned} C = \max \left( C_1,C_2,C_3,C_4\right) ,\quad \alpha = \max \left( \alpha _1,\alpha _2\right) . \end{aligned}$$

Since the objective functions mentioned in Lemma 14 (the lower level function g) and 15 (the objective in (9)) are strongly convex, we know C and \(\alpha\) only depend on \(L, \mu , \rho\) and the stepsize (only when it is a constant). For example \(\alpha _2\) in Lemma 15 only depends on the spectral radius of \(H_i\), smallest eigenvalue of \(H_i\), \(\rho\) and \(\gamma\).

For heterogeneous data (i.e., no Assumption 2.3) on g we have a different error estimation. We first notice that for each JHIP oracle, the following lemma holds:

Lemma 16

Suppose Assumptions 2.1 holds. In Algorithm 3 and 4 we have:

$$\begin{aligned}&\Vert \left( Z_{k}^*\right) ^{{\textsf{T}}}- \nabla _{xy} g({{\bar{x}}}_k, {\tilde{y}}_k^*)\left( \nabla _y^2g({{\bar{x}}}_k, {\tilde{y}}_k^*)\right) ^{-1}\Vert _2^2\\&\hspace{10em}\le \frac{2L_{g,2}^2(1 + \kappa ^2)}{\mu ^2}\left( \frac{1}{n}\Vert Q_k\Vert ^2 + \frac{1}{n}\sum _{j=1}^{n}\Vert y_{j,k}^{(T)} - {\tilde{y}}_k^*\Vert ^2\right) , \end{aligned}$$

where \(Z_{k}^*\) denotes the optimal solution of Algorithm 1 in iteration k:

$$\begin{aligned} \left( Z_{k}^*\right) ^{{\textsf{T}}}= \left( \frac{1}{n}\sum _{j=1}^{n}\nabla _{xy} g_j(x_{j,k}, y_{j,k}^{(T)})\right) \left( \frac{1}{n}\sum _{j=1}^{n}\nabla _y^2g_j(x_{j,k}, y_{j,k}^{(T)})\right) ^{-1}. \end{aligned}$$

Proof

Notice that we have

$$\begin{aligned}&\Vert \left( Z_{k}^*\right) ^{{\textsf{T}}}- \nabla _{xy} g({{\bar{x}}}_k, {\tilde{y}}_k^*)\left[ \nabla _y^2g({{\bar{x}}}_k, {\tilde{y}}_k^*)\right] ^{-1}\Vert _2^2 \\&\quad \le 2\left\| \left( \frac{1}{n}\sum _{j=1}^{n}\nabla _{xy} g_j(x_{j,k}, y_{j,k}^{(T)}) - \nabla _{xy} g({{\bar{x}}}_k, {\tilde{y}}_k^*)\right) \left( \frac{1}{n}\sum _{j=1}^{n}\nabla _y^2g_j(x_{j,k}, y_{j,k}^{(T)})\right) ^{-1}\right\| _2^2 \\&\qquad +\,2\left\| \nabla _{xy} g({{\bar{x}}}_k, {\tilde{y}}_k^*)\left[ \left( \frac{1}{n}\sum _{j=1}^{n}\nabla _y^2g_j(x_{j,k}, y_{j,k}^{(T)})\right) ^{-1} -\left( \nabla _y^2g({{\bar{x}}}_k, {\tilde{y}}_k^*)\right) ^{-1}\right] \right\| _2^2 \\&\quad \le \frac{2L_{g,2}^2}{n\mu ^2}\sum _{j=1}^{n}(\Vert x_{j,k} - {{\bar{x}}}_k\Vert ^2 + \Vert y_{j,k}^{(T)} - {\tilde{y}}_k^*\Vert ^2) \\&\qquad +\, \frac{2L_{g,1}^2L_{g,2}^2}{n\mu ^4}\sum _{j=1}^{n}(\Vert x_{j,k} - {{\bar{x}}}_k\Vert ^2 + \Vert y_{j,k}^{(T)} - {\tilde{y}}_k^*\Vert ^2) \\&\quad \le \frac{2L_{g,2}^2(1 + \kappa ^2)}{\mu ^2}\left( \frac{1}{n}\Vert Q_k\Vert ^2 + \frac{1}{n}\sum _{j=1}^{n}\Vert y_{j,k}^{(T)} - {\tilde{y}}_k^*\Vert ^2\right) \end{aligned}$$

where the second inequality holds due to Assumption 2.1 and the following inequality:

$$\begin{aligned}&\left\| \left( \frac{1}{n}\sum _{j=1}^{n}\nabla _y^2g_j(x_{j,k}, y_{j,k}^{(T)})\right) ^{-1} -\left( \nabla _y^2g({{\bar{x}}}_k, {\tilde{y}}_k^*)\right) ^{-1}\right\| _2^2 \\&\quad =\bigg \Vert \bigg (\frac{1}{n}\sum _{j=1}^{n}\nabla _y^2g_j(x_{j,k}, y_{j,k}^{(T)})\bigg )^{-1}\cdot \\&\qquad \bigg (\nabla _y^2g({{\bar{x}}}_k, {\tilde{y}}_k^*) - \frac{1}{n}\sum _{j=1}^{n}\nabla _y^2g_j(x_{j,k}, y_{j,k}^{(T)})\bigg )\left( \nabla _y^2g({{\bar{x}}}_k, {\tilde{y}}_k^*)\right) ^{-1}\bigg \Vert _2^2 \\&\quad \le \frac{L_{g,2}^2}{n\mu ^4}\sum _{j=1}^{n}(\Vert x_{j,k} - \bar{x}_k\Vert ^2 + \Vert y_{j,k}^{(T)} - {\tilde{y}}_k^*\Vert ^2). \end{aligned}$$

\(\square\)

Lemma 17

Suppose Assumption 2.1 holds. In Algorithm 3 and 4 we have:

$$\begin{aligned}&\frac{1}{n}\sum _{i=1}^{n}\Vert {\hat{\nabla }}f_i(x_{i,k}, y_{i,k}^{(T)}) - {\bar{\nabla }} f_i(x_{i,k}, {\tilde{y}}_k^*)\Vert ^2 \nonumber \\&\quad \le \frac{18L_{f,0}^2L_{g,2}^2(1 + \kappa ^2)}{n\mu ^2}\Vert Q_k\Vert ^2 + \frac{6L_{f,0}^2}{n}\sum _{i=1}^{n}\Vert Z_{i,k}^{(N)} - Z_k^*\Vert ^2 \nonumber \\&\quad +\, \left( 6 + 6L^2\kappa ^2 + \frac{12L_{f,0}^2L_{g,2}^2(1+\kappa ^2)}{\mu ^2}\right) \cdot \left( \frac{1}{n}\sum _{i=1}^{n}\Vert y_{i,k}^{(T)} - \tilde{y}_k^*\Vert ^2\right) . \end{aligned}$$
(22)

Proof

Note that

$$\begin{aligned} {\hat{\nabla }}f_i(x_{i,k}, y_{i,k}^{(T)})&= \nabla _x f_i(x_{i,k},y_{i,k}^{(T)}) -\left( Z_{i,k}^{(N)}\right) ^{{\textsf{T}}}\nabla _y f_i(x_{i,k},y_{i,k}^{(T)}), \\ {\bar{\nabla }} f_i(x_{i,k}, {\tilde{y}}_k^*)&= \nabla _x f_i(x_{i,k}, {\tilde{y}}_k^*) - \nabla _{xy} g(x_{i,k}, {\tilde{y}}_k^*)\nabla _y^2g(x_{i,k}, {\tilde{y}}_k^*)^{-1}\nabla _yf_i(x_{i,k},{\tilde{y}}_k^*). \end{aligned}$$

Then we know

$$\begin{aligned}&{\hat{\nabla }}f_i(x_{i,k}, y_{i,k}^{(T)}) - {\bar{\nabla }} f_i(x_{i,k}, {\tilde{y}}_k^*)\\&\quad = \nabla _x f_i(x_{i,k},y_{i,k}^{(T)}) - \nabla _x f_i(x_{i,k}, {\tilde{y}}_k^*) \\&\qquad - \left( Z_{i,k}^{(N)}\right) ^{{\textsf{T}}}\nabla _y f_i(x_{i,k},y_{i,k}^{(T)}) + \left( Z_{k}^*\right) ^{{\textsf{T}}}\nabla _y f_i(x_{i,k},y_{i,k}^{(T)}) \\&\qquad - \left( Z_{k}^*\right) ^{{\textsf{T}}}\nabla _y f_i(x_{i,k},y_{i,k}^{(T)}) + \left( Z_{k}^*\right) ^{{\textsf{T}}}\nabla _y f_i(x_{i,k},{\tilde{y}}_k^*) \\&\qquad - \left( Z_{k}^*\right) ^{{\textsf{T}}}\nabla _y f_i(x_{i,k},{\tilde{y}}_k^*) + \nabla _{xy} g({{\bar{x}}}_k,{\tilde{y}}_k^*)\nabla _y^2g({{\bar{x}}}_k, {\tilde{y}}_k^*)^{-1}\nabla _y f_i(x_{i,k}, {\tilde{y}}_k^*) \\&\qquad - \nabla _{xy} g({{\bar{x}}}_k,{\tilde{y}}_k^*)\nabla _y^2g({{\bar{x}}}_k, {\tilde{y}}_k^*)^{-1}\nabla _y f_i(x_{i,k}, {\tilde{y}}_k^*) \\&\qquad + \nabla _{xy} g(x_{i,k}, {\tilde{y}}_k^*)\nabla _y^2g({{\bar{x}}}_k, {\tilde{y}}_k^*)^{-1}\nabla _y f_i(x_{i,k},{\tilde{y}}_k^*) \\&\qquad -\nabla _{xy} g(x_{i,k}, {\tilde{y}}_k^*)\nabla _y^2g({{\bar{x}}}_k, {\tilde{y}}_k^*)^{-1}\nabla _y f_i(x_{i,k},{\tilde{y}}_k^*) \\&\qquad + \nabla _{xy} g(x_{i,k}, {\tilde{y}}_k^*)\nabla _y^2g(x_{i,k}, {\tilde{y}}_k^*)^{-1}\nabla _y f(x_{i,k},{\tilde{y}}_k^*), \end{aligned}$$

which gives

$$\begin{aligned}&\Vert {\hat{\nabla }}f_i(x_{i,k}, y_{i,k}^{(T)}) - {\bar{\nabla }} f_i(x_{i,k}, {\tilde{y}}_k^*)\Vert ^2 \\&\quad \le 6 \left(\Vert y_{i,k}^{(T)} - {\tilde{y}}_k^*\Vert ^2 + L_{f,0}^2\Vert Z_{i,k}^{(N)} - Z_{k}^*\Vert ^2 + L^2\Vert \left( Z_{k}^*\right) ^{{\textsf{T}}}\Vert _2^2\Vert y_{i,k}^{(T)} - {\tilde{y}}_k^*\Vert ^2 \right. \\&\left. \qquad +L_{f,0}^2\Vert \left( Z_{k}^*\right) ^{{\textsf{T}}}- \nabla _{xy} g({{\bar{x}}}_k, {{\tilde{y}}}_k^*)\nabla _y^2g({{\bar{x}}}_k, {\tilde{y}}_k^*)^{-1}\Vert _2^2 + \frac{L_{g,2}^2L_{f,0}^2}{\mu ^2}\Vert x_{i,k} - {{\bar{x}}}_k\Vert ^2 \right. \\&\left. \qquad + L^2L_{f,0}^2\Vert \nabla _y^2g({{\bar{x}}}_k, {\tilde{y}}_k^*)^{-1}- \nabla _y^2g(x_{i,k}, {\tilde{y}}_k^*)^{-1}\Vert _2^2 \right) \\&\quad \le 6 \left(\Vert y_{i,k}^{(T)} - {\tilde{y}}_k^*\Vert ^2 + L_{f,0}^2\Vert Z_{i,k}^{(N)} - Z_{k}^*\Vert ^2 + \frac{L^4}{\mu ^2}\Vert y_{i,k}^{(T)} - {\tilde{y}}_k^*\Vert ^2 \right. \\& \left. \qquad + \frac{2L_{f,0}^2L_{g,2}^2(1 + \kappa ^2)}{\mu ^2}\left( \frac{1}{n}\Vert Q_k\Vert ^2 + \frac{1}{n}\sum _{j=1}^{n}\Vert y_{j,k}^{(T)} - {\tilde{y}}_k^*\Vert ^2\right)\right. \\&\left. \qquad + \frac{L_{g,2}^2L_{f,0}^2}{\mu ^2}\Vert x_{i,k}-{{\bar{x}}}_k\Vert ^2 + \frac{L^2L_{f,0}^2L_{g,2}^2}{\mu ^4}\Vert x_{i,k} - {{\bar{x}}}_k\Vert ^2 \right). \end{aligned}$$

The second inequality uses Lemma 16, Assumption 2.1. Taking summation on both sides and using Lemma 15, we know

$$\begin{aligned}&\frac{1}{n}\sum _{i=1}^{n}\Vert {\hat{\nabla }}f_i(x_{i,k}, y_{i,k}^{(T)}) - {\bar{\nabla }} f_i(x_{i,k}, {\tilde{y}}_k^*)\Vert ^2 \\&\quad \le \frac{18L_{f,0}^2L_{g,2}^2(1 + \kappa ^2)}{n\mu ^2}\Vert Q_k\Vert ^2 + \frac{6L_{f,0}^2}{n}\sum _{i=1}^{n}\Vert Z_{i,k}^{(N)} - Z_k^*\Vert ^2 \\&\qquad + \left( 6 + 6L^2\kappa ^2 + \frac{12L_{f,0}^2L_{g,2}^2(1+\kappa ^2)}{\mu ^2}\right) \cdot \left( \frac{1}{n}\sum _{i=1}^{n}\Vert y_{i,k}^{(T)} - \tilde{y}_k^*\Vert ^2\right) . \end{aligned}$$

\(\square\)

Lemma 18

Suppose Assumption 2.3 does not hold, then in Algorithms 3 and 4 we have:

$$\begin{aligned} \begin{aligned}\Vert \overline{\partial \Phi (X_k)} - \nabla \Phi ({{\bar{x}}}_k)\Vert ^2 &\le \frac{(1+\kappa ^2)}{n} \cdot \left( \frac{36L_{f,0}^2L_{g,2}^2}{\mu ^2} + 2L_{f}^2\right) \Vert Q_k\Vert ^2 \\&\quad +12C\left[ \left( 1 + L^2\kappa ^2 + \frac{2L_{f,0}^2L_{g,2}^2(1+\kappa ^2)}{\mu ^2}\right) \alpha ^T + L_{f,0}^2\alpha ^N\right] . \end{aligned} \end{aligned}$$
(23)

Proof

We have

$$\begin{aligned}&\Vert \overline{\partial \Phi (X_k)} - \nabla \Phi ({{\bar{x}}}_k)\Vert ^2 = \frac{1}{n^2}\left\| \sum _{i=1}^{n}\left( {\hat{\nabla }}f_i(x_{i,k}, y_{i,k}^{(T)}) - \nabla f_i({{\bar{x}}}_k, y^*(\bar{x}_k))\right) \right\| ^2 \nonumber \\&\quad \le \frac{1}{n}\sum _{i=1}^{n}\Vert {\hat{\nabla }}f_i(x_{i,k}, y_{i,k}^{(T)}) - \nabla f_i({{\bar{x}}}_k, y^*({{\bar{x}}}_k))\Vert ^2 \nonumber \\&\quad \le \frac{2}{n}\sum _{i=1}^n( \Vert {\hat{\nabla }}f_i(x_{i,k}, y_{i,k}^{(T)}) - {\bar{\nabla }} f_i(x_{i,k}, {\tilde{y}}_k^*))\Vert ^2 + \Vert {\bar{\nabla }} f_i(x_{i,k}, {\tilde{y}}_k^*)- \nabla f_i({{\bar{x}}}_k, y^*({{\bar{x}}}_k)) \Vert ^2) \nonumber \\&\quad \le \frac{36L_{f,0}^2L_{g,2}^2(1 + \kappa ^2)}{n\mu ^2}\Vert Q_k\Vert ^2 \nonumber \\&\qquad + 12\left( 1 + L^2\kappa ^2 + \frac{2L_{f,0}^2L_{g,2}^2(1+\kappa ^2)}{\mu ^2}\right) \cdot \frac{1}{n}\sum _{i=1}^{n}\Vert y_{i,k}^{(T)} - {{\tilde{y}}}_k^*\Vert ^2 \nonumber \\&\qquad + \frac{12L_{f,0}^2}{n}\sum _{i=1}^{n}\Vert Z_{i,k}^{(N)} - Z_k^*\Vert ^2 + \frac{2}{n}\sum _{i=1}^{n}(L_{f}^2\Vert x_{i,k} - \bar{x}_k\Vert ^2 + L_{f}^2\Vert {\tilde{y}}_k^* - y^*({{\bar{x}}}_k)\Vert ^2) \nonumber \\&\quad \le 12\left( 1 + L^2\kappa ^2 + \frac{2L_{f,0}^2L_{g,2}^2(1+\kappa ^2)}{\mu ^2}\right) \cdot \frac{1}{n}\sum _{i=1}^{n}\Vert y_{i,k}^{(T)} - {{\tilde{y}}}_k^*\Vert ^2 \nonumber \\&\qquad + \frac{12L_{f,0}^2}{n}\sum _{i=1}^{n}\Vert Z_{i,k}^{(N)} - Z_k^*\Vert ^2 + \frac{(1+\kappa ^2)}{n}\cdot \left( \frac{36L_{f,0}^2L_{g,2}^2}{\mu ^2} + 2L_{f}^2\right) \Vert Q_k\Vert ^2, \end{aligned}$$
(24)

where the third inequality is due to Lemma 17 and Lemma 6, and the fourth inequality is by Lemma 13. Notice that \(\frac{1}{n}\sum _{i=1}^{n}\Vert y_{i,k}^{(T)} - {\tilde{y}}_k^*\Vert ^2\) in the first term denotes the error of the inner loop iterates. In both DBO (Algorithm 3) and DBOGT (Algorithm 4), the inner loop performs a decentralized gradient descent with gradient tracking. By Lemmas 14 and 15, we have the error bounds \(\frac{1}{n}\sum _{i=1}^{n}\Vert y_{i,k}^{(T)} - {\tilde{y}}_k^*\Vert ^2\le C\alpha ^T\) and \(\frac{1}{n}\sum _{i=1}^{n}\Vert Z_{i,k}^{(N)} - Z_k^*\Vert ^2\le C\alpha ^N\), which complete the proof. \(\square\)

1.1 Proof of the DBO convergence

In this section we will prove the following convergence result of the DBO algorithm:

Theorem 19

In Algorithm 3, suppose Assumptions 2.1 and 2.2 hold. If Assumption 2.3 holds, then by setting \(0<\eta _x\le \frac{1-\rho }{130L_{\Phi }},\ 0<\eta _y< \frac{2}{\mu + L},\ T=\Theta (\kappa \log \kappa ),\ N=\Theta (\sqrt{\kappa }\log \kappa )\), we have:

$$\begin{aligned}&\frac{1}{K+1}\sum _{j=0}^{K}\Vert \nabla \Phi ({{\bar{x}}}_j)\Vert ^2 \le \frac{4}{\eta _x(K+1)}(\Phi ({{\bar{x}}}_0)-\inf _x \Phi (x))+ \eta _x^2\cdot \frac{1272L_{\Phi }^2L_{f,0}^2(1+\kappa )^2}{(1-\rho )^2} + \frac{C_1}{K+1}. \end{aligned}$$

If Assumption 2.3 does not hold, then by setting \(0< \eta _x\le \frac{1}{L_{\Phi }}, \eta _y^{(t)} = {\mathcal {O}}(\frac{1}{t})\), we have:

$$\begin{aligned}\frac{1}{K+1}\sum _{j=0}^{K}\Vert \nabla \Phi (\bar{x}_j)\Vert ^2 &\le \frac{2}{\eta _x(K+1)}(\Phi ({{\bar{x}}}_0) - \inf _x\Phi (x)) \\&\quad + \eta _x^2\left( \frac{18L_{f,0}^2L_{g,2}^2}{\mu ^2} + L_{f}^2\right) \frac{4(1+\kappa ^2)((1+\kappa )^2 + C\alpha ^N)L_{f,0}^2 }{(1-\rho )^2}+ \tilde{C_1}, \end{aligned}$$

where \(C_1 = \Theta (1), C = \Theta (1)\) and \(\tilde{C_1} = {\mathcal {O}}(\alpha ^T + \alpha ^N)\).

We first consider bounding the consensus error estimation for DBO:

Lemma 20

In Algorithm 3, we have

$$\begin{aligned} S_{K}:= \sum _{k=1}^{K}\Vert Q_k\Vert ^2< \frac{\eta _x^2}{(1-\rho )^2}\sum _{j=0}^{K-1}\sum _{i=1}^{n}\Vert {\hat{\nabla }}f_i(x_{i,j}, y_{i,j}^{(T)})\Vert ^2. \end{aligned}$$

Proof

Note that the x update can be written as

$$\begin{aligned} X_k = X_{k-1}W - \eta _x\partial \Phi (X_{k-1}), \end{aligned}$$

which indicates

$$\begin{aligned} {{\bar{x}}}_k = {{\bar{x}}}_{k-1} - \eta _x\overline{\partial \Phi (X_{k-1})}. \end{aligned}$$

By definition of \(q_{i,k}\), we have

$$\begin{aligned} q_{i,k+1}&= \sum _{j=1}^{n}w_{ij}x_{j,k} - \eta _x {\hat{\nabla }}f(x_{i,k}, y_{i,k}^{(T)}) - ({{\bar{x}}}_k - \eta _x\overline{\partial \Phi (X_k)} ) \\&=\sum _{j=1}^{n}w_{ij}(x_{j,k} - {{\bar{x}}}_k) - \eta _x({\hat{\nabla }}f(x_{i,k}, y_{i,k}^{(T)}) - \overline{\partial \Phi (X_k)}) \\&= Q_kWe_i - \eta _x\partial \Phi (X_k)\left( e_i - \frac{{\textbf{1}}_n}{n} \right) , \end{aligned}$$

where the last equality uses the fact that W is symmetric. Therefore, for \(Q_{k+1}\) we have

$$\begin{aligned} Q_{k+1}&= Q_kW - \eta _x\partial \Phi (X_k)\left( I - \frac{{\textbf{1}}_n{\textbf{1}}^{\top }_n}{n}\right) \\&=\left( Q_{k-1}W - \eta _x\partial \Phi (X_{k-1})\left( I - \frac{{\textbf{1}}_n{\textbf{1}}^{\top }_n}{n}\right) \right) W - \eta _x\partial \Phi (X_k)\left( I - \frac{{\textbf{1}}_n{\textbf{1}}^{\top }_n}{n}\right) \\&=Q_0W^{k+1} - \eta _x\sum _{i=0}^{k}\left( \partial \Phi (X_{i})\left( I - \frac{{\textbf{1}}_n{\textbf{1}}^{\top }_n}{n}\right) W^{k-i}\right) \\&=-\eta _x\sum _{i=0}^{k}\partial \Phi (X_{i})\left( W^{k-i} - \frac{{\textbf{1}}_n{\textbf{1}}^{\top }_n}{n}\right) , \end{aligned}$$

where the last equality is obtained by \(Q_0 = 0\) and \({\textbf{1}}_n{\textbf{1}}^{\top }_n W = {\textbf{1}}_n{\textbf{1}}^{\top }_n.\) By Cauchy–Schwarz inequality, we have the following estimate

$$\begin{aligned}&\Vert Q_{k+1}\Vert ^2 = \eta _x^2 \Vert \sum _{i=0}^{k} \partial \Phi (X_{i})\left( W^{k-i} - \frac{{\textbf{1}}_n{\textbf{1}}^{\top }_n}{n}\right) \Vert ^2\\&\quad \le \eta _x^2 \left( \sum _{i=0}^{k} \Vert \partial \Phi (X_{i})\left( W^{k-i} - \frac{{\textbf{1}}_n{\textbf{1}}^{\top }_n}{n}\right) \Vert \right) ^2 \\&\quad \le \eta _x^2 \left( \sum _{i=0}^{k} \Vert \partial \Phi (X_{i})\Vert \Vert \left( W^{k-i} - \frac{{\textbf{1}}_n{\textbf{1}}^{\top }_n}{n}\right) \Vert _2\right) ^2 \\&\quad \le \eta _x^2 \left( \sum _{i=0}^{k}\rho ^{k-i}\Vert \partial \Phi (X_i)\Vert ^2 \right) \left( \sum _{i=0}^{k}\frac{1}{\rho ^{k-i}}\left\| W^{k-i} - \frac{{\textbf{1}}_n{\textbf{1}}^{\top }_n}{n}\right\| _2^2\right) \\&\quad \le \eta _x^2 \left( \sum _{i=0s}^{k}\rho ^{k-i}\Vert \partial \Phi (X_i)\Vert ^2\right) \left( \sum _{i=0}^{k}\rho ^{k-i}\right) < \frac{\eta _x^2}{1-\rho }\left( \sum _{i=0}^{k}\rho ^{k-i}\Vert \partial \Phi (X_i)\Vert ^2\right) \\&\quad = \frac{\eta _x^2}{1-\rho }\left( \sum _{j=0}^{k}\rho ^{k-j}\sum _{i=1}^{n}\Vert {\hat{\nabla }}f_i(x_{i,j}, y_{i,j}^{(T)})\Vert ^2\right) = \frac{\eta _x^2}{1-\rho }\sum _{i=1}^{n}\sum _{j=0}^{k}\rho ^{k-j}\Vert {\hat{\nabla }}f_i(x_{i,j}, y_{i,j}^{(T)})\Vert ^2. \end{aligned}$$

where the fourth inequality is obtained by Lemma 4. Summing the above inequality yields

$$\begin{aligned} S_K&= \sum _{k=0}^{K-1}\Vert Q_{k+1}\Vert ^2< \frac{\eta _x^2}{1-\rho }\sum _{k=0}^{K-1}\sum _{i=1}^{n}\sum _{j=0}^k\rho ^{k-j}\Vert {\hat{\nabla }}f_i(x_{i,j}, y_{i,j}^{(T)})\Vert ^2 \nonumber \\&=\frac{\eta _x^2}{1-\rho }\sum _{j=0}^{K-1}\sum _{i=1}^{n}\sum _{k=j}^{K-1}\rho ^{k-j}\Vert {\hat{\nabla }}f_i(x_{i,j}, y_{i,j}^{(T)})\Vert ^2\nonumber \\&< \frac{\eta _x^2}{(1-\rho )^2}\sum _{j=0}^{K-1}\sum _{i=1}^{n}\Vert {\hat{\nabla }}f_i(x_{i,j}, y_{i,j}^{(T)})\Vert ^2, \end{aligned}$$
(25)

where the second equality holds since we can change the order of summation. \(\square\)

1.1.1 Case 1: Assumption 2.3 holds

We first consider the case when Assumption 2.3 holds.

Lemma 21

Suppose Assumptions 2.1 and 2.3 hold, then we have:

$$\begin{aligned} \Vert {\hat{\nabla }}f_i(x_{i,j}, y_{i,j}^{(T)})\Vert ^2\le 2(L^2\kappa \delta _\kappa ^N\Vert v_{i,j}^{(0)}- v_{i,j}^*\Vert ^2 + (1+\kappa )^2L_{f,0}^2). \end{aligned}$$

Proof

Notice that we have:

$$\begin{aligned}&\Vert {\hat{\nabla }}f_i(x_{i,j}, y_{i,j}^{(T)})\Vert ^2 \le 2\Vert {\hat{\nabla }}f_i(x_{i,j}, y_{i,j}^{(T)}) - {\bar{\nabla }} f_i(x_{i,j}, y_{i,j}^{(T)})\Vert ^2 + 2\Vert {\bar{\nabla }} f_i(x_{i,j}, y_{i,j}^{(T)})\Vert ^2 \\&\quad \le 2\Vert \nabla _{xy} g_i(x_{i,j},y_{i,j}^{(T)})(v_{i,j}^{(N)} - v_{i,j}^*)\Vert ^2 \\&\qquad + 2\Vert \nabla _x f_i(x_{i,j}, y_{i,j}^{(T)}) - \nabla _{xy} g_i(x_{i,j}, y_{i,j}^{(T)}) \left( \nabla _y^2g_i(x_{i,j}, y_{i,j}^{(T)})\right) ^{-1}\nabla _y f_i(x_{i,j}, y_{i,j}^{(T)})\Vert ^2 \\&\quad \le 2(L^2\Vert v_{i,j}^{(N)} - v_{i,j}^*\Vert ^2 + (L_{f,0} + \frac{L}{\mu }L_{f,0})^2)\le 2(L^2\kappa \delta _\kappa ^N\Vert v_{i,j}^{(0)}- v_{i,j}^*\Vert ^2 + (1+\kappa )^2L_{f,0}^2), \end{aligned}$$

where the second inequality is via the Assumption 2.1, and the last inequality is based on the convergence result of CG for the quadratic programming, e.g., eq. (17) in [24]. \(\square\)

Next we obtain the upper bound for \(S_K\).

Lemma 22

Suppose Assumptions 2.1 and 2.3 hold, then we have:

$$\begin{aligned} S_K < \frac{2\eta _x^2}{(1-\rho )^2} (L^2\kappa \delta _\kappa ^NB_{K-1} + nK(1+\kappa )^2L_{f,0}^2). \end{aligned}$$

Proof

By Lemmas 20 and 21, we have:

$$\begin{aligned} S_K <&\frac{\eta _x^2}{(1-\rho )^2}\sum _{j=0}^{K-1}\sum _{i=1}^{n}\Vert {\hat{\nabla }}f_i(x_{i,j}, y_{i,j}^{(T)})\Vert ^2 \\ \le&\frac{\eta _x^2}{(1-\rho )^2}\sum _{j=0}^{K-1}\sum _{i=1}^{n}2(L^2\kappa \delta _\kappa ^N\Vert v_{i,j}^{(0)}- v_{i,j}^*\Vert ^2 + (1+\kappa )^2L_{f,0}^2) \\ =&\frac{2\eta _x^2}{(1-\rho )^2} (L^2\kappa \delta _\kappa ^NB_{K-1} + nK(1+\kappa )^2L_{f,0}^2), \end{aligned}$$

which completes the proof. \(\square\)

We are ready to prove the main results in Theorem 19. We first summarize main results in Lemmas 22, 10 and 9:

$$\begin{aligned} S_{K}&< \frac{2\eta _x^2}{(1-\rho )^2} (L^2\kappa \delta _\kappa ^NB_{K-1} + nK(1+\kappa )^2L_{f,0}^2), \nonumber \\ A_K&\le 3\delta _y^T(c_1 + 2\kappa ^2E_K),\ B_K\le 2c_2 + 2d_1A_{K-1} + 2d_2E_K, \nonumber \\ E_K&\le 8S_K + 4n\eta _x^2\sum _{j=0}^{K-1}\Vert \overline{\partial \Phi (X_{j})} - \nabla \Phi ({{\bar{x}}}_j)\Vert ^2 + 4n\eta _x^2T_{K-1}. \end{aligned}$$
(26)

The next lemma proves the first part of Theorem 19.

Lemma 23

Suppose the assumptions of Lemma 10 hold. Furthermore, if we set \(N = \Theta (\sqrt{\kappa }\log \kappa ), T=\Theta (\kappa \log \kappa ), \eta _x = {\mathcal {O}}(\kappa ^{-3})\) such that:

$$\begin{aligned}&\delta _{\kappa }^N<\min \left( \frac{L_{\Phi }^2}{L^2\kappa (4d_1\kappa ^2 + 2d_2)}, \kappa ^{-6}\right) =\Theta (\kappa ^{-6}),\\&\delta _y^T<\min \left( \frac{L_{\Phi }^2}{12\Gamma \kappa ^2}, \kappa ^{-5}, \frac{1}{3}\right) =\Theta (\kappa ^{-5}),\ \eta _x< \frac{1-\rho }{130L_{\Phi }}, \end{aligned}$$

we have:

$$\begin{aligned}&\frac{1}{K+1}\sum _{j=0}^{K}\Vert \nabla \Phi ({{\bar{x}}}_j)\Vert ^2 \le \frac{4}{\eta _x(K+1)}(\Phi ({{\bar{x}}}_0)-\inf _x \Phi (x)) + \eta _x^2\cdot \frac{1272L_{\Phi }^2L_{f,0}^2(1+\kappa )^2}{(1-\rho )^2} + \frac{C_1}{K+1}, \end{aligned}$$

where the constant is given by:

$$\begin{aligned} C_1&= 106L_{\Phi }^2\cdot \frac{6\eta _x^2}{(1-\rho )^2}L^2\kappa \delta _{\kappa }^N(2c_2 + 2d_1c_1) +\frac{18L^2\kappa \delta _{\kappa }^N(2c_2 + 2d_1c_1) + 9\Gamma c_1\delta _y^T}{n}\\&= \Theta (\eta _x^2\delta _{\kappa }^{N}\kappa ^{12} + \kappa ^5\delta _y^T) = \Theta (1). \end{aligned}$$

Proof

For \(B_K\) we know:

$$\begin{aligned} B_K&\le 2c_2 + 2d_1A_K + 2d_2E_K \le 2c_2 + \frac{2}{3}d_1(3c_1 + 6\kappa ^2E_K) + 2d_2E_K \nonumber \\&= 2c_2 + 2d_1c_1 + (4d_1\kappa ^2 + 2d_2)E_K. \end{aligned}$$
(27)

We first eliminate \(B_K\) in the upper bound of \(S_K\). Pick NT such that:

$$\begin{aligned} \delta _{\kappa }^N\cdot (4d_1\kappa ^2 + 2d_2)\cdot L^2\kappa<L_{\Phi }^2\quad \Rightarrow \quad \delta _{\kappa }^N< \frac{L_{\Phi }^2}{L^2\kappa (4d_1\kappa ^2 + 2d_2)}. \end{aligned}$$
(28)

Therefore, we have

$$\begin{aligned} S_K&\le \frac{2\eta _x^2}{(1-\rho )^2}(L^2\kappa \delta _{\kappa }^N(2c_2 + 2d_1c_1) + L^2\kappa \delta _{\kappa }^N(4d_1\kappa ^2 + 2d_2)E_K + nK(1+\kappa )^2L_{f,0}^2) \\&\le \frac{2\eta _x^2}{(1-\rho )^2}(L_{\Phi }^2 E_K + L^2\kappa \delta _{\kappa }^N(2c_2 + 2d_1c_1) + nK(1+\kappa )^2L_{f,0}^2), \end{aligned}$$

where in the first inequality we use (27) to eliminate \(B_K\). Next we eliminate \(E_K\) in this bound. By the definition of \(\eta _x\), we know:

$$\begin{aligned} \eta _x<\frac{(1-\rho )}{4\sqrt{2}L_{\Phi }}\quad \Rightarrow \quad \frac{16\eta _x^2L_{\Phi }^2}{(1-\rho )^2}<\frac{1}{2}, \end{aligned}$$

which, together with (26), yields

$$\begin{aligned} S_{K}&\le \frac{2\eta _x^2}{(1-\rho )^2}(L_{\Phi }^2(8S_K + 4n\eta _x^2\sum _{j=0}^{K-1}\Vert \overline{\partial \Phi (X_{j})} - \nabla \Phi ({{\bar{x}}}_j)\Vert ^2 + 4n\eta _x^2T_{K-1}) \nonumber \\&\quad + L^2\kappa \delta _{\kappa }^N(2c_2 + 2d_1c_1) + nK(1+\kappa )^2L_{f,0}^2) \nonumber \\&< \frac{1}{2}S_K + \frac{2\eta _x^2}{(1-\rho )^2}(4n\eta _x^2L_{\Phi }^2\sum _{j=0}^{K-1}\Vert \overline{\partial \Phi (X_{j})} - \nabla \Phi ({{\bar{x}}}_j)\Vert ^2 + 4n\eta _x^2L_{\Phi }^2T_{K-1} \nonumber \\&\quad + L^2\kappa \delta _{\kappa }^N(2c_2 + 2d_1c_1) + nK(1+\kappa )^2L_{f,0}^2). \end{aligned}$$
(29)

The above inequality indicates

$$\begin{aligned} S_K&\le \,\frac{4\eta _x^2}{(1-\rho )^2}\left( 4n\eta _x^2L_{\Phi }^2\sum _{j=0}^{K-1}\Vert \overline{\partial \Phi (X_{j})} - \nabla \Phi ({{\bar{x}}}_j)\Vert ^2 + 4n\eta _x^2L_{\Phi }^2T_{K-1}\right) \nonumber \\&\quad +\, \frac{4\eta _x^2}{(1-\rho )^2}\left( L^2\kappa \delta _{\kappa }^N(2c_2 + 2d_1c_1) + nK(1+\kappa )^2L_{f,0}^2\right) . \end{aligned}$$
(30)

Note that we have

$$\begin{aligned} \delta _y^T<\frac{L_{\Phi }^2}{12\Gamma \kappa ^2}\quad \Rightarrow \quad \delta _y^T\cdot 6\kappa ^2\cdot 2\Gamma < L_{\Phi }^2. \end{aligned}$$
(31)

Define

$$\begin{aligned} \Lambda = \frac{12L^2\kappa \delta _{\kappa }^N(2c_2 + 2d_1c_1) + 6\Gamma c_1\delta _y^T}{n}. \end{aligned}$$

By Lemma 12,

$$\begin{aligned}&\sum _{k=0}^{K}\Vert \overline{\partial \Phi (X_k)} - \nabla \Phi (\bar{x}_k)\Vert ^2\le \frac{2L_{\Phi }^2}{n}S_K + \frac{2\Gamma }{n}A_K + \frac{12L^2\kappa }{n}\delta _\kappa ^NB_K \\&\quad \le \frac{2L_{\Phi }^2}{n}S_K + \left( \frac{2\Gamma }{n}\cdot 6\kappa ^2\delta _y^T + \frac{12L^2\kappa }{n}\cdot \delta _{\kappa }^N\cdot (4d_1\kappa ^2 + 2d_2)\right) E_K \\&\qquad +\, \frac{12L^2\kappa \delta _{\kappa }^N(2c_2 + 6d_1c_1\delta _y^T) + 6\Gamma c_1\delta _y^T}{n} \\&\quad \le \frac{2L_{\Phi }^2}{n}S_K + \left( \frac{L_{\Phi }^2}{n} + \frac{12L_{\Phi }^2}{n}\right) E_K + \frac{12L^2\kappa \delta _{\kappa }^N(2c_2 + 2d_1c_1) + 6\Gamma c_1\delta _y^T}{n} \\&\quad \le \frac{2L_{\Phi }^2}{n}S_K + \frac{13L_{\Phi }^2}{n}\left( 8S_K + 4n\eta _x^2\sum _{j=0}^{K-1}\Vert \overline{\partial \Phi (X_{j})} - \nabla \Phi ({{\bar{x}}}_j)\Vert ^2 + 4n\eta _x^2T_{K-1}\right) + \Lambda \\&\quad < \frac{106L_{\Phi }^2}{n}S_K + 52\eta _x^2L_{\Phi }^2 \left( \sum _{j=0}^{K}\Vert \overline{\partial \Phi (X_{j})} - \nabla \Phi ({{\bar{x}}}_j)\Vert ^2 + T_{K}\right) + \Lambda \\&\quad \le \left( \frac{106L_{\Phi }^2}{n}\cdot \frac{16nL_{\Phi }^2\eta _x^4}{(1-\rho )^2} + 52\eta _x^2L_{\Phi }^2\right) \left( \sum _{j=0}^{K}\Vert \overline{\partial \Phi (X_{j})} - \nabla \Phi ({{\bar{x}}}_j)\Vert ^2 + T_K\right) \\&\qquad +\, \frac{106L_{\Phi }^2}{n}\cdot \frac{4\eta _x^2}{(1-\rho )^2}\left( L^2\kappa \delta _{\kappa }^N(2c_2 + 2d_1c_1) + nK(1+\kappa )^2L_{f,0}^2\right) + \Lambda , \end{aligned}$$

where the second inequality is by (26) and (27), the third inequality is by (28) and (31), the fourth inequality is obtained by (26) and the last inequality is by (30). Note that the definition of \(\eta _x\) also indicates:

$$\begin{aligned} 106L_{\Phi }^2\cdot \frac{16L_{\Phi }^2\eta _x^4}{(1-\rho )^2} + 52\eta _x^2L_{\Phi }^2 < \frac{1}{3}. \end{aligned}$$

Therefore,

$$\begin{aligned}&\sum _{k=0}^{K}\Vert \overline{\partial \Phi (X_k)} - \nabla \Phi (\bar{x}_k)\Vert ^2<\frac{1}{3}\left( \sum _{k=0}^{K}\Vert \overline{\partial \Phi (X_k)} - \nabla \Phi ({{\bar{x}}}_k)\Vert ^2 + T_K\right) \\&\quad +\frac{106L_{\Phi }^2}{n}\cdot \frac{4\eta _x^2}{(1-\rho )^2}(L^2\kappa \delta _{\kappa }^N(2c_2 + 2d_1c_1) + nK(1+\kappa )^2L_{f,0}^2) + \Lambda ,\\ \end{aligned}$$

which leads to

$$\begin{aligned}&\sum _{k=0}^{K}\Vert \overline{\partial \Phi (X_k)} - \nabla \Phi ({{\bar{x}}}_k)\Vert ^2\\&\quad \le \frac{1}{2}T_K + 106L_{\Phi }^2\cdot \frac{6\eta _x^2}{(1-\rho )^2}\left( \frac{L^2\kappa \delta _{\kappa }^N(2c_2 + 2d_1c_1)}{n} + K(1+\kappa )^2L_{f,0}^2\right) \\&\qquad +\frac{18L^2\kappa \delta _{\kappa }^N(2c_2 + 2d_1c_1) + 9\Gamma c_1\delta _y^T}{n}. \end{aligned}$$

Combining this bound with (12), we can obtain

$$\begin{aligned} T_{K}&\le \frac{2}{\eta _x}(\Phi ({{\bar{x}}}_0) - \inf _{x}\Phi (x)) + \sum _{k=0}^{K}\Vert \overline{\partial \Phi (X_k)} - \nabla \Phi (\bar{x}_k)\Vert ^2 \\&\le \frac{2}{\eta _x}(\Phi ({{\bar{x}}}_0) - \inf _{x}\Phi (x)) + \eta _x^2\cdot \frac{636L_{\Phi }^2L_{f,0}^2(1+\kappa )^2}{(1-\rho )^2}K + \frac{1}{2}T_K + \frac{1}{2}C_1, \end{aligned}$$

which implies

$$\begin{aligned}&\frac{1}{K+1}\sum _{j=0}^{K}\Vert \nabla \Phi ({{\bar{x}}}_j)\Vert ^2\\&\quad \le \frac{4}{\eta _x(K+1)}(\Phi ({{\bar{x}}}_0)-\inf _x \Phi (x)) + \eta _x^2\cdot \frac{1272L_{\Phi }^2L_{f,0}^2(1+\kappa )^2}{(1-\rho )^2} + \frac{C_1}{K+1}. \end{aligned}$$

The constant \(C_1\) satisfies

$$\begin{aligned} \frac{1}{2}C_1&= 106L_{\Phi }^2\cdot \frac{6\eta _x^2L^2\kappa \delta _{\kappa }^N(2c_2 + 2d_1c_1)}{n(1-\rho )^2} +\frac{18L^2\kappa \delta _{\kappa }^N(2c_2 + 2d_1c_1) + 9\Gamma c_1\delta _y^T}{n}\\&= {\mathcal {O}}(\eta _x^2\delta _{\kappa }^{N}\kappa ^{12} + \kappa ^5\delta _y^T) = {\mathcal {O}}(1). \end{aligned}$$

Moreover, we notice that by setting

$$\begin{aligned} N = \Theta (\sqrt{\kappa }\log \kappa ),\ T=\Theta (\kappa \log \kappa ),\ \eta _x = \Theta (K^{-\frac{1}{3}}\kappa ^{-\frac{8}{3}}),\ \eta _y = \frac{1}{\mu + L}, \end{aligned}$$

for sufficiently large K the conditions on algorithm parameters in Lemma 23 hold and

$$\begin{aligned} \frac{1}{K+1}\sum _{j=0}^{K}\Vert \nabla \Phi ({{\bar{x}}}_j)\Vert ^2 = {\mathcal {O}}\left( \frac{\kappa ^{\frac{8}{3}}}{K^{\frac{2}{3}}}\right) , \end{aligned}$$

which proves the first case of Theorems 3.1 and 19. \(\square\)

1.1.2 Case 2: Assumption 2.3 does not hold

Now we consider the case when Assumption 2.3 does not hold.

Lemma 24

$$\begin{aligned} S_{K}< \frac{\eta _x^2}{(1-\rho )^2}\sum _{j=0}^{K-1}\sum _{i=1}^{n}\Vert {\hat{\nabla }}f_i(x_{i,j}, y_{i,j}^{(T)})\Vert ^2 < \frac{\eta _x^2L_{f,0}^2}{(1-\rho )^2}nK\left( 2(1+\kappa )^2 + 2C\alpha ^N\right) . \end{aligned}$$

Proof

The first inequality follows from Lemma 20. For the second one observe that:

$$\begin{aligned}&\Vert {\hat{\nabla }}f_i(x_{i,j},y_{i,j}^{(T)})\Vert = \left\| \nabla _x f_i(x_{i,k},y_{i,k}^{(T)}) -\left( Z_{i,k}^{(N)}\right) ^{{\textsf{T}}}\nabla _y f_i(x_{i,k},y_{i,k}^{(T)})\right\| \\&\quad \le \Vert \nabla _x f_i(x_{i,k},y_{i,k}^{(T)})\Vert + \Vert (Z_{i,k}^{(N)} - Z_k^*)^{{\textsf{T}}}\nabla _y f_i(x_{i,k},y_{i,k}^{(T)})\Vert + \Vert \left( Z_k^*\right) ^{{\textsf{T}}}\nabla _y f_i(x_{i,k},y_{i,k}^{(T)})\Vert \\&\quad \le \left( 1 + \left\| \left( Z_{i,k}^{(N)}\right) ^{{\textsf{T}}}- \left( Z_k^*\right) ^{{\textsf{T}}}\right\| _2 + \kappa \right) L_{f,0}, \end{aligned}$$

where we use \(\left( Z_{i,k}^{(N)}\right) ^{{\textsf{T}}}\) to denote the output of Algorithm 1 in outer loop iteration k of agent i, and \(\left( Z_k^*\right) ^{{\textsf{T}}}\) denotes the optimal solution. By Cauchy–Schwarz inequality we know:

$$\begin{aligned} \Vert {\hat{\nabla }}f_i(x_{i,j},y_{i,j}^{(T)})\Vert ^2&\le (1+\kappa +\Vert \left( Z_{i,k}^{(N)}\right) ^{{\textsf{T}}}- \left( Z_k^*\right) ^{{\textsf{T}}}\Vert _2)^2L_{f,0}^2\\&\le (2(1+\kappa )^2 + 2\Vert \left( Z_{i,k}^{(N)}\right) ^{{\textsf{T}}}- \left( Z_k^*\right) ^{{\textsf{T}}}\Vert _2^2)L_{f,0}^2\\&\le (2(1+\kappa )^2+2C\alpha ^N)L_{f,0}^2, \end{aligned}$$

which completes the proof. \(\square\)

Taking summation on both sides of (23) and applying Lemma 24 we know:

$$\begin{aligned}&\sum _{k=0}^{K}\Vert \overline{\partial \Phi (X_k)} - \nabla \Phi (\bar{x}_k)\Vert ^2 \\&\quad \le \frac{(1+\kappa ^2)}{n}\cdot \left( \frac{36L_{f,0}^2L_{g,2}^2}{\mu ^2} + 2L_{f}^2\right) S_K \\&\qquad + 12(K+1)C\left[ \left( 1 + L^2\kappa ^2 + \frac{2L_{f,0}^2L_{g,2}^2(1+\kappa ^2)}{\mu ^2}\right) \alpha ^T + L_{f,0}^2\alpha ^N\right] \\&\quad \le \left( \frac{36L_{f,0}^2L_{g,2}^2}{\mu ^2} + 2L_{f}^2\right) \frac{(1+\kappa ^2)\eta _x^2L_{f,0}^2}{(1-\rho )^2}K(2(1+\kappa )^2 + 2C\alpha ^N) + (K+1){{\tilde{C}}}_1, \end{aligned}$$

where we define:

$$\begin{aligned} \tilde{C_1} = 12C\left[ \left( 1 + L^2\kappa ^2 + \frac{2L_{f,0}^2L_{g,2}^2(1+\kappa ^2)}{\mu ^2}\right) \alpha ^T + L_{f,0}^2\alpha ^N\right] = {\mathcal {O}}(\alpha ^T + \alpha ^N). \end{aligned}$$

The above inequality together with (12) gives

$$\begin{aligned}&\frac{1}{K+1}\sum _{k=0}^{K}\Vert \nabla \Phi ({{\bar{x}}}_k)\Vert ^2\\&\quad \le \frac{2}{\eta _x(K+1)}(\Phi ({{\bar{x}}}_0) - \inf _x\Phi (x)) + \frac{1}{K+1}\sum _{k=0}^{K}\Vert \overline{\partial \Phi (X_k)} - \nabla \Phi ({{\bar{x}}}_k)\Vert ^2 \\&\quad \le \frac{2}{\eta _x(K+1)}(\Phi ({{\bar{x}}}_0) - \inf _x\Phi (x)) \\&\hspace{3em}+ \frac{4\eta _x^2(1+\kappa ^2)L_{f,0}^2 }{(1-\rho )^2}((1+\kappa )^2 + C\alpha ^N)\left( \frac{18L_{f,0}^2L_{g,2}^2}{\mu ^2} + L_{f}^2\right) + \tilde{C_1}. \end{aligned}$$

Moreover, if we choose

$$\begin{aligned} N = \Theta (\log K),\ T=\Theta (\log K),\ \eta _x = \Theta (K^{-\frac{1}{3}}\kappa ^{-\frac{8}{3}}),\ \eta _y^{(t)} = \Theta (1) \end{aligned}$$

then we can get:

$$\begin{aligned} \frac{1}{K+1}\sum _{j=0}^{K}\Vert \nabla \Phi ({{\bar{x}}}_j)\Vert ^2 = {\mathcal {O}}\left( \frac{\kappa ^{\frac{8}{3}}}{K^{\frac{2}{3}}}\right) , \end{aligned}$$

which proves the second case of Theorems 3.1 and 19.

1.2 Proof of the convergence of DBOGT

In this section we will prove the following convergence result of Algorithm 4

Theorem 25

In Algorithm 4, suppose Assumptions 2.1 and 2.2 hold. If Assumption 2.3 holds, then by setting \(0< \eta _x<\frac{(1-\rho )^2}{8L_{\Phi }},\ 0<\eta _y< \frac{2}{\mu + L},\ T = \Theta (\kappa \log \kappa ),\ N = \Theta (\sqrt{\kappa }\log \kappa )\), we have:

$$\begin{aligned} \frac{1}{K+1}\sum _{j=0}^{K}\Vert \nabla \Phi ({{\bar{x}}}_j)\Vert ^2\le \frac{4}{\eta _x(K+1)}(\Phi ({{\bar{x}}}_0)-\inf _x \Phi (x)) + \frac{C_2}{K+1}. \end{aligned}$$

If Assumption 2.3 does not hold, then by setting

$$\begin{aligned} 0<\eta _x<\min \left( \frac{(1-\rho )^2}{14\kappa L_f}, \frac{\mu (1-\rho )^2}{21L_{f,0}L_{g,2}\kappa }\right) ,\ \eta _y=\Theta (1), \end{aligned}$$

we have:

$$\begin{aligned} \frac{1}{K+1}\sum _{j=0}^{K}\Vert \nabla \Phi (\bar{x}_j)\Vert ^2\le \frac{4}{\eta _x(K+1)}(\Phi ({{\bar{x}}}_0) - \inf _x\Phi (x)) +{\tilde{C}}_2. \end{aligned}$$

Here \(C_2 = \Theta (1)\) and \({\tilde{C}}_2 = \Theta (\alpha ^T + \alpha ^N + \frac{1}{K+1})\).

We first bound the consensus estimation error in the following lemma.

Lemma 26

In Algorithm 4, we have the following inequality holds:

$$\begin{aligned} S_K \le \frac{\eta _x^2}{(1-\rho )^4}\left(\sum _{j=1}^{K-1}\sum _{i=1}^{n}\Vert {\hat{\nabla }} f_i(x_{i,j}, y_{i,j}^{(T)}) - {\hat{\nabla }} f_i(x_{i,j-1}, y_{i,j-1}^{(T)}) \Vert ^2 + \Vert \partial \Phi (X_0)\Vert ^2\right). \end{aligned}$$

Proof

From the updates of x and u, we have:

$$\begin{aligned} {{\bar{u}}}_k = {{\bar{u}}}_{k-1} + \overline{\partial \Phi (X_k)} - \overline{\partial \Phi (X_{k-1})},\quad {{\bar{u}}}_0 = \overline{\partial \Phi (X_0)},\quad {{\bar{x}}}_{k+1} = {{\bar{x}}}_k - \eta _x{{\bar{u}}}_k, \end{aligned}$$

which implies:

$$\begin{aligned} {{\bar{u}}}_k = \overline{\partial \Phi (X_k)},\quad {{\bar{x}}}_{k+1} = \bar{x}_k - \eta _x\overline{\partial \Phi (X_k)}. \end{aligned}$$

Hence by definition of \(q_{i,k+1}\):

$$\begin{aligned} q_{i,k+1} =&\, x_{i,k+1} - {{\bar{x}}}_{k+1} =\sum _{j=1}^{n}w_{ij}x_{j,k} - \eta _x u_{i,k} - {{\bar{x}}}_k + \eta _x{{\bar{u}}}_k\\ =&\sum _{j=1}^{n}w_{ij}(x_{j,k} - {{\bar{x}}}_k) - \eta _x(u_{i,k} - {{\bar{u}}}_k)\\ =&\sum _{j=1}^{n}w_{ij}q_{j,k} - \eta _xr_{i,k} =Q_{k}We_i - \eta _x R_k e_i. \\ \end{aligned}$$

Therefore, we can write the update of the matrix \(Q_{k+1}\) as

$$\begin{aligned} Q_{k+1} = Q_kW - \eta _x R_k,\quad Q_1 = -\eta _x R_0. \end{aligned}$$

Note that \(Q_{k+1}\) takes the form of

$$\begin{aligned} Q_{k+1} = (Q_{k-1}W - \eta _x R_{k-1})W - \eta _x R_k = -\eta _x\sum _{i=0}^{k}R_iW^{k-i}. \end{aligned}$$
(32)

We then compute \(r_{i,k}\) as following

$$\begin{aligned}&r_{i,k+1} = u_{i,k+1} - {{\bar{u}}}_{k+1} \\&\quad = \sum _{j=1}^{n}w_{ij}u_{j,k} + {\hat{\nabla }} f_i(x_{i,k+1},y_{i,k+1}^{(T)}) - {\hat{\nabla }} f_i(x_{i,k},y_{i,k}^{(T)}) - {{\bar{u}}}_k - (\overline{\partial \Phi (X_{k+1})} - \overline{\partial \Phi (X_{k})})\\&\quad =\sum _{j=1}^{n}w_{ij}(u_{j,k} - {{\bar{u}}}_k) + (\partial \Phi (X_{k+1}) - \partial \Phi (X_{k}))\left( e_i - \frac{{{\textbf {1}}}_n}{n}\right) \\&\quad = R_kWe_i + (\partial \Phi (X_{k+1}) - \partial \Phi (X_{k}))\left( e_i - \frac{{{\textbf {1}}}_n}{n}\right) . \end{aligned}$$

The matrix \(R_{k+1}\) can be written as

$$\begin{aligned} R_{k+1}&= R_kW + (\partial \Phi (X_{k+1}) - \partial \Phi (X_{k}))(I - \frac{{\textbf{1}}_n{\textbf{1}}^{\top }_n}{n}) \nonumber \\&=R_0W^{k+1} + \sum _{j=0}^k(\partial \Phi (X_{j+1}) - \partial \Phi (X_{j}))\left( I - \frac{{\textbf{1}}_n{\textbf{1}}^{\top }_n}{n}\right) W^{k-j} \nonumber \\&=\partial \Phi (X_0)\left( I - \frac{{\textbf{1}}_n{\textbf{1}}^{\top }_n}{n}\right) W^{k+1} + \sum _{j=0}^k(\partial \Phi (X_{j+1}) - \partial \Phi (X_{j}))\left( I - \frac{{\textbf{1}}_n{\textbf{1}}^{\top }_n}{n}\right) W^{k-j} \nonumber \\&=\sum _{j=0}^{k+1}(\partial \Phi (X_{j}) - \partial \Phi (X_{j-1}))\left( I - \frac{{\textbf{1}}_n{\textbf{1}}^{\top }_n}{n}\right) W^{k+1-j}, \end{aligned}$$
(33)

where the third equality holds because of the initialization \(u_{i,0} = {\hat{\nabla }} f_i(x_{i,0},y_{i,0}^{(T)})\) and we denote \(\partial \Phi (X_{-1})=0\). Plugging (33) into (32) yields

$$\begin{aligned} Q_{k+1}&= -\eta _x\sum _{i=0}^{k}\sum _{j=0}^{i}(\partial \Phi (X_{j}) - \partial \Phi (X_{j-1})) \left(I - \frac{{\textbf{1}}_n{\textbf{1}}^{\top }_n}{n} \right)W^{k-j} \\&=-\eta _x\sum _{j=0}^{k}\sum _{i=j}^{k}(\partial \Phi (X_{j}) - \partial \Phi (X_{j-1})) \left(W^{k-j} - \frac{{\textbf{1}}_n{\textbf{1}}^{\top }_n}{n} \right) \\&=-\eta _x\sum _{j=0}^{k}(k+1-j)(\partial \Phi (X_{j}) - \partial \Phi (X_{j-1})) \left(W^{k-j} - \frac{{\textbf{1}}_n{\textbf{1}}^{\top }_n}{n} \right), \end{aligned}$$

where the second equality is obtained by \({\textbf{1}}_n{\textbf{1}}^{\top }_n W = {\textbf{1}}_n{\textbf{1}}^{\top }_n\) and switching the order of the summations. Therefore, we have

$$\begin{aligned}\Vert Q_{k+1}\Vert ^2 &= \eta _x^2\left\| \sum _{j=0}^{k}(k+1-j)(\partial \Phi (X_{j}) - \partial \Phi (X_{j-1}))\left( W^{k-j} - \frac{{\textbf{1}}_n{\textbf{1}}^{\top }_n}{n}\right) \right\| ^2 \nonumber \\& \le \eta _x^2\left( \sum _{j=0}^k\left\| (k+1-j)(\partial \Phi (X_{j}) - \partial \Phi (X_{j-1}))\left( W^{k-j} - \frac{{\textbf{1}}_n{\textbf{1}}^{\top }_n}{n}\right) \right\| \right) ^2 \nonumber \\& \le \eta _x^2\left( \sum _{j=0}^k\left\| (k+1-j)(\partial \Phi (X_{j}) - \partial \Phi (X_{j-1}))\right\| \left\| W^{k-j} - \frac{{\textbf{1}}_n{\textbf{1}}^{\top }_n}{n}\right\| _2\right) ^2 \nonumber \\& \le \eta _x^2 \left( \sum _{j=0}^{k} \rho ^{k-j}(k+1-j)\Vert \partial \Phi (X_{j}) - \partial \Phi (X_{j-1})\Vert ^2\right) \cdot \nonumber \\&\qquad \left( \sum _{j=0}^{k}\frac{(k+1-j)}{\rho ^{k-j}}\left\| W^{k-j}-\frac{{\textbf{1}}_n{\textbf{1}}^{\top }_n}{n}\right\| _2^2\right) \nonumber \\& \le \eta _x^2\left( \sum _{j=0}^{k} \rho ^{k-j}(k+1-j)\Vert \partial \Phi (X_{j}) - \partial \Phi (X_{j-1})\Vert ^2\right) \left( \sum _{j=0}^{k}(k+1-j)\rho ^{k-j}\right) \nonumber \\& < \frac{\eta _x^2}{(1-\rho )^2}\left( \sum _{j=0}^{k} \rho ^{k-j}(k+1-j)\Vert \partial \Phi (X_{j}) - \partial \Phi (X_{j-1})\Vert ^2\right) , \end{aligned}$$
(34)

where the second inequality is by Lemma 1, the third inequality is by Lemma 4, and the last inequality uses the fact that:

$$\begin{aligned} \sum _{j=0}^{k}(k+1-j)\rho ^{k-j} = \sum _{m=0}^{k}(m+1)\rho ^m = \frac{1 - (k+2)\rho ^{k+1} + (k+1)\rho ^{k+2}}{(1-\rho )^2}<\frac{1}{(1-\rho )^2}. \end{aligned}$$
(35)

Summing (34) over \(k=0,\ldots , K-1\), we get:

$$\begin{aligned} S_K&= \sum _{k=0}^{K-1}\Vert Q_{k+1}\Vert ^2\\&\le \frac{\eta _x^2}{(1-\rho )^2}\left( \sum _{k=0}^{K-1}\sum _{j=0}^{k} \rho ^{k-j}(k+1-j)\Vert \partial \Phi (X_{j}) - \partial \Phi (X_{j-1})\Vert ^2\right) \\&=\frac{\eta _x^2}{(1-\rho )^2}\left( \sum _{j=0}^{K-1}\sum _{k=j}^{K-1} \rho ^{k-j}(k+1-j)\Vert \partial \Phi (X_{j}) - \partial \Phi (X_{j-1})\Vert ^2\right) \\&<\frac{\eta _x^2}{(1-\rho )^4}\sum _{j=0}^{K-1}\Vert \partial \Phi (X_{j}) - \partial \Phi (X_{j-1})\Vert ^2 \\&= \frac{\eta _x^2}{(1-\rho )^4}\left( \sum _{j=1}^{K-1}\sum _{i=1}^{n}\Vert {\hat{\nabla }} f_i(x_{i,j}, y_{i,j}^{(T)}) - {\hat{\nabla }} f_i(x_{i,j-1}, y_{i,j-1}^{(T)}) \Vert ^2 + \Vert \partial \Phi (X_0)\Vert ^2\right) , \\ \end{aligned}$$

which completes the proof. \(\square\)

1.2.1 Case 1: Assumption 2.3 holds

When Assumption 2.3 holds, we have the following lemmas.

Lemma 27

Under Assumption 2.3, the following inequality holds for Algorithm 4:

$$\begin{aligned}&\sum _{j=1}^{K-1}\sum _{i=1}^{n}\Vert {\hat{\nabla }} f_i(x_{i,j}, y_{i,j}^{(T)}) - {\hat{\nabla }} f_i(x_{i,j-1}, y_{i,j-1}^{(T)})\Vert ^2\\&\hspace{6em}\le 6\Gamma A_{K-1} + 36L^2\kappa \delta _\kappa ^NB_{K-1} + 3L_{\Phi }^2E_{K-1}. \end{aligned}$$

Moreover, we have:

$$\begin{aligned} S_K \le \frac{\eta _x^2}{(1-\rho )^4}(6\Gamma A_{K-1} + 36L^2\kappa \delta _\kappa ^NB_{K-1} + 3L_{\Phi }^2E_{K-1} + \Vert \partial \Phi (X_0)\Vert ^2). \end{aligned}$$

Proof

For each term, we know that for \(j\ge 1\):

$$\begin{aligned}&\Vert {\hat{\nabla }} f_i(x_{i,j}, y_{i,j}^{(T)}) - {\hat{\nabla }} f_i(x_{i,j-1}, y_{i,j-1}^{(T)}) \Vert ^2 \\&\quad \le 3(\Vert {\hat{\nabla }} f_i(x_{i,j}, y_{i,j}^{(T)}) - \nabla \Phi _i(x_{i,j})\Vert ^2 + \Vert \nabla \Phi _i(x_{i,j}) - \nabla \Phi _i(x_{i,j-1}) \Vert ^2\\&\qquad + \Vert \nabla \Phi _i(x_{i,j-1}) - {\hat{\nabla }} f_i(x_{i,j-1}, y_{i,j-1}^{(T)})\Vert ^2) \\&\quad \le 3(\Gamma (\Vert y_i^*(x_{i,j}) - y_{i, j}^{(T)}\Vert ^2 + \Vert y_i^*(x_{i,j-1}) - y_{i, j-1}^{(T)}\Vert ^2) \\&\qquad + 6L^2\kappa \delta _\kappa ^N(\Vert v_{i,j}^* - v_{i,j}^{(0)}\Vert ^2 + \Vert v_{i,j-1}^* - v_{i,j-1}^{(0)}\Vert ^2) + L_{\Phi }^2\Vert x_{i,j} - x_{i,j-1}\Vert ^2), \end{aligned}$$

where the last inequality uses Lemmas 11 and 5. Taking summation (\(j=1,2,\ldots ,K-1\) and \(i=1,2,\ldots ,n\)) on both sides, we have:

$$\begin{aligned}&\sum _{j=1}^{K-1}\sum _{i=1}^{n}\Vert {\hat{\nabla }} f_i(x_{i,j}, y_{i,j}^{(T)}) - {\hat{\nabla }} f_i(x_{i,j-1}, y_{i,j-1}^{(T)})\Vert ^2 \\&\quad \le 6\Gamma A_{K-1} + 36L^2\kappa \delta _\kappa ^NB_{K-1} + 3L_{\Phi }^2E_{K-1}. \end{aligned}$$

Together with Lemma 26, we can prove the second inequality for \(S_K\). \(\square\)

The above lemma together with Lemma 10 and 9 gives

$$\begin{aligned} S_K&\le \frac{\eta _x^2}{(1-\rho )^4}(6\Gamma A_{K-1} + 36L^2\kappa \delta _\kappa ^NB_{K-1} + 3L_{\Phi }^2E_{K-1} + \Vert \partial \Phi (X_0)\Vert ^2) \nonumber \\ A_K&\le \delta _y^T (3c_1 + 6\kappa ^2E_K)\quad B_K\le 2c_2 + 2d_1A_{K-1} + 2d_2E_K \nonumber \\ E_K&\le 8S_K + 4n\eta _x^2\sum _{j=0}^{K-1}\Vert \overline{\partial \Phi (X_{j})} - \nabla \Phi ({{\bar{x}}}_j)\Vert ^2 + 4n\eta _x^2T_{K-1}. \end{aligned}$$
(36)

Now we can obtain the following result.

Lemma 28

Suppose Assumptions 2.1, 2.2 and 2.3 hold. Set:

$$\begin{aligned}&\delta _y^{T}<\min \left( \frac{L_{\Phi }^2}{72\kappa ^2\Gamma }, \kappa ^{-5}\right) =\Theta (\kappa ^{-5}),\\&\delta _{\kappa }^N<\min \left( \frac{L_{\Phi }^2}{72L^2\kappa (4d_1\kappa ^2 + 2d_2)}, \kappa ^{-4}\right) =\Theta (\kappa ^{-4}),\ \eta _x<\frac{(1-\rho )^2}{8L_{\Phi }}. \end{aligned}$$

For Algorithm 4, we have:

$$\begin{aligned} \frac{1}{K+1}\sum _{k=0}^{K}\Vert \nabla \Phi ({{\bar{x}}}_k)\Vert ^2\le \frac{4}{\eta _x(K+1)}(\Phi ({{\bar{x}}}_0)-\inf _x \Phi (x)) + \frac{C_2}{K+1}, \end{aligned}$$

where the constant is defined as:

$$\begin{aligned} \frac{1}{2}C_2&= \frac{15\eta _x^2L_{\Phi }^2}{n(1-\rho )^4}(\Vert \partial \Phi (X_0)\Vert ^2 + 18\Gamma c_1\delta _y^T + 36L^2\kappa \delta _{\kappa }^N(2c_2 + 2d_1c_1))\\&\quad +\frac{18L^2\kappa \delta _{\kappa }^N(2c_2 + 2d_1c_1) + 9\Gamma c_1\delta _y^T}{n} \\&=\Theta (\eta _x^2\kappa ^6 + (\eta _x^2\kappa ^6 + 1)(\kappa ^5\delta _y^T + \kappa ^4\delta _{\kappa }^N)) = \Theta (1). \end{aligned}$$

Proof

We first bound \(B_K\) as

$$\begin{aligned} \begin{aligned} B_K&\le 2c_2 + 2d_1A_K + 2d_2E_K \le 2c_2 + \frac{2}{3}d_1(3c_1 + 6\kappa ^2E_K) + 2d_2E_K \\&= 2c_2 + 2d_1c_1 + (4d_1\kappa ^2 + 2d_2)E_K. \end{aligned} \end{aligned}$$
(37)

Next we eliminate \(A_K\) and \(B_K\) in the upper bound of \(S_K\). Choose NT such that

$$\begin{aligned}&\delta _y^T\cdot 6\kappa ^2\cdot 6\Gamma< \frac{L_{\Phi }^2}{2},\quad \delta _{\kappa }^N\cdot (4d_1\kappa ^2 + 2d_2)\cdot 36L^2\kappa <\frac{L_{\Phi }^2}{2}, \end{aligned}$$

which implies

$$\begin{aligned} \begin{aligned}&\delta _y^{T}< \frac{L_{\Phi }^2}{72\kappa ^2\Gamma },\quad \delta _{\kappa }^N< \frac{L_{\Phi }^2}{72L^2\kappa (4d_1\kappa ^2 + 2d_2)}. \end{aligned} \end{aligned}$$
(38)

By (36) and the, we have

$$\begin{aligned} S_K\le \frac{\eta _x^2}{(1-\rho )^4}(4L_{\Phi }^2E_{K-1} + \Vert \partial \Phi (X_0)\Vert ^2 + 18\Gamma c_1\delta _y^T + 36L^2\kappa \delta _{\kappa }^N(2c_2 + 2d_1c_1)). \end{aligned}$$

Next we eliminate \(E_{K-1}\) in this bound. The definition of \(\eta _x\) gives \(\eta _x<\frac{(1-\rho )^2}{8L_{\Phi }}\), which implies \(\frac{32L_{\Phi }^2\eta _x^2}{(1-\rho )^4}<\frac{1}{2}.\) Together with (36) and \(E_{K-1}\le E_K\), we have:

$$\begin{aligned} S_{K}&\le \frac{\eta _x^2}{(1-\rho )^4}\left( 4L_{\Phi }^2(8S_K + 4n\eta _x^2\sum _{j=0}^{K-1}\Vert \overline{\partial \Phi (X_{j})} - \nabla \Phi ({{\bar{x}}}_j)\Vert ^2 + 4n\eta _x^2T_{K-1}\right) \\&\qquad + \Vert \partial \Phi (X_0)\Vert ^2 + 18\Gamma c_1\delta _y^T + 36L^2\kappa \delta _{\kappa }^N(2c_2 + 2d_1c_1) ) \\&\le \frac{1}{2}S_K + \frac{\eta _x^2}{(1-\rho )^4}\left( 4L_{\Phi }^2(4n\eta _x^2\sum _{j=0}^{K}\Vert \overline{\partial \Phi (X_{j})} - \nabla \Phi ({{\bar{x}}}_j)\Vert ^2 + 4n\eta _x^2T_{K}\right) \\&\qquad +\, \Vert \partial \Phi (X_0)\Vert ^2 + 18\Gamma c_1\delta _y^T + 36L^2\kappa \delta _{\kappa }^N(2c_2 + 2d_1c_1) ), \end{aligned}$$

which immediately implies

$$\begin{aligned} \begin{aligned} S_K&< \frac{2\eta _x^2}{(1-\rho )^4}(16n\eta _x^2L_{\Phi }^2\left( \sum _{j=0}^{K}\Vert \overline{\partial \Phi (X_{j})} - \nabla \Phi ({{\bar{x}}}_j)\Vert ^2 + T_{K}\right) \\&\qquad + \Vert \partial \Phi (X_0)\Vert ^2 + 18\Gamma c_1\delta _y^T + 36L^2\kappa \delta _{\kappa }^N(2c_2 + 2d_1c_1) ). \end{aligned} \end{aligned}$$
(39)

Moreover, by (19) we have

$$\begin{aligned}&\sum _{k=0}^{K}\Vert \overline{\partial \Phi (X_k)} - \nabla \Phi (\bar{x}_k)\Vert ^2\le \frac{2L_{\Phi }^2}{n}S_K + \frac{2\Gamma }{n}A_K + \frac{12L^2\kappa }{n}\delta _{\kappa }^NB_K \\&\quad \le \frac{2L_{\Phi }^2}{n}S_K + \left( \frac{L_{\Phi }^2}{6n} + \frac{L_{\Phi }^2}{6n}\right) E_K + \frac{12L^2\kappa \delta _{\kappa }^N(2c_2 + 2d_1c_1) + 6\Gamma c_1\delta _y^T}{n} \\&\quad \le \frac{2L_{\Phi }^2}{n}S_K + \frac{L_{\Phi }^2}{3n}\left( 8S_K + 4n\eta _x^2\sum _{j=0}^{K-1}\Vert \overline{\partial \Phi (X_{j})} - \nabla \Phi ({{\bar{x}}}_j)\Vert ^2 + 4n\eta _x^2T_{K-1}\right) \\&\qquad +\, \frac{12L^2\kappa \delta _{\kappa }^N(2c_2 + 2d_1c_1) + 6\Gamma c_1\delta _y^T}{n}\\&\quad < \frac{5L_{\Phi }^2}{n}S_K + \frac{4\eta _x^2L_{\Phi }^2}{3}\left( \sum _{j=0}^{K}\Vert \overline{\partial \Phi (X_{j})} - \nabla \Phi ({{\bar{x}}}_j)\Vert ^2 + T_{K}\right) \\&\qquad +\, \frac{12L^2\kappa \delta _{\kappa }^N(2c_2 + 2d_1c_1) + 6\Gamma c_1\delta _y^T}{n} \\&\quad \le \left( \frac{5L_{\Phi }^2}{n}\cdot \frac{32nL_{\Phi }^2\eta _x^4}{(1-\rho )^4} + \frac{4\eta _x^2L_{\Phi }^2}{3}\right) \left( \sum _{j=0}^{K}\Vert \overline{\partial \Phi (X_{j})} - \nabla \Phi ({{\bar{x}}}_j)\Vert ^2 + T_K\right) \\&\qquad +\, \frac{5L_{\Phi }^2}{n}\cdot \frac{2\eta _x^2}{(1-\rho )^4}\left( \Vert \partial \Phi (X_0)\Vert ^2 + 18\Gamma c_1\delta _y^T + 36L^2\kappa \delta _{\kappa }^N(2c_2 + 2d_1c_1)\right) \\&\qquad +\, \frac{12L^2\kappa \delta _{\kappa }^N(2c_2 + 2d_1c_1) + 6\Gamma c_1\delta _y^T}{n}, \end{aligned}$$

where the second inequality is by (36), (37) and (38), and the third inequality uses (36). Note that \(\eta _x\) satisfies:

$$\begin{aligned} \eta _x<\frac{(1-\rho )^2}{8L_{\Phi }}\quad \Rightarrow \quad \frac{160L_{\Phi }^4\eta _x^4}{(1-\rho )^4} + \frac{8\eta _x^2L_{\Phi }^2}{3} < \frac{1}{3}. \end{aligned}$$

Therefore, we have:

$$\begin{aligned}\sum _{k=0}^{K}\Vert \overline{\partial \Phi (X_k)} - \nabla \Phi (\bar{x}_k)\Vert ^2 &\le \frac{1}{3}\sum _{k=0}^{K}\Vert \overline{\partial \Phi (X_k)} - \nabla \Phi ({{\bar{x}}}_k)\Vert ^2 + \frac{1}{3}T_K \\&\quad + \frac{10\eta _x^2L_{\Phi }^2}{n(1-\rho )^4}(\Vert \partial \Phi (X_0)\Vert ^2 + 18\Gamma c_1\delta _y^T + 36L^2\kappa \delta _{\kappa }^N(2c_2 + 2d_1c_1)) \\&\quad + \frac{12L^2\kappa \delta _{\kappa }^N(2c_2 + 2d_1c_1) + 6\Gamma c_1\delta _y^T}{n}, \end{aligned}$$

which leads to

$$\begin{aligned}&\sum _{k=0}^{K}\Vert \overline{\partial \Phi (X_k)} - \nabla \Phi (\bar{x}_k)\Vert ^2\\&\quad \le \frac{1}{2}T_K + \frac{15\eta _x^2L_{\Phi }^2}{n(1-\rho )^4}\left( \Vert \partial \Phi (X_0)\Vert ^2 + 18\Gamma c_1\delta _y^T + 36L^2\kappa \delta _{\kappa }^N(2c_2 + 2d_1c_1)\right) \\&\qquad +\frac{18L^2\kappa \delta _{\kappa }^N(2c_2 + 2d_1c_1) + 9\Gamma c_1\delta _y^T}{n}. \end{aligned}$$

Recall (12), we have

$$\begin{aligned} \frac{1}{K+1}T_K&\le \frac{2}{\eta _x(K+1)}(\Phi ({{\bar{x}}}_0)-\inf _x \Phi (x)) + \frac{1}{K+1}\sum _{k=0}^{K}\Vert \overline{\partial \Phi (X_k)} - \nabla \Phi ({{\bar{x}}}_k)\Vert ^2 \\&\le \frac{2}{\eta _x(K+1)}(\Phi ({{\bar{x}}}_0)-\inf _x \Phi (x)) + \frac{1}{2(K+1)}T_K + \frac{1}{2(K+1)}C_2. \\ \end{aligned}$$

Therefore, we get

$$\begin{aligned} \frac{1}{K+1}\sum _{j=0}^{K}\Vert \nabla \Phi ({{\bar{x}}}_j)\Vert ^2\le \frac{4}{\eta _x(K+1)}(\Phi ({{\bar{x}}}_0)-\inf _x \Phi (x)) + \frac{C_2}{K+1}, \end{aligned}$$

where the constant is defined as following

$$\begin{aligned} \frac{1}{2}C_2&= \frac{15\eta _x^2L_{\Phi }^2}{n(1-\rho )^4}(\Vert \partial \Phi (X_0)\Vert ^2 + 18\Gamma c_1\delta _y^T + 36L^2\kappa \delta _{\kappa }^N(2c_2 + 2d_1c_1)) \\&\qquad +\frac{18L^2\kappa \delta _{\kappa }^N(2c_2 + 2d_1c_1) + 9\Gamma c_1\delta _y^T}{n} \\&=\Theta (\eta _x^2\kappa ^6 + (\eta _x^2\kappa ^6 + 1)(\kappa ^5\delta _y^T + \kappa ^4\delta _{\kappa }^N)) = \Theta (1). \end{aligned}$$

Then if we choose

$$\begin{aligned} T=\Theta (\kappa \log \kappa ), N=\Theta (\sqrt{\kappa }\log \kappa ), \eta _x = \Theta (\kappa ^{-3}), \eta _y = \frac{1}{\mu + L}, \end{aligned}$$

then the restrictions on algorithm parameters in Lemma 28 hold and we have

$$\begin{aligned} \frac{1}{K+1}\sum _{j=0}^{K}\Vert \nabla \Phi ({{\bar{x}}}_j)\Vert ^2 = {\mathcal {O}}\left( \frac{1}{K}\right) , \end{aligned}$$

which proves the first case of Theorems 3.2 and 25. \(\square\)

1.2.2 Case 2: Assumption 2.3 does not hold

We first give a bound for \(\Vert {\tilde{y}}_{j}^* - {\tilde{y}}_{j-1}^*\Vert\) in the following lemma.

Lemma 29

Recall that \({\tilde{y}}_j^* = \mathop {\mathrm {arg\,min}}\limits \frac{1}{n}\sum _{i=1}^{n}g_i(x_{i,j}, y)\). We have:

$$\begin{aligned} \Vert {\tilde{y}}_j^* - {\tilde{y}}_{j-1}^*\Vert ^2\le \frac{\kappa ^2}{n}\sum _{i=1}^{n}\Vert x_{i,j} - x_{i,j-1}\Vert ^2. \end{aligned}$$

Proof

The proof technique is similar to Lemma 13. Consider:

$$\begin{aligned}&\Vert \frac{1}{n}\sum _{i=1}^{n}\nabla _yg_i(x_{i,j-1}, {\tilde{y}}_j^*)\Vert \\&\quad = \Vert \frac{1}{n}\sum _{i=1}^{n}\nabla _yg_i(x_{i,j-1}, {\tilde{y}}_j^*) - \frac{1}{n}\sum _{i=1}^{n}\nabla _yg_i(x_{i,j-1}, {\tilde{y}}_{j-1}^* )\Vert \ge \mu \Vert {\tilde{y}}_j^* - {\tilde{y}}_{j-1}^* \Vert , \\&\Vert \frac{1}{n}\sum _{i=1}^{n}\nabla _yg_i(x_{i,j-1}, {\tilde{y}}_j^*)\Vert \\&\quad = \Vert \frac{1}{n}\sum _{i=1}^{n}\nabla _yg_i(x_{i,j-1}, {\tilde{y}}_j^*) - \frac{1}{n}\sum _{i=1}^{n}\nabla _yg_i(x_{i,j}, {\tilde{y}}_j^*) \Vert \le \frac{L}{n}\sum _{i=1}^{n}\Vert x_{i,j} - x_{i,j-1}\Vert , \end{aligned}$$

which implies:

$$\begin{aligned} \Vert {\tilde{y}}_j^* - {\tilde{y}}_{j-1}^* \Vert ^2\le \frac{\kappa ^2}{n^2}\left( \sum _{i=1}^{n}\Vert x_{i,j} - x_{i,j-1}\Vert \right) ^2\le \frac{\kappa ^2}{n}\sum _{i=1}^{n}\Vert x_{i,j}-x_{i,j-1}\Vert ^2. \end{aligned}$$

\(\square\)

Lemma 30

Suppose \(\eta _x\) satisfies

$$\begin{aligned} \eta _x \le \frac{\mu (1-\rho )^2}{21L_{f,0}L_{g,2}\kappa }. \end{aligned}$$
(40)

When the Assumption 2.3 does not hold, we have for Algorithm 4:

$$\begin{aligned} S_K&\le \frac{2\eta _x^2}{(1-\rho )^4}\left[ 3L_f^2(1+\kappa ^2)\sum _{j=1}^{K-1}\sum _{i=1}^{n}E_{K-1} + \Vert \partial \Phi (X_0)\Vert ^2\right] \\&\quad +\frac{72nKC\eta _x^2}{(1-\rho )^4}\left( \left( 1 + L^2\kappa ^2 + \frac{2L_{f,0}^2L_{g,2}^2(1+\kappa ^2)}{\mu ^2}\right) \alpha ^T + L_{f,0}^2\alpha ^N\right) \end{aligned}$$

Proof

We first notice that

$$\begin{aligned}&\Vert {\hat{\nabla }} f_i(x_{i,j}, y_{i,j}^{(T)}) - {\hat{\nabla }} f_i(x_{i,j-1}, y_{i,j-1}^{(T)}) \Vert ^2\\&\quad \le 3\Vert {\hat{\nabla }} f_i(x_{i,j}, y_{i,j}^{(T)}) - {\bar{\nabla }} f_i(x_{i,j}, {\tilde{y}}_j^*)\Vert ^2 +3\Vert {\bar{\nabla }} f_i(x_{i,j}, {\tilde{y}}_j^*) - {\bar{\nabla }} f_i(x_{i,j-1}, {\tilde{y}}_{j-1}^*)\Vert ^2 \\&\qquad + 3\Vert {\bar{\nabla }} f_i(x_{i,j-1}, {\tilde{y}}_{j-1}^*) - {\hat{\nabla }} f_i(x_{i,j-1}, y_{i,j-1}^{(T)})\Vert ^2. \end{aligned}$$

Taking summation on both sides and using Lemma 17, we have

$$\begin{aligned}&\frac{1}{n}\sum _{j=1}^{K-1}\sum _{i=1}^{n}\Vert {\hat{\nabla }} f_i(x_{i,j}, y_{i,j}^{(T)}) - {\hat{\nabla }} f_i(x_{i,j-1}, y_{i,j-1}^{(T)})\Vert ^2\\&\quad \le \frac{108L_{f,0}^2L_{g,2}^2(1 + \kappa ^2)}{n\mu ^2}S_{K-1} \\&\qquad + 36(K-1)\left( 1 + L^2\kappa ^2 + \frac{2L_{f,0}^2L_{g,2}^2(1+\kappa ^2)}{\mu ^2}\right) \cdot \frac{1}{n}\sum _{i=1}^{n}\Vert y_{i,k}^{(T)} - {{\tilde{y}}}_k^*\Vert ^2\\&\qquad + 36(K-1)CL_{f,0}^2\alpha ^N + \frac{3L_f^2}{n}\sum _{j=1}^{K-1}\sum _{i=1}^{n}(\Vert x_{i,j}-x_{i,j-1}\Vert ^2 + \Vert {\tilde{y}}_j^* - {\tilde{y}}_{j-1}^*\Vert ^2)\\&\quad \le \frac{(1-\rho )^4}{2n\eta _x^2}S_{K-1}+36KC\left( \left( 1 + L^2\kappa ^2 + \frac{2L_{f,0}^2L_{g,2}^2(1+\kappa ^2)}{\mu ^2}\right) \alpha ^T + L_{f,0}^2\alpha ^N\right) \\&\qquad +\frac{3L_f^2(1+\kappa ^2)}{n}E_{K-1}, \end{aligned}$$

where the second inequality uses Lemma 14, 29 and (40). This completes the proof together with Lemma 26. \(\square\)

Lemma 31

When the Assumption 2.3 does not hold, we further have for Algorithm 4:

$$\begin{aligned}&\frac{1}{K+1}\sum _{k=0}^{K}\Vert \overline{\partial \Phi (X_k)} - \nabla \Phi ({{\bar{x}}}_k)\Vert ^2 \\&\quad \le \frac{(1+\kappa ^2)}{n(K+1)}\cdot \left( \frac{36L_{f,0}^2L_{g,2}^2}{\mu ^2} + 2L_{f}^2\right) S_K \\&\qquad + 12C\left[ \left( 1 + L^2\kappa ^2 + \frac{2L_{f,0}^2L_{g,2}^2(1+\kappa ^2)}{\mu ^2}\right) \alpha ^T + L_{f,0}^2\alpha ^N\right] . \end{aligned}$$

Proof

Note that the above inequality is a direct result of Lemma 18. \(\square\)

Now we are ready to provide the convergence rate. Recall that from Lemma 30, 9 and inequality (12), we have:

$$\begin{aligned}&\frac{1}{K+1}\sum _{k=0}^{K}\Vert \nabla \Phi ({{\bar{x}}}_k)\Vert ^2 \nonumber \\&\quad \le \frac{2}{\eta _x(K+1)}(\Phi ({{\bar{x}}}_0) - \inf _x\Phi (x)) + \frac{1}{K+1}\sum _{k=0}^{K}\Vert \overline{\partial \Phi (X_k)} - \nabla \Phi ({{\bar{x}}}_k)\Vert ^2, \nonumber \\&S_K \le \frac{2\eta _x^2}{(1-\rho )^4}\left( 3L_f^2(1+\kappa ^2)E_{K-1} + \Vert \partial \Phi (X_0)\Vert ^2\right) \nonumber \\&\qquad +\frac{72nKC\eta _x^2}{(1-\rho )^4}\left( \left( 1 + L^2\kappa ^2 + \frac{2L_{f,0}^2L_{g,2}^2(1+\kappa ^2)}{\mu ^2}\right) \alpha ^T + L_{f,0}^2\alpha ^N\right) , \nonumber \\&E_K \le 8S_K + 4n\eta _x^2\sum _{j=0}^{K-1}\Vert \overline{\partial \Phi (X_{j})} - \nabla \Phi ({{\bar{x}}}_j)\Vert ^2 + 4n\eta _x^2T_{K-1}. \end{aligned}$$
(41)

The following lemma proves the convergence results in Theorem 25.

Lemma 32

Suppose the Assumption 2.3 does not hold. We set \(\eta _x\) as

$$\begin{aligned} \eta _x<\min \left( \frac{(1-\rho )^2}{14\kappa L_f},\ \frac{\mu (1-\rho )^2}{21L_{f,0}L_{g,2}\kappa }\right) . \end{aligned}$$
(42)

Then we have:

$$\begin{aligned} \frac{1}{K+1}\sum _{k=0}^{K}\Vert \nabla \Phi ({{\bar{x}}}_k)\Vert ^2\le \frac{6}{\eta _x(K+1)}(\Phi ({{\bar{x}}}_0) - \inf _x\Phi (x)) + \frac{\Vert \partial \Phi (X_0)\Vert ^2 }{K+1} +{\tilde{C}}_2, \end{aligned}$$

where the constant is given by:

$$\begin{aligned} \frac{{\tilde{C}}_2}{6} =&\, 6L^2(1+\kappa ^2)C\alpha ^T + 6L_{f,0}^2C\alpha ^N \\&\quad + 2L_f^2(1 + \kappa ^2)\cdot \frac{2\eta _x^2}{(1-\rho )^4}\left[ 6L^2(1 + \kappa ^2)nC\alpha ^T + 6nL_{f,0}^2C\alpha ^N + \frac{\Vert \partial \Phi (X_0)\Vert ^2}{K+1}\right] \\&=\Theta \left(\alpha ^T + \alpha ^N + \frac{1}{K+1} \right). \end{aligned}$$

Proof

We first eliminate \(E_{K-1}\) in the upper bound of \(S_K\). Note that (42) implies

$$\begin{aligned} \frac{2\eta _x^2}{(1-\rho )^4}\cdot 3L_f^2(1+\kappa ^2)\cdot 8<\frac{1}{2}, \end{aligned}$$

which together with \(E_{K-1}\le E_K\) and the upper bounds of \(S_K\) and \(E_K\) in (41) gives

$$\begin{aligned} S_K&\le \frac{1}{2}\left( S_K + \frac{\eta _x^2}{2}\sum _{j=0}^{K-1}\Vert \overline{\partial \Phi (X_{j})} - \nabla \Phi ({{\bar{x}}}_j)\Vert ^2 + \frac{\eta _x^2}{2}T_{K-1}\right) + \frac{2\eta _x^2}{(1-\rho )^4}\Vert \partial \Phi (X_0)\Vert ^2\\&\quad + \frac{2\eta _x^2}{(1-\rho )^4}\left( 36nKC\left( \left( 1 + L^2\kappa ^2 + \frac{2L_{f,0}^2L_{g,2}^2(1+\kappa ^2)}{\mu ^2}\right) \alpha ^T + L_{f,0}^2\alpha ^N\right) \right) . \end{aligned}$$

Hence we know

$$\begin{aligned} S_K&\le \frac{\eta _x^2}{2}\sum _{j=0}^{K-1}\Vert \overline{\partial \Phi (X_{j})} - \nabla \Phi ({{\bar{x}}}_j)\Vert ^2 + \frac{\eta _x^2}{2}T_{K-1} + \frac{4\eta _x^2}{(1-\rho )^4}\Vert \partial \Phi (X_0)\Vert ^2 \\&\quad + \frac{4\eta _x^2}{(1-\rho )^4}\left( 36nKC\left( \left( 1 + L^2\kappa ^2 + \frac{2L_{f,0}^2L_{g,2}^2(1+\kappa ^2)}{\mu ^2}\right) \alpha ^T + L_{f,0}^2\alpha ^N\right) \right) . \end{aligned}$$

By Lemma 31, we have

$$\begin{aligned}&\frac{1}{K+1}\sum _{k=0}^{K}\Vert \overline{\partial \Phi (X_k)} - \nabla \Phi ({{\bar{x}}}_k)\Vert ^2\nonumber \\&\quad \le \frac{(1+\kappa ^2)}{n(K+1)}\cdot \left( \frac{36L_{f,0}^2L_{g,2}^2}{\mu ^2} + 2L_{f}^2\right) S_K \nonumber \\&\qquad +\, 12C\left[ \left( 1 + L^2\kappa ^2 + \frac{2L_{f,0}^2L_{g,2}^2(1+\kappa ^2)}{\mu ^2}\right) \alpha ^T + L_{f,0}^2\alpha ^N\right] \nonumber \\&\quad \le \frac{1}{3(K+1)}\left( \sum _{j=0}^{K-1}\Vert \overline{\partial \Phi (X_{j})} - \nabla \Phi ({{\bar{x}}}_j)\Vert ^2 + T_{K-1}\right) + \frac{\tilde{C}_2}{3}, \end{aligned}$$
(43)

where the second inequality holds since we have (42), which implies

$$\begin{aligned} \eta _x^2(1+\kappa ^2)\cdot \left( \frac{36L_{f,0}^2L_{g,2}^2}{\mu ^2} + 2L_{f}^2\right) \le \frac{1}{4}. \end{aligned}$$

The constant is defined as:

$$\begin{aligned} \frac{{\tilde{C}}_2}{3}&=\, 12C\left[ \left( 1 + L^2\kappa ^2 + \frac{2L_{f,0}^2L_{g,2}^2(1+\kappa ^2)}{\mu ^2}\right) \alpha ^T + L_{f,0}^2\alpha ^N\right] + \frac{1}{(1-\rho )^4}\frac{\Vert \partial \Phi (X_0)\Vert ^2}{n(K+1)}\\&\quad + \frac{1}{(1-\rho )^4}\left( 36C\left( \left( 1 + L^2\kappa ^2 + \frac{2L_{f,0}^2L_{g,2}^2(1+\kappa ^2)}{\mu ^2}\right) \alpha ^T + L_{f,0}^2\alpha ^N\right) \right) \\&=\Theta \left(\alpha ^T + \alpha ^N + \frac{1}{K+1} \right). \end{aligned}$$

From (43) we know

$$\begin{aligned} \frac{1}{K+1}\sum _{k=0}^{K}\Vert \overline{\partial \Phi (X_k)} - \nabla \Phi ({{\bar{x}}}_k)\Vert ^2< \frac{{\tilde{C}}_2}{2} + \frac{1}{2(K+1)}T_{K-1}. \end{aligned}$$

Combining the above inequality, Lemma 8, and \(T_{K-1}\le T_K\), we have

$$\begin{aligned} \frac{1}{K+1}\sum _{k=0}^{K}\Vert \nabla \Phi (\bar{x}_k)\Vert ^2<\frac{2}{\eta _x(K+1)}(\Phi ({{\bar{x}}}_0) - \inf _x\Phi (x)) + \frac{{\tilde{C}}_2}{2} + \frac{1}{2(K+1)}T_K. \end{aligned}$$

Hence

$$\begin{aligned} \frac{1}{K+1}T_K<\frac{4}{\eta _x(K+1)}(\Phi ({{\bar{x}}}_0) - \inf _x\Phi (x))+{\tilde{C}}_2. \end{aligned}$$

Furthermore, by setting

$$\begin{aligned} N = \Theta (\log K),\ T = \Theta (\log K),\ \eta _x=\Theta (\kappa ^{-3}),\ \eta _y=\Theta (1) \end{aligned}$$

we have

$$\begin{aligned} \frac{1}{K+1}\sum _{j=0}^{K}\Vert \nabla \Phi ({{\bar{x}}}_j)\Vert ^2 = {\mathcal {O}}\left( \frac{1}{K}\right) , \end{aligned}$$

which proves the second case of Theorems 3.2 and 25. \(\square\)

1.3 Proof of the convergence of DSBO

In this section we will prove the convergence result of the DSBO algorithm.

Theorem 33

In Algorithm 5, suppose Assumptions 2.1 and 2.2 hold. If Assumption 2.3 holds, then by setting \(M = \Theta (\log K),\ T = \Omega (\kappa \log \kappa ),\ \beta \le \min \left( \frac{\mu }{\mu ^2 + \sigma _{g,2}^2},\ \frac{1}{L}\right) ,\ \eta _x\le \frac{1}{L_{\Phi }},\ \eta _y< \frac{2}{\mu + L}\), we have:

$$\begin{aligned}&\frac{1}{K+1}\sum _{k=0}^{K}{\mathbb {E}}\left[ \Vert \nabla \Phi ({{\bar{x}}}_k)\Vert ^2\right] \\&\quad \le \frac{2}{\eta _x(K+1)}({\mathbb {E}}\left[ \Phi ({{\bar{x}}}_0)\right] -\inf _x \Phi (x)) + \frac{3\eta _y L_f^2 \sigma _{g,1}^2}{\mu } + \frac{3\eta _x^2 L_{\Phi }^2 }{(1-\rho )^2}{\tilde{C}}_f^2 + L\eta _x{\tilde{\sigma }}_f^2 + C_3. \end{aligned}$$

If Assumption 2.3 does not hold, then by setting \(\eta _x\le \frac{1}{L_{\Phi }},\ \eta _y^{(t)} = {\mathcal {O}}(\frac{1}{t})\), we have:

$$\begin{aligned}&\frac{1}{K+1}\sum _{k=0}^{K}{\mathbb {E}}\left[ \Vert \nabla \Phi ({{\bar{x}}}_k)\Vert ^2\right] \\&\quad \le \frac{2}{\eta _x(K+1)}({\mathbb {E}}\left[ \Phi (\overline{x_{0}})\right] - \inf _x\Phi (x)) + \left( \frac{36L_{f,0}^2L_{g,2}^2}{\mu ^2} + 2L_{f}^2\right) \frac{\eta _x^2(1+\kappa ^2)}{(1-\rho )^2}{\tilde{C}}_f^2 \\&\qquad +L\eta _x\left( 4\sigma _f^2(1+\kappa ^2) + (8L_{f,0}^2 + 4\sigma _f^2)\frac{C}{N}\right) + {\tilde{C}}_3. \end{aligned}$$

Here \(C = \Theta (1),\ C_3 = \Theta (\eta _x^2 + \frac{1}{K+1})\) and \({\tilde{C}}_3 = {\mathcal {O}}\left( \frac{1}{T}+\alpha ^N\right)\).

We first define the following filtration:

$$\begin{aligned}&{\mathcal {F}}_k = \sigma \left( \bigcup _{i=1}^{n}\{x_{i,0}, x_{i,1},\ldots ,x_{i,k}\}\right) , \\&{\mathcal {G}}_{i,j}^{(t)} = \sigma \left( \{y_{i,l}^{(s)}: 0\le l\le j, 0\le s\le t\}\bigcup \{x_{i,l}: 0\le l\le j\}\right) . \end{aligned}$$

Then in both cases we have the following lemma.

Lemma 34

If \(\eta _x\le \frac{1}{L_{\Phi }}\), then we have:

$$\begin{aligned}&{\mathbb {E}}\left[ \Vert \nabla \Phi ({{\bar{x}}}_k)\Vert ^2\right] \\&\quad \le \frac{2}{\eta _x}({\mathbb {E}}\left[ \Phi ({{\bar{x}}}_k)\right] - {\mathbb {E}}\left[ \Phi ({{\bar{x}}}_{k+1})\right] ) + {\mathbb {E}}\left[ \Vert {\mathbb {E}}\left[ \overline{\partial \Phi (X_k; \phi )}\vert {\mathcal {F}}_k\right] - \nabla \Phi ({{\bar{x}}}_k)\Vert ^2\right] \\&\qquad + L\eta _x {\mathbb {E}}\Vert \left[ \overline{\partial \Phi (X_k; \phi )}\right] - {\mathbb {E}}\left[ \overline{\partial \Phi (X_k; \phi )}\vert {\mathcal {F}}_k\right] \Vert ^2. \end{aligned}$$

Proof

In each iteration of Algorithm 5, we have:

$$\begin{aligned} {{\bar{x}}}_{k+1} = {{\bar{x}}}_k - \eta _x \overline{\partial \Phi (X_k;\phi )}. \end{aligned}$$
(44)

The \(L_{\Phi }\)-smoothness of \(\Phi\) indicates that

$$\begin{aligned} \Phi ({{\bar{x}}}_{k+1}) - \Phi ({{\bar{x}}}_k) \le \nabla \Phi (\bar{x}_k)^{{\textsf{T}}}(-\eta _x\overline{\partial \Phi (X_k;\phi )}) + \frac{L_{\Phi }\eta _x^2}{2}\Vert \overline{\partial \Phi (X_k;\phi )}\Vert ^2. \end{aligned}$$

Taking conditional expectation with respect to \({\mathcal {F}}_k\) on both sides, we have the following

$$\begin{aligned}&{\mathbb {E}}\left[ \Phi ({{\bar{x}}}_{k+1})\vert {\mathcal {F}}_k \right] - \Phi (\bar{x}_k)\\&\quad \le \nabla \Phi ({{\bar{x}}}_k)^{{\textsf{T}}}(-\eta _x{\mathbb {E}}\left[ \overline{\partial \Phi (X_k; \phi )}\vert {\mathcal {F}}_k\right] ) + \frac{L_{\Phi }\eta _x^2}{2} {\mathbb {E}}\left[ \Vert \overline{\partial \Phi (X_k;\phi )}\Vert ^2\vert {\mathcal {F}}_k\right] \\&\quad =-\frac{\eta _x}{2}(\Vert \nabla \Phi ({{\bar{x}}}_k)\Vert ^2 + \Vert {\mathbb {E}}\left[ \overline{\partial \Phi (X_k; \phi )}\vert {\mathcal {F}}_k\right] \Vert ^2 - \Vert {\mathbb {E}}\left[ \overline{\partial \Phi (X_k; \phi )}\vert {\mathcal {F}}_k\right] - \nabla \Phi ({{\bar{x}}}_k)\Vert ^2) \\&\qquad + \frac{L_{\Phi }\eta _x^2}{2}(\Vert {\mathbb {E}}\left[ \overline{\partial \Phi (X_k; \phi )}\vert {\mathcal {F}}_k\right] \Vert ^2 + {\mathbb {E}}\left[ \Vert \overline{\partial \Phi (X_k; \phi )} - {\mathbb {E}}\left[ \overline{\partial \Phi (X_k; \phi )}\vert {\mathcal {F}}_k\right] \Vert ^2\vert {\mathcal {F}}_k\right] ) \\&\quad = \left( \frac{L_{\Phi }\eta _x^2}{2} - \frac{\eta _x}{2}\right) \Vert {\mathbb {E}}\left[ \overline{\partial \Phi (X_k; \phi )}\vert {\mathcal {F}}_k\right] \Vert ^2 \\&\qquad + \frac{L_{\Phi }\eta _x^2}{2} {\mathbb {E}}\left[ \Vert \overline{\partial \Phi (X_k; \phi )} - {\mathbb {E}}\left[ \overline{\partial \Phi (X_k; \phi )}\vert {\mathcal {F}}_k\right] \Vert ^2\vert {\mathcal {F}}_k\right] \\&\qquad - \frac{\eta _x}{2}(\Vert \nabla \Phi ({{\bar{x}}}_k)\Vert ^2 - \Vert {\mathbb {E}}\left[ \overline{\partial \Phi (X_k; \phi )}\vert {\mathcal {F}}_k\right] - \nabla \Phi ({{\bar{x}}}_k)\Vert ^2) \\&\quad \le \frac{L_{\Phi }\eta _x^2}{2} {\mathbb {E}}\left[ \Vert \overline{\partial \Phi (X_k; \phi )} - {\mathbb {E}}\left[ \overline{\partial \Phi (X_k; \phi )}\vert {\mathcal {F}}_k\right] \Vert ^2\vert {\mathcal {F}}_k\right] \\&\qquad -\frac{\eta _x}{2}(\Vert \nabla \Phi ({{\bar{x}}}_k)\Vert ^2 - \Vert {\mathbb {E}}\left[ \overline{\partial \Phi (X_k; \phi )}\vert {\mathcal {F}}_k\right] - \nabla \Phi ({{\bar{x}}}_k)\Vert ^2), \end{aligned}$$

where the second inequality holds since we pick \(\eta _x\le \frac{1}{L}\). Thus we can take expectation again and use tower property to obtain:

$$\begin{aligned}&\frac{\eta _x}{2}{\mathbb {E}}\left[ \Vert \nabla \Phi ({{\bar{x}}}_k)\Vert ^2\right] \nonumber \\&\quad \le {\mathbb {E}}\left[ \Phi ({{\bar{x}}}_k)\right] - {\mathbb {E}}\left[ \Phi (\bar{x}_{k+1})\right] + \frac{\eta _x}{2}{\mathbb {E}}\left[ \Vert {\mathbb {E}}\left[ \overline{\partial \Phi (X_k; \phi )}\vert {\mathcal {F}}_k\right] - \nabla \Phi ({{\bar{x}}}_k)\Vert ^2\right] \nonumber \\&\qquad +\frac{L_{\Phi }\eta _x^2}{2} {\mathbb {E}}\Vert \left[ \overline{\partial \Phi (X_k; \phi )}\right] - {\mathbb {E}}\left[ \overline{\partial \Phi (X_k; \phi )}\vert {\mathcal {F}}_k\right] \Vert ^2. \end{aligned}$$
(45)

which completes the proof. \(\square\)

1.3.1 Case 1: Assumption 2.3 holds

Lemma 35

Suppose \(\beta \le \frac{1}{L}\) and Assumption 2.3 holds, we have:

$$\begin{aligned} \left\| {\mathbb {E}}\left[ {\hat{\nabla }} f_i(x_{i,k},y_{i,k}^{(T)};\phi _{i,k})\vert {\mathcal {F}}_k\right] - {\bar{\nabla }} f_i(x_{i,k},y_{i,k}^{(T)})\right\| _2 \le L_{f,0}(1-\beta \mu )^M\kappa . \end{aligned}$$

Proof

We first consider the expectation

$$\begin{aligned}&{\mathbb {E}}\left[ {\hat{\nabla }} f_i(x_{i,k},y_{i,k}^{(T)};\phi _{i,k})\vert {\mathcal {F}}_k\right] \nonumber \\&\quad = \nabla _xf_i(x_{i,k}, y_{i,k}^{(T)})\nonumber \\&\qquad - \beta \nabla _{xy}g(x_{i,k}, y_{i,k}^{(T)})\sum _{j=0}^{M-1} \big (I - \beta \nabla _y^2g(x_{i,k}, y_{i,k}^{(T)}) \big )^j\nabla _yf_i(x_{i,k}, y_{i,k}^{(T)}). \end{aligned}$$
(46)

Notice that for the finite sum we have:

$$\begin{aligned}\beta \sum _{j=0}^{M-1} \left( I - \beta \nabla _y^2g(x_{i,k}, y_{i,k}^{(T)}) \right) ^j &= \beta \left( \beta \nabla _y^2g(x_{i,k}, y_{i,k}^{(T)})\right) ^{-1}\left( I - (I - \beta \nabla _y^2g(x_{i,k}, y_{i,k}^{(T)}))^M\right) \\& =\left( \nabla _y^2g(x_{i,k}, y_{i,k}^{(T)})\right) ^{-1}\left( I - (I - \beta \nabla _y^2g(x_{i,k}, y_{i,k}^{(T)}))^M\right) , \end{aligned}$$

which implies:

$$\begin{aligned} \left\| \beta \sum _{j=0}^{M-1} \left( I - \beta \nabla _y^2g(x_{i,k}, y_{i,k}^{(T)}) \right) ^j - \left( \nabla _y^2g(x_{i,k}, y_{i,k}^{(T)})\right) ^{-1}\right\| _2\le \frac{ (1-\beta \mu )^M}{\mu }. \end{aligned}$$
(47)

The above inequality and the fact that

$$\begin{aligned} {\bar{\nabla }} f_i(x_{i,k},y_{i,k}^{(T)}) = \nabla _xf_i(x_{i,k}, y_{i,k}^{(T)}) - \nabla _{xy}g(x_{i,k}, y_{i,k}^{(T)})\left( \nabla _y^2g(x_{i,k}, y_{i,k}^{(T)})\right) ^{-1}\nabla _yf_i(x_{i,k}, y_{i,k}^{(T)}) \end{aligned}$$

imply

$$\begin{aligned} \left\| {\mathbb {E}}\left[ {\hat{\nabla }} f_i(x_{i,k},y_{i,k}^{(T)};\phi _{i,k})\vert {\mathcal {F}}_k\right] - {\bar{\nabla }} f_i(x_{i,k},y_{i,k}^{(T)})\right\| _2 \le L_{f,0}(1-\beta \mu )^M\kappa , \end{aligned}$$

which completes the proof. \(\square\)

Lemma 36

Under Assumption 2.3, we have:

$$\begin{aligned}&\sum _{k=0}^{K}\Vert {\mathbb {E}}\left[ \overline{\partial \Phi (X_k; \phi )}\vert {\mathcal {F}}_k\right] - \nabla \Phi ({{\bar{x}}}_k)\Vert ^2\nonumber \\&\quad \le 3 \left((K+1)L_{f,0}^2(1-\beta \mu )^{2M}\kappa ^2 + \frac{L_f^2}{n}A_K + \frac{L_{\Phi }^2}{n}S_K \right). \end{aligned}$$
(48)

Proof

We first bound each component of the gradient error as

$$\begin{aligned}&\Vert {\mathbb {E}}\left[ {\hat{\nabla }}f_i(x_{i,k}, y_{i,k}^{(T)};\phi _{i,k})\vert {\mathcal {F}}_k\right] - \nabla \Phi _i({{\bar{x}}}_k)\Vert ^2 \\&\quad \le 3(\Vert {\mathbb {E}}\left[ {\hat{\nabla }}f_i(x_{i,k}, y_{i,k}^{(T)};\phi _{i,k})\vert {\mathcal {F}}_k\right] - {\bar{\nabla }} f_i(x_{i,k},y_{i,k}^{(T)})\Vert ^2 \\&\qquad +\Vert {\bar{\nabla }} f_i(x_{i,k},y_{i,k}^{(T)}) - \nabla f_i(x_{i,k}, y_i^*(x_{i,k}))\Vert ^2+\Vert \nabla f_i(x_{i,k}, y_i^*(x_{i,k})) - \nabla \Phi _i({{\bar{x}}}_k)\Vert ^2) \\&\quad \le 3(L_{f,0}^2(1-\beta \mu )^{2M}\kappa ^2 + L_f^2\Vert y_{i,k}^{(T)} - y_i^*(x_{i,k})\Vert ^2 + L_{\Phi }^2\Vert x_{i,k} - {{\bar{x}}}_k\Vert ^2), \end{aligned}$$

where the second inequality is obtained by Lemmas 35 and 5. Taking summation on both sides over \(i=1,\ldots ,n\), we have:

$$\begin{aligned}&\Vert {\mathbb {E}}\left[ \overline{\partial \Phi (X_k; \phi )}\vert {\mathcal {F}}_k\right] - \nabla \Phi ({{\bar{x}}}_k)\Vert ^2 \\&\quad \le \frac{1}{n}\sum _{i=1}^{n}\Vert {\mathbb {E}}\left[ {\hat{\nabla }}f_i(x_{i,k}, y_{i,k}^{(T)};\phi _{i,k})\vert {\mathcal {F}}_k\right] - \nabla \Phi _i({{\bar{x}}}_k)\Vert ^2 \\&\quad \le 3\left( L_{f,0}^2(1-\beta \mu )^{2M}\kappa ^2 + \frac{L_f^2}{n} \sum _{i=1}^{n}\Vert y_{i,k}^{(T)} - y_i^*(x_{i,k})\Vert ^2 + \frac{L_{\Phi }^2}{n} \sum _{i=1}^{n}\Vert x_{i,k} - {{\bar{x}}}_k\Vert ^2\right) . \\ \end{aligned}$$

Taking summation on both sides over \(k=0,\ldots ,K\), we know

$$\begin{aligned}&\sum _{k=0}^{K}\Vert {\mathbb {E}}\left[ \overline{\partial \Phi (X_k; \phi )}\vert {\mathcal {F}}_k\right] - \nabla \Phi ({{\bar{x}}}_k)\Vert ^2\\&\quad \le 3 \left((K+1)L_{f,0}^2(1-\beta \mu )^{2M}\kappa ^2 + \frac{L_f^2}{n}A_K + \frac{L_{\Phi }^2}{n}S_K \right), \end{aligned}$$

which completes the proof. \(\square\)

The following lemma characterizes the variance of the hypergradient estimation.

Lemma 37

Suppose \(\beta\) in Algorithm 2 satisfies

$$\begin{aligned} \beta \le \min \left( \frac{\mu }{\mu ^2 + \sigma _{g,2}^2},\ \frac{1}{L}\right) \end{aligned}$$
(49)

Under Assumptions 2.12.4, we have:

$$\begin{aligned} \begin{aligned} {\mathbb {E}}\left[ \Vert {\mathbb {E}}\left[ {\hat{\nabla }} f_i(x_{i,k},y_{i,k}^{(T)};\phi _{i,k})\vert {\mathcal {F}}_k\right] - {\hat{\nabla }} f_i(x_{i,k},y_{i,k}^{(T)};\phi _{i,k})\Vert ^2\right]&\le {\tilde{\sigma }}_f^2, \\ {\mathbb {E}}\left[ \Vert \left[ \overline{\partial \Phi (X_k; \phi )}\right] - {\mathbb {E}}\left[ \overline{\partial \Phi (X_k; \phi )}\vert {\mathcal {F}}_k\right] \Vert ^2\right]&\le \frac{{\tilde{\sigma }}_f^2}{n}, \end{aligned} \end{aligned}$$
(50)

where the constants are defined as

$$\begin{aligned} {\tilde{\sigma }}_f^2 = \sigma _{f,1}^2 + \frac{2(\sigma _{g,2}^2 + L^2)(\sigma _{f,1}^2 + L_{f,0}^2)}{\mu ^2} = {\mathcal {O}}(\kappa ^2). \end{aligned}$$

Proof

We first notice that in the stochastic case of Algorithm 2 under Assumption 2.3, for each agent i we have

$$\begin{aligned} H_M\cdot \nabla _yf_i(x, y;\phi ^{(0)}) = \beta \sum _{s=0}^{M-1}\prod _{n=1}^{s}(I - \beta \nabla _y^2g_i(x, y;\phi ^{(M+1-n)}))\nabla _yf_i(x, y;\phi ^{(0)}). \end{aligned}$$
(51)

For \(m=1, 2,\ldots , M-1\) we define

$$\begin{aligned}&A = \nabla _y^2 g_i(x,y),\ A_m = \nabla _y^2g_i(x, y;\phi ^{(m+1)}),\ b_0 = \nabla _yf_i(x, y;\phi ^{(0)}), \\&x_m = \beta \sum _{s=0}^{m-1}\prod _{n=1}^{s}(I - \beta A_{m-n})b_0,\ x_0 = 0, \end{aligned}$$

which gives

$$\begin{aligned} x_{m+1} = (I - \beta A_m)x_m + \beta b_0. \end{aligned}$$
(52)

For simplicity in the proof of this lemma we denote by \({\mathbb {E}}_0\) the conditional expectation given \(\phi ^{(0)}\). In other words we have \({\mathbb {E}}_0\left[ x\right] = {\mathbb {E}}\left[ x\vert \phi ^{(0)}\right]\) for any random vector (or matrix) x. From (52) we know

$$\begin{aligned} \Vert {\mathbb {E}}_0\left[ x_m\right] \Vert = \beta \left\| \sum _{n=1}^{M-1}\left( I - \beta A\right) ^nb_0\right\| = \left\| A^{-1}\left( I - (I - \beta A)^M\right) b_0\right\| \le \frac{\Vert b_0\Vert }{\mu }. \end{aligned}$$
(53)

Combining (52) and (53), we know

$$\begin{aligned}&{\mathbb {E}}_0\left[ \Vert x_{m+1} - {\mathbb {E}}_0\left[ x_{m+1}\right] \Vert ^2\right] \\&\quad = {\mathbb {E}}_0\left[ \Vert (I - \beta A)(x_m - {\mathbb {E}}_0\left[ x_m\right] ) + \beta (A - A_m)x_m\Vert ^2\right] \\&\quad = {\mathbb {E}}_0\left[ \Vert (I - \beta A)(x_m - {\mathbb {E}}_0\left[ x_m\right] )\Vert ^2\right] + \beta ^2{\mathbb {E}}_0\left[ \Vert (A-A_m)x_m\Vert ^2\right] \\&\quad \le (1-\beta \mu )^2{\mathbb {E}}_0\left[ \Vert x_m - {\mathbb {E}}_0\left[ x_m\right] \Vert ^2\right] + \beta ^2\sigma _{g,2}^2({\mathbb {E}}_0\left[ \Vert x_m - {\mathbb {E}}_0\left[ x_m\right] \Vert ^2\right] + \Vert {\mathbb {E}}_0\left[ x_m\right] \Vert ^2) \\&\quad \le (1-\beta \mu ){\mathbb {E}}_0\left[ \Vert x_m - {\mathbb {E}}_0\left[ x_m\right] \Vert ^2\right] + \frac{\beta ^2\sigma _{g,2}^2\Vert b_0\Vert ^2}{\mu ^2} \\&\quad \le (1-\beta \mu )^{m+1}{\mathbb {E}}\left[ \Vert x_0 - {\mathbb {E}}_0\left[ x_0\right] \Vert ^2\right] + \frac{\beta ^2\sigma _{g,2}^2\Vert b_0\Vert ^2}{\mu ^2}\left( \sum _{i=0}^{m}(1-\beta \mu )^i\right) \le \frac{\beta \sigma _{g,2}^2\Vert b_0\Vert ^2}{\mu ^3}. \end{aligned}$$

The second equality uses the independence, the second inequality uses (49), and the third inequality repeats the second inequality for m times. From the above inequality we know that the variance of \(x_m\), namely, (51), has bounded variance since

$$\begin{aligned} {\mathbb {E}}\left[ \Vert x_{M} - {\mathbb {E}}_0\left[ x_{M}\right] \Vert ^2\right] \le \frac{\beta \sigma _{g,2}^2{\mathbb {E}}\left[ \Vert b_0\Vert ^2\right] }{\mu ^3}\le \frac{\beta \sigma _{g,2}^2(\sigma _{f,1}^2 + L_{f,0}^2)}{\mu ^3}\le \frac{\sigma _{f,1}^2 + L_{f,0}^2}{\mu ^2}, \end{aligned}$$

where the second inequality uses Assumption 2.1 and the third inequality uses (49). We further know from the above conclusion and (53) that

$$\begin{aligned} \begin{aligned}&{\mathbb {E}}\left[ \Vert x_M - {\mathbb {E}}\left[ x_M\right] \Vert ^2\right] \\&\quad \le {\mathbb {E}}\left[ \Vert x_M\Vert ^2\right] = {\mathbb {E}}\left[ \Vert x_M - {\mathbb {E}}_0\left[ x_M\right] \Vert ^2\right] + {\mathbb {E}}\left[ \Vert {\mathbb {E}}_0\left[ x_M\right] \Vert ^2\right] \le \frac{2(\sigma _{f,1}^2 + L_{f,0}^2)}{\mu ^2}. \end{aligned} \end{aligned}$$
(54)

Hence in Algorithm 2 (stochastic case under Assumption 2.3) we have the following decomposition:

$$\begin{aligned}{\hat{\nabla }} f_i - {\mathbb {E}}\left[ {\hat{\nabla }} f_i\right] &= \nabla _xf_i(x, y;\phi ^{(0)}) - \nabla _x f_i(x,y) + \nabla _{xy}g_i(x,y){\mathbb {E}}\left[ x_M\right] - \nabla _{xy}g_i(x, y;\phi ^{(1)})x_{M} \\& = \nabla _xf_i(x, y;\phi ^{(0)}) - \nabla _x f_i(x,y) + (\nabla _{xy}g_i(x,y) - \nabla _{xy}g_i(x, y;\phi ^{(1)}))x_{M} \\&\quad + \nabla _{xy}g_i(x,y)({\mathbb {E}}\left[ x_M\right] - x_M), \end{aligned}$$

which implies

$$\begin{aligned}&{\mathbb {E}}\left[ \left\| {\hat{\nabla }} f_i - {\mathbb {E}}\left[ {\hat{\nabla }} f_i\right] \right\| ^2\vert x, y\right] \\&\quad = {\mathbb {E}}\left[ \Vert \nabla _xf_i(x, y;\phi ^{(0)}) - \nabla _x f_i(x,y)\Vert ^2\vert x, y\right] \\&\qquad + {\mathbb {E}}\left[ \Vert (\nabla _{xy}g_i(x,y) - \nabla _{xy}g_i(x, y;\phi ^{(1)}))x_{M}\Vert ^2\vert x, y\right] \\&\qquad + {\mathbb {E}}\left[ \Vert \nabla _{xy}g_i(x,y)({\mathbb {E}}\left[ x_M\right] - x_M)\Vert ^2\vert x, y\right] \\&\quad \le \sigma _{f,1}^2 + (\sigma _{g,2}^2 + L^2){\mathbb {E}}\left[ \Vert x_M\Vert ^2\right] \le \sigma _{f,1}^2 + \frac{2(\sigma _{g,2}^2 + L^2)(\sigma _{f,1}^2 + L_{f,0}^2)}{\mu ^2} = {{\tilde{\sigma }}}_f^2, \end{aligned}$$

where the first inequality uses the independence between different samples, the first inequality uses Assumptions 2.1 and 2.4 and the second inequality uses (54). Hence we know the first inequality of (50) holds. Furthermore the second inequality of (50) is true since for any n independent random vectors \(v_1,\ldots ,v_n\) with variance bounded by \(\sigma _v^2\) if we define \({{\bar{v}}} = \frac{1}{n}\sum _{i=1}^{n}v_i\) we have

$$\begin{aligned} {\mathbb {E}}\left[ \Vert {{\bar{v}}} - {\mathbb {E}}\left[ {{\bar{v}}}\right] \Vert ^2\right] = \frac{1}{n^2}\sum _{i=1}^{n}{\mathbb {E}}\left[ \Vert v_i-{\mathbb {E}}\left[ v_i\right] \Vert ^2\right] \le \frac{\sigma _v^2}{n}. \end{aligned}$$

\(\square\)

The following lemmas give the estimation bound of \(A_K\) and \(S_K\) in the stochastic case.

Lemma 38

In Algorithm 5, we have

$$\begin{aligned} {\mathbb {E}}\left[ S_{K}\right]&< \frac{\eta _x^2}{(1-\rho )^2}\sum _{j=0}^{K-1}\sum _{i=1}^{n}{\mathbb {E}}\left[ \Vert {\hat{\nabla }}f_i(x_{i,j}, y_{i,j}^{(T)};\phi _{i,j})\Vert ^2\right] \le \frac{\eta _x^2nK}{(1-\rho )^2}{\tilde{C}}_f^2, \end{aligned}$$

where the constant is defined as

$$\begin{aligned} {\tilde{C}}_f^2 = \left( L_{f,0} + \frac{LL_{f,1}}{\mu } +\frac{ LL_{f,1}}{\mu } \right) ^2 + {\tilde{\sigma }}_f^2 = {\mathcal {O}}(\kappa ^2). \end{aligned}$$

Proof

Observe that in this stochastic case, we can replace \({\hat{\nabla }}f_i(x_{i,j}, y_{i,j}^{(T)})\) with \({\hat{\nabla }}f_i(x_{i,j}, y_{i,j}^{(T)};\phi _{i,j})\) in Lemma 20 to get the first inequality. For the second inequality, we adopt the bound in Lemma 2 of [14]. \(\square\)

Lemma 39

Set parameters in Algorithm 5 as

$$\begin{aligned} \eta _y< \frac{2}{\mu + L},\quad \delta _y^{T}\le \frac{1}{3}. \end{aligned}$$
(55)

Then we have the following inequalities

$$\begin{aligned} {\mathbb {E}}\left[ A_K\right] \le \delta _y^{T}(2{\mathbb {E}}\left[ c_1\right] + 6\kappa ^2{\mathbb {E}}\left[ E_K\right] ) + \frac{\eta _y nK\sigma _{g,1}^2}{\mu },\ {\mathbb {E}}\left[ E_K\right] \le \frac{9n\eta _x^2K{\tilde{C}}_f^2}{(1-\rho )^2}. \end{aligned}$$

Proof

The proof is based on Lemma 10. Taking conditional expectation with respect to the filtration \({\mathcal {G}}_{i,j}^{(t-1)}\), we get

$$\begin{aligned}&{\mathbb {E}}\left[ \Vert y_{i,j}^{(t)} - y_i^*(x_{i,j})\Vert ^2\vert {\mathcal {G}}_{i,j}^{(t-1)}\right] \\&\quad = {\mathbb {E}}\left[ \Vert y_{i,j}^{(t-1)} - \eta _y\nabla _y g(x_{i,j}, y_{i,j}^{(t-1)};\xi _{i,j}^{(t-1)}) - y_i^*(x_{i,j}) \Vert ^2\vert {\mathcal {G}}_{i,j}^{(t-1)}\right] \\&\quad = \Vert y_{i,j}^{(t-1)} - \eta _y\nabla _y g(x_{i,j}, y_{i,j}^{(t-1)}) - y_i^*(x_{i,j})\Vert ^2 \\&\qquad + \eta _y^2{\mathbb {E}}\left[ \Vert \nabla _y g(x_{i,j}, y_{i,j}^{(t-1)}) - \nabla _y g(x_{i,j}, y_{i,j}^{(t-1)};\xi _{i,j}^{(t-1)})\Vert ^2\vert {\mathcal {G}}_{i,j}^{(t-1)}\right] \\&\quad \le (1 - \eta _y\mu )^2\Vert y_{i,j}^{(t-1)} - y_i^*(x_{i,j})\Vert ^2 + \eta _y^2\sigma _{g,1}^2, \end{aligned}$$

where the inequality uses Lemma 3. Taking expectation on both sides and using the tower property, we have

$$\begin{aligned}&{\mathbb {E}}\left[ \Vert y_{i,j}^{(T)} - y_i^*(x_{i,j})\Vert ^2\right] \nonumber \\&\quad \le (1 - \eta _y\mu )^2{\mathbb {E}}\left[ \Vert y_{i,j}^{(T-1)} - y_i^*(x_{i,j})\Vert ^2\right] + \eta _y^2\sigma _{g,1}^2 \nonumber \\&\quad \le (1 - \eta _y\mu )^{2T}{\mathbb {E}}\left[ \Vert y_{i,j}^{(0)} - y_i^*(x_{i,j})\Vert ^2\right] + \eta _y^2\sigma _{g,1}^2\sum _{s=0}^{T-1}(1 - \eta _y\mu )^{2s} \nonumber \\&\quad \le \delta _y^{T}{\mathbb {E}}\left[ \Vert y_{i,j}^{(0)} - y_i^*(x_{i,j})\Vert ^2\right] + \frac{\eta _y\sigma _{g,1}^2}{\mu }. \end{aligned}$$
(56)

Moreover, by the warm-start strategy, we have \(y_{i,j}^{(0)} = y_{i,j-1}^{(T)}\) and thus

$$\begin{aligned}&{\mathbb {E}}\left[ \Vert y_{i,j}^{(0)} - y_i^*(x_{i,j})\Vert ^2\right] \nonumber \\&\quad = {\mathbb {E}}\left[ \Vert y_{i,j-1}^{(T)} - y_i^*(x_{i,j-1}) + y_i^*(x_{i,j-1}) - y_i^*(x_{i,j})\Vert ^2\right] \nonumber \\&\quad \le 2{\mathbb {E}}\left[ \Vert y_{i,j-1}^{(T)} - y_i^*(x_{i,j-1})\Vert ^2\right] + 2{\mathbb {E}}\left[ \Vert y_i^*(x_{i,j-1}) - y_i^*(x_{i,j})\Vert ^2\right] \nonumber \\&\quad \le 2\delta _y^{T} {\mathbb {E}}\left[ \Vert y_{i,j-1}^{(0)} - y_i^*(x_{i,j-1})\Vert ^2\right] + 2\kappa ^2{\mathbb {E}}\left[ \Vert x_{i,j-1} - x_{i,j}\Vert ^2\right] \nonumber \\&\quad \le \frac{2}{3}{\mathbb {E}}\left[ \Vert y_{i,j-1}^{(0)} - y_i^*(x_{i,j-1})\Vert ^2\right] + 2\kappa ^2{\mathbb {E}}\left[ \Vert x_{i,j-1} - x_{i,j}\Vert ^2\right] , \end{aligned}$$
(57)

where the second inequality is by Lemma 7 and (57) and the last inequality is by (55). Taking summation over ij, we have:

$$\begin{aligned}&\sum _{j=1}^{K}\sum _{i=1}^{n}{\mathbb {E}}\left[ \Vert y_{i,j}^{(0)} - y_i^*(x_{i,j})\Vert ^2\right] \\&\quad \le \frac{2}{3} \sum _{j=1}^{K}\sum _{i=1}^{n}{\mathbb {E}}\left[ \Vert y_{i,j-1}^{(0)} - y_i^*(x_{i,j-1})\Vert ^2\right] + 2\kappa ^2{\mathbb {E}}\left[ E_K\right] \\&\quad \le \frac{2}{3}{\mathbb {E}}\left[ c_1\right] + \frac{2}{3}\sum _{j=1}^{K}\sum _{i=1}^{n}{\mathbb {E}}\left[ \Vert y_{i,j}^{(0)} - y_i^*(x_{i,j})\Vert ^2\right] + 2\kappa ^2{\mathbb {E}}\left[ E_K\right] , \end{aligned}$$

which leads to

$$\begin{aligned} \sum _{j=1}^{K}\sum _{i=1}^{n}{\mathbb {E}}\left[ \Vert y_{i,j}^{(0)} - y_i^*(x_{i,j})\Vert ^2\right] \le 2{\mathbb {E}}\left[ c_1\right] + 6\kappa ^2{\mathbb {E}}\left[ E_K\right] . \end{aligned}$$
(58)

Combining (58) with (56) and taking summation over ij, we have

$$\begin{aligned} {\mathbb {E}}\left[ A_K\right]&\le \delta _y^{T}\sum _{j=1}^{K}\sum _{i=1}^{n}{\mathbb {E}}\left[ \Vert y_{i,j}^{(0)} - y_i^*(x_{i,j})\Vert ^2\right] + \frac{\eta _y nK\sigma _{g,1}^2}{\mu } \\&\le \delta _y^{T}(2{\mathbb {E}}\left[ c_1\right] + 6\kappa ^2{\mathbb {E}}\left[ E_K\right] ) + \frac{\eta _y nK\sigma _{g,1}^2}{\mu }. \end{aligned}$$

Recall that for \(E_K\) we have:

$$\begin{aligned} E_K&= \sum _{j=1}^{K}\sum _{i=1}^{n}\Vert x_{i,j} - x_{i,j-1}\Vert ^2 \\&= \sum _{j=1}^{K}\sum _{i=1}^{n}\Vert x_{i,j} - {{\bar{x}}}_j + {{\bar{x}}}_j - {{\bar{x}}}_{j-1} + {{\bar{x}}}_{j-1} - x_{i,j-1}\Vert ^2 \\&= \sum _{j=1}^{K}\sum _{i=1}^{n}\Vert q_{i,j} -\eta _x\overline{\partial \Phi (X_{j-1};\phi )} - q_{i,j-1}\Vert ^2\\&\le 3\sum _{j=1}^{K}\sum _{i=1}^{n}(\Vert q_{i,j}\Vert ^2 + \eta _x^2\Vert \overline{\partial \Phi (X_{j-1};\phi )}\Vert ^2 + \Vert q_{i,j-1}\Vert ^2) \\&\le 3\sum _{j=1}^{K}(\Vert Q_j\Vert ^2 + \Vert Q_{j-1}\Vert ^2 + \eta _x^2\Vert \overline{\partial \Phi (X_{j-1};\phi )}\Vert ^2 \\&\le 6S_K + \frac{3\eta _x^2}{n}\sum _{j=0}^{K-1}\sum _{i=1}^{n}\Vert {\hat{\nabla }} f_i(x_{i,j}, y_{i,j}^{(T)};\phi _{i,j})\Vert ^2. \end{aligned}$$

Taking expectation on both sides yields

$$\begin{aligned} {\mathbb {E}}\left[ E_K\right]&\le 6{\mathbb {E}}\left[ S_K\right] + \frac{3\eta _x^2}{n}\sum _{j=0}^{K-1}\sum _{i=1}^{n}{\mathbb {E}}\left[ \Vert {\hat{\nabla }} f_i(x_{i,j}, y_{i,j}^{(T)};\phi _{i,j})\Vert ^2 \right] \\&\le \frac{6\eta _x^2nK}{(1-\rho )^2}{\tilde{C}}_f^2 + 3\eta _x^2K {\tilde{C}}_f^2 = \frac{9n\eta _x^2K{\tilde{C}}_f^2}{(1-\rho )^2},\\ \end{aligned}$$

which completes the proof. \(\square\)

Next, we prove the main convergence results in Theorem 33. Taking expectation on both sides in (48), we have:

$$\begin{aligned} \begin{aligned}&\frac{1}{K+1}\sum _{k=0}^{K}{\mathbb {E}}\left[ \Vert {\mathbb {E}}\left[ \overline{\partial \Phi (X_k; \phi )}\vert {\mathcal {F}}_k\right] - \nabla \Phi (\bar{x}_k)\Vert ^2\right] \\&\quad \le 3\left( L_{f,0}^2(1-\beta \mu )^{2M}\kappa ^2 + \frac{L_f^2}{n(K+1)}{\mathbb {E}}\left[ A_K\right] + \frac{L_{\Phi }^2}{n(K+1)}{\mathbb {E}}\left[ S_K\right] \right) \\&\quad \le C_3 + \frac{3\eta _y L_f^2 \sigma _{g,1}^2}{\mu } + \frac{3\eta _x^2L_{\Phi }^2 }{(1-\rho )^2}{\tilde{C}}_f^2, \end{aligned} \end{aligned}$$
(59)

where the constant is defined as:

$$\begin{aligned} C_3&= 3L_{f,0}^2(1-\beta \mu )^{2M}\kappa ^2 + \frac{3L_f^2}{n(K+1)}\delta _y^{T}(2{\mathbb {E}}\left[ c_1\right] + 6\kappa ^2{\mathbb {E}}\left[ E_K\right] ) \\&\le 3L_{f,0}^2(1-\beta \mu )^{2M}\kappa ^2 + \frac{3L_f^2}{n(K+1)}\delta _y^{T}\left( 2{\mathbb {E}}\left[ c_1\right] +\frac{54\kappa ^2n\eta _x^2K{\tilde{C}}_f^2}{(1-\rho )^2}\right) \\&=\Theta (\delta _{\beta }^{M}\kappa ^2 + \eta _x^2\delta _y^T\kappa ^8). \end{aligned}$$

Here we denote \(\delta _{\beta } = (1-\beta \mu )^2\) for simplicity. Therefore, we set \(M = \Theta (\log K)\) and \(T = \Theta (\log \kappa )\) such that \(C_3 = \Theta (\eta _x^2 + \frac{1}{K+1})\). Recall that (45) yields:

$$\begin{aligned}&{\mathbb {E}}\left[ \Vert \nabla \Phi ({{\bar{x}}}_k)\Vert ^2\right] \\&\quad \le \frac{2}{\eta _x}\left( {\mathbb {E}}\left[ \Phi ({{\bar{x}}}_k)\right] - {\mathbb {E}}\left[ \Phi ({{\bar{x}}}_{k+1})\right] \right) + {\mathbb {E}}\left[ \Vert {\mathbb {E}}\left[ \overline{\partial \Phi (X_k; \phi )}\vert {\mathcal {F}}_k\right] - \nabla \Phi ({{\bar{x}}}_k)\Vert ^2\right] \\&\qquad +L\eta _x{\mathbb {E}}\Vert \left[ \overline{\partial \Phi (X_k; \phi )}\right] - {\mathbb {E}}\left[ \overline{\partial \Phi (X_k; \phi )}\vert {\mathcal {F}}_k\right] \Vert ^2. \end{aligned}$$

Taking summation on both sides and using (59) and Lemma 37, we have

$$\begin{aligned} \frac{1}{K+1}\sum _{k=0}^{K}{\mathbb {E}}\left[ \Vert \nabla \Phi (\bar{x}_k)\Vert ^2\right] &\le \frac{2}{\eta _x(K+1)}({\mathbb {E}}\left[ \Phi (\bar{x}_0)\right] -\inf _x \Phi (x))+ \frac{3\eta _y L_f^2 \sigma _{g,1}^2}{\mu } \\&\quad + \frac{3\eta _x^2 L_{\Phi }^2 }{(1-\rho )^2}{\tilde{C}}_f^2 + \frac{L\eta _x{\tilde{\sigma }}_f^2}{n} + C_3. \end{aligned}$$

By setting

$$\begin{aligned} M = \Theta (\log K),\ T=\Theta (K^{\frac{1}{2}}),\ \eta _x = \Theta (K^{-\frac{1}{2}}),\ \eta _y = \Theta (K^{-\frac{1}{2}}) \end{aligned}$$

we know that the restrictions on algorithm parameters in Lemmas 35, 37, and 39 hold and we have

$$\begin{aligned} \frac{1}{K+1}\sum _{k=0}^{K}{\mathbb {E}}\left[ \Vert \nabla \Phi ({{\bar{x}}}_k)\Vert ^2\right] = {\mathcal {O}}\left( \frac{1}{\sqrt{K}}\right) , \end{aligned}$$

which proves the first case of Theorems 3.3 and 33.

1.3.2 Case 2: Assumption 2.3 does not hold

Lemma 40

Suppose the Assumption 2.3 does not hold in Algorithm 5, we have

$$\begin{aligned}&\frac{1}{K+1}\sum _{k=0}^{K}{\mathbb {E}}\left[ \Vert {\mathbb {E}}\left[ \overline{\partial \Phi (X_k; \phi )}\vert {\mathcal {F}}_k\right] - \nabla \Phi (\bar{x}_k)\Vert ^2\right] \\&\quad \le 12\left( 1 + L^2\kappa ^2 + \frac{2L_{f,0}^2L_{g,2}^2(1+\kappa ^2)}{\mu ^2}\right) \cdot \frac{C}{T} + 12CL_{f,0}^2\alpha ^N \\&\qquad + \left( \frac{36L_{f,0}^2L_{g,2}^2}{\mu ^2} + 2L_{f}^2\right) \frac{\eta _x^2(1+\kappa ^2)}{(1-\rho )^2}{\tilde{C}}_f^2. \end{aligned}$$

Proof

Denote by \({\hat{Z}}_{i,k}^{(N)}\) the output of each stochastic JHIP oracle 1 in Algorithm 5. Then

$$\begin{aligned} {\mathbb {E}}\left[ {\hat{Z}}_{i,k}^{(N)}\right] =Z_{i,k}^{(N)}, \end{aligned}$$

which implies

$$\begin{aligned} {\mathbb {E}}\left[ \overline{\partial \Phi (X_k; \phi )}\vert {\mathcal {F}}_k\right] = \overline{\partial \Phi (X_k)}. \end{aligned}$$

Hence we can follow the same process in case 2 of DBO to get (24) and thus

$$\begin{aligned}&\sum _{k=0}^{K}\Vert {\mathbb {E}}\left[ \overline{\partial \Phi (X_k; \phi )}\vert {\mathcal {F}}_k\right] - \nabla \Phi ({{\bar{x}}}_k)\Vert ^2 = \sum _{k=0}^{K}\Vert \overline{\partial \Phi (X_k)} - \nabla \Phi (\bar{x}_k)\Vert ^2 \\&\quad \le 12\left( 1 + L^2\kappa ^2 + \frac{2L_{f,0}^2L_{g,2}^2(1+\kappa ^2)}{\mu ^2}\right) \cdot \frac{1}{n}\sum _{k=0}^{K}\sum _{i=1}^{n}\Vert y_{i,k}^{(T)} - \tilde{y}_k^*\Vert ^2 \\&\qquad + \frac{12L_{f,0}^2}{n}\sum _{k=0}^{K}\sum _{i=1}^{n}\Vert Z_{i,k}^{(N)} - Z_k^*\Vert ^2+ \frac{(1+\kappa ^2)}{n}\cdot \left( \frac{36L_{f,0}^2L_{g,2}^2}{\mu ^2} + 2L_{f}^2\right) S_K. \\&\quad \le 12\left( 1 + L^2\kappa ^2 + \frac{2L_{f,0}^2L_{g,2}^2(1+\kappa ^2)}{\mu ^2}\right) \cdot \frac{(K+1)C}{T} + 12(K+1)CL_{f,0}^2\alpha ^N\\&\qquad + \frac{(1+\kappa ^2)}{n}\cdot \left( \frac{36L_{f,0}^2L_{g,2}^2}{\mu ^2} + 2L_{f}^2\right) S_K. \end{aligned}$$

The second inequality uses Lemmaa 14 and 15. Taking expectation, multiplying by \(\frac{1}{K+1}\), and using Lemma 38 we complete the proof. \(\square\)

The next lemma characterizes the variance of the gradient estimation.

Lemma 41

Suppose the Assumption 2.3 does not hold in Algorithm 5, then there exists \(\gamma _t = {\mathcal {O}}\left( \frac{1}{t}\right)\) such that

$$\begin{aligned} {\mathbb {E}}\Vert \overline{\partial \Phi (X_k; \phi )} - {\mathbb {E}}\left[ \overline{\partial \Phi (X_k; \phi )}\vert {\mathcal {F}}_k\right] \Vert ^2 \le 4\sigma _f^2(1+\kappa ^2) + (8L_{f,0}^2 + 4\sigma _f^2)\frac{C}{N}. \end{aligned}$$

Proof

Recall that we have:

$$\begin{aligned}&{\hat{\nabla }}f_i(x_{i,k}, y_{i,k}^{(T)};\phi ) = \nabla _x f_i(x_{i,k},y_{i,k}^{(T)};\phi _{i,k}^{(0)}) -\left[ {\hat{Z}}_{i,k}^{(N)}\right] ^{{\textsf{T}}}\nabla _y f_i(x_{i,k},y_{i,k}^{(T)};\phi _{i,k}^{(0)}) \\&{\hat{\nabla }}f_i(x_{i,k}, y_{i,k}^{(T)}) = \nabla _x f_i(x_{i,k},y_{i,k}^{(T)}) -\left( Z_{i,k}^{(N)}\right) ^{{\textsf{T}}}\nabla _y f_i(x_{i,k},y_{i,k}^{(T)}). \end{aligned}$$

By introducing intermediate terms we have

$$\begin{aligned}&{\hat{\nabla }}f_i(x_{i,k}, y_{i,k}^{(T)};\phi ) - {\hat{\nabla }}f_i(x_{i,k}, y_{i,k}^{(T)}) \\&\quad = \nabla _x f_i(x_{i,k},y_{i,k}^{(T)};\phi _{i,k}^{(0)}) - \nabla _x f_i(x_{i,k},y_{i,k}^{(T)}) - \left[ {\hat{Z}}_{i,k}^{(N)}\right] ^{{\textsf{T}}}\nabla _y f_i(x_{i,k},y_{i,k}^{(T)};\phi _{i,k}^{(0)}) \\&\qquad + \left( Z_k^*\right) ^{{\textsf{T}}}\nabla _y f_i(x_{i,k},y_{i,k}^{(T)};\phi _{i,k}^{(0)})- \left( Z_k^*\right) ^{{\textsf{T}}}\nabla _y f_i(x_{i,k},y_{i,k}^{(T)};\phi _{i,k}^{(0)}) \\&\qquad + \left( Z_k^*\right) ^{{\textsf{T}}}\nabla _y f_i(x_{i,k},y_{i,k}^{(T)}) - \left( Z_k^*\right) ^{{\textsf{T}}}\nabla _y f_i(x_{i,k},y_{i,k}^{(T)}) + \left( Z_{i,k}^{(N)}\right) ^{{\textsf{T}}}\nabla _y f_i(x_{i,k},y_{i,k}^{(T)}). \end{aligned}$$

Hence we know

$$\begin{aligned}&\Vert {\hat{\nabla }}f_i(x_{i,k}, y_{i,k}^{(T)};\phi ) - {\hat{\nabla }}f_i(x_{i,k}, y_{i,k}^{(T)})\Vert ^2 \\&\quad \le 4\Vert \nabla _x f_i(x_{i,k},y_{i,k}^{(T)};\phi _{i,k}^{(0)}) - \nabla _x f_i(x_{i,k},y_{i,k}^{(T)})\Vert ^2\\&\qquad +4\Vert \left( {\hat{Z}}_{i,k}^{(N)} - Z_k^* \right) ^{{\textsf{T}}}\nabla _y f_i(x_{i,k},y_{i,k}^{(T)};\phi _{i,k}^{(0)})\Vert ^2 \\&\qquad +4\Vert \left( Z_k^*\right) ^{{\textsf{T}}}(\nabla _y f_i(x_{i,k},y_{i,k}^{(T)};\phi _{i,k}^{(0)}) - \nabla _y f_i(x_{i,k},y_{i,k}^{(T)}))\Vert ^2\\&\qquad +4\Vert \left( Z_k^* - Z_{i,k}^{(N)}\right) ^{{\textsf{T}}}\nabla _y f_i(x_{i,k},y_{i,k}^{(T)})\Vert ^2. \end{aligned}$$

For the first term and the third term we use \({\mathbb {E}}\left[ \Vert \nabla f_i(x,y;\phi ) - \nabla f_i(x,y)\Vert ^2\right] \le \sigma _f^2\). For the second term (and the fourth term) we use the fact that stochastic (and deterministic) decentralized algorithm achieves sublinear rate (Lemma 15). Without loss of generality we can set C such that: \(\max \left( \frac{1}{n}\sum _{i=1}^{n}{\mathbb {E}}\left[ \Vert {\hat{Z}}_{i,k}^{(N)} - Z_k^*\Vert ^2\right] , \Vert Z_{i,k}^{(N)}-Z_k^*\Vert ^2\right) \le \frac{C}{N}\). For partial gradients in the second and fourth terms, we use Assumption 2.1 and the fact that

$$\begin{aligned} {\mathbb {E}}\left[ \Vert X\Vert ^2\right] = {\mathbb {E}}\left[ \Vert X - {\mathbb {E}}\left[ X\right] \Vert ^2\right] + \Vert {\mathbb {E}}\left[ X\right] \Vert ^2 \end{aligned}$$

for any random vector X. Taking summation and expectation on both sides, we have

$$\begin{aligned}&\frac{1}{n}\sum _{i=1}^{n}{\mathbb {E}}\left[ \Vert {\hat{\nabla }}f_i(x_{i,k}, y_{i,k}^{(T)};\phi ) - {\hat{\nabla }}f_i(x_{i,k}, y_{i,k}^{(T)})\Vert ^2\right] \\&\quad \le 4\sigma _f^2 + 4(L_{f,0}^2 + \sigma _f^2)\frac{C}{N} + \frac{4L^2}{\mu ^2}\sigma _f^2 + 4L_{f,0}^2\frac{C}{N}, \end{aligned}$$

which, together with

$$\begin{aligned}&{\mathbb {E}}\Vert \left[ \overline{\partial \Phi (X_k; \phi )}\right] - {\mathbb {E}}\left[ \overline{\partial \Phi (X_k; \phi )}\vert {\mathcal {F}}_k\right] \Vert ^2\\&\quad \le \frac{1}{n}\sum _{i=1}^{n}{\mathbb {E}}\left[ \Vert {\hat{\nabla }}f_i(x_{i,k}, y_{i,k}^{(T)};\phi ) - {\hat{\nabla }}f_i(x_{i,k}, y_{i,k}^{(T)})\Vert ^2\right] , \end{aligned}$$

proves the lemma. \(\square\)

Now we are ready to give the final proof. Taking summation on both sides of (45) and putting Lemma 40 and 41 together we know:

$$\begin{aligned}&\frac{1}{K+1}\sum _{k=0}^{K}{\mathbb {E}}\left[ \Vert \nabla \Phi ({{\bar{x}}}_k)\Vert ^2\right] \\&\quad \le \frac{1}{K+1}\big (\frac{2}{\eta _x}({\mathbb {E}}\left[ \Phi (\overline{x_{0}})\right] - \inf _x\Phi (x)) +\sum _{k=0}^{K}{\mathbb {E}}[\Vert {\mathbb {E}}\left[ \overline{\partial \Phi (X_k; \phi )}\vert {\mathcal {F}}_k\right] - \nabla \Phi (\bar{x}_k)\Vert ^2]\big ) \\&\qquad +\frac{L\eta _x}{K+1} \sum _{k=0}^{K}{\mathbb {E}}\Vert \left[ \overline{\partial \Phi (X_k; \phi )}\right] - {\mathbb {E}}\left[ \overline{\partial \Phi (X_k; \phi )}\vert {\mathcal {F}}_k\right] \Vert ^2 \\&\quad \le \frac{2}{\eta _x(K+1)}({\mathbb {E}}\left[ \Phi (\overline{x_{0}})\right] - \inf _x\Phi (x)) \\&\qquad +12\left( 1 + L^2\kappa ^2 + \frac{2L_{f,0}^2L_{g,2}^2(1+\kappa ^2)}{\mu ^2}\right) \cdot \frac{C}{T} + 12CL_{f,0}^2\alpha ^N \\&\qquad + \left( \frac{36L_{f,0}^2L_{g,2}^2}{\mu ^2} + 2L_{f}^2\right) \frac{\eta _x^2(1+\kappa ^2)}{(1-\rho )^2}{\tilde{C}}_f^2 +L\eta _x \left(4\sigma _f^2(1+\kappa ^2) + (8L_{f,0}^2 + 4\sigma _f^2)\frac{C}{N}\right) \\&\quad = \frac{2}{\eta _x(K+1)}({\mathbb {E}}\left[ \Phi (\overline{x_{0}})\right] - \inf _x\Phi (x)) + \left( \frac{36L_{f,0}^2L_{g,2}^2}{\mu ^2} + 2L_{f}^2\right) \frac{\eta _x^2(1+\kappa ^2)}{(1-\rho )^2}{\tilde{C}}_f^2 \\&\qquad + L\eta _x\left( 4\sigma _f^2(1+\kappa ^2) + (8L_{f,0}^2 + 4\sigma _f^2)\frac{C}{N}\right) + {\tilde{C}}_3, \end{aligned}$$

which completes the proof. Here the constant is defined as

$$\begin{aligned} {\tilde{C}}_3 = 12\left( 1 + L^2\kappa ^2 + \frac{2L_{f,0}^2L_{g,2}^2(1+\kappa ^2)}{\mu ^2}\right) \cdot \frac{C}{T} + 12CL_{f,0}^2\alpha ^N = {\mathcal {O}}\left( \frac{1}{T}+\alpha ^N\right) . \end{aligned}$$

By setting

$$\begin{aligned} N=\Theta (\log K),\ T=\Theta (K^{\frac{1}{2}}),\ \eta _x = \Theta (K^{-\frac{1}{2}}),\ \eta _y^{(t)} = {\mathcal {O}}\left( \frac{1}{t}\right) ,\ \end{aligned}$$

we have:

$$\begin{aligned} \frac{1}{K+1}\sum _{k=0}^{K}{\mathbb {E}}\left[ \Vert \nabla \Phi ({{\bar{x}}}_k)\Vert ^2\right] = {\mathcal {O}}\left( \frac{1}{\sqrt{K}}\right) , \end{aligned}$$

which proves the second case of Theorems 3.3 and 33.

Rights and permissions

Springer Nature or its licensor (e.g. a society or other partner) holds exclusive rights to this article under a publishing agreement with the author(s) or other rightsholder(s); author self-archiving of the accepted manuscript version of this article is solely governed by the terms of such publishing agreement and applicable law.

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Chen, X., Huang, M. & Ma, S. Decentralized bilevel optimization. Optim Lett (2024). https://doi.org/10.1007/s11590-024-02101-4

Download citation

  • Received:

  • Accepted:

  • Published:

  • DOI: https://doi.org/10.1007/s11590-024-02101-4

Keywords

Navigation