Existence, uniqueness, and convergence rates for gradient flows in the training of artificial neural networks with ReLU activation

Simon Eberle1, Arnulf Jentzen2,3, Adrian Riekert4, and Georg S. Weiss5

1 Faculty of Mathematics, AG Analysis of Partial Differential Equations, University of Duisburg-Essen, Germany, e-mail: simon.eberle○a uni-due.de 2 : Institute for Analysis and Numerics, University of M¨unster,Germany, e-mail: ajentzen○a uni-muenster.de 3 School of Data Science and Shenzhen Research Institute of Big Data, The Chinese University of Hong Kong, Shenzhen, China, e-mail: ajentzen○a cuhk.edu.cn 4 Applied Mathematics: Institute for Analysis and Numerics, University of M¨unster,Germany, e-mail: ariekert○a uni-muenster.de 5 Faculty of Mathematics, AG Analysis of Partial Differential Equations, University of Duisburg-Essen, Germany, e-mail: georg.weiss○a uni-due.de August 21, 2021

Abstract The training of artificial neural networks (ANNs) with rectified linear unit (ReLU) activation via gradient descent (GD) type optimization schemes is nowadays a common industrially relevant procedure which appears, for example, in the context of natural language processing, image processing, fraud detection, and game intelligence. Although there exist a large number of numerical simulations in which GD type optimization schemes are effectively used to train ANNs with ReLU activation, till this day in the scientific literature there is in general no mathematical convergence analysis which explains the success of GD type optimization schemes in the training of such ANNs. GD type optimization schemes can be regarded as temporal methods for the gradient flow (GF) differential equations associated to the considered optimization problem and, in view of this, it seems to be a natural direction of research to first aim to develop a mathematical convergence theory for time-continuous GF differential equations and, thereafter, to aim to extend such a time-continuous convergence theory to implementable time-discrete GD type optimization methods. In this article we establish two basic results for GF differential equations in the training of fully-connected feedforward ANNs with one hidden layer and ReLU activation. In the first main result of this article we establish in the training of such ANNs under the assumption that the probability distribution of the input data of the considered supervised learning problem is absolutely arXiv:2108.08106v1 [cs.LG] 18 Aug 2021 continuous with a bounded density function that every GF differential equation admits for every initial value a solution which is also unique among a suitable class of solutions. In the second main result of this article we prove in the training of such ANNs under the assumption that the target function and the density function of the probability distribution of the input data are piecewise that every non-divergent GF trajectory converges with an appropriate rate of convergence to a critical point and that the risk of the non-divergent GF trajectory converges with rate 1 to the risk of the critical point.

1 Contents

1 Introduction 2

2 Properties of the risk function and its generalized gradient function6 2.1 Mathematical description of artificial neural networks (ANNs) ...... 6 2.2 Differentiability properties of the risk function ...... 7 2.3 Local Lipschitz continuity of active neuron regions ...... 7 2.4 Local Lipschitz continuity properties for the generalized gradient function . . . . 10 2.5 Subdifferentials ...... 12

3 Existence and uniqueness properties for solutions of gradient flows (GFs) 13 3.1 Existence properties for solutions of GF differential equations ...... 13 3.2 Uniqueness properties for solutions of GF differential equations ...... 15 3.3 Existence and uniqueness properties for solutions of GF differential equations . . 16

4 Semialgebraic sets and functions 16 4.1 Semialgebraic sets and functions ...... 17 4.2 On the semialgebraic property of certain parametric ...... 17 4.3 On the semialgebraic property of the risk function ...... 21

5 Convergence rates for solutions of GF differential equations 22 5.1 GeneralizedLojasiewicz inequality for the risk function ...... 22 5.2 Local convergence for solutions of GF differential equations ...... 23 5.3 Global convergence for solutions of GF differential equations ...... 26

1 Introduction

The training of artificial neural networks (ANNs) with rectified linear unit (ReLU) activation via gradient descent (GD) type optimization schemes is nowadays a common industrially rel- evant procedure which appears, for instance, in the context of natural language processing, face recognition, fraud detection, and game intelligence. Although there exist a large number of numerical simulations in which GD type optimization schemes are effectively used to train ANNs with ReLU activation, till this day in the scientific literature there is in general no math- ematical convergence analysis which explains the success of GD type optimization schemes in the training of such ANNs. GD type optimization schemes can be regarded as temporal discretization methods for the gradient flow (GF) differential equations associated to the considered optimization problem and, in view of this, it seems to be a natural direction of research to first aim to develop a mathematical convergence theory for time-continuous GF differential equations and, thereafter, to aim to extend such a time-continuous convergence theory to implementable time-discrete GD type optimization methods. Although there is in general no theoretical analysis which explains the success of GD type optimization schemes in the training of ANNs in the literature, there are several auspicious analysis approaches as well as several promising partial error analyses regarding the training of ANNs via GD type optimization schemes and GFs, respectively, in the literature. For convex objective functions, the convergence of GF and GD processes to the global minimum in different settings has been proved, e.g., in [5, 23, 34, 35, 38]. For general non-convex objective functions, even under smoothness assumptions GF and GD processes can show wild oscillations and admit infinitely many points, cf., e.g., [1]. A standard condition which excludes this undesirable behavior is theLojasiewicz inequality and we point to [1,3,4,8, 16, 28, 29, 30, 31, 33, 36] for convergence results for GF and GD processes underLojasiewicz type assumptions. It is

2 in fact one of the main contributions of this work to demonstrate that the objective functions occurring in the training of ANNs with ReLU activation satisfy an appropriateLojasiewicz inequality, provided that both the target function and the density of the probability distribution of the input data are piecewise polynomial. For further abstract convergence results for GF and GD processes in the non-convex setting we refer, e.g., to [6, 20, 32, 37, 40] and the references mentioned therein. In the overparametrized regime, where the number of training parameters is much larger than the number of training data points, GF and GD processes can be shown to converge to global minima in the training of ANNs with high probability, cf., e.g., [2, 14, 17, 19, 21, 22, 41]. As the number of neurons increases to infinity, the corresponding GF processes converge (with appropriate rescaling) to a measure-valued process which is known in the scientific literature as Wasserstein gradient flow. For results on the convergence behavior of Wasserstein gradient flows in the training of ANNs we point, e.g., to [9], [12], [13], [18, Section 5.1], and the references mentioned therein. A different approach is to consider only very special target functions and we refer, in par- ticular, to [10, 25] for a convergence analysis for GF and GD processes in the case of constant target functions and to [26] for a convergence analysis for GF and GD processes in the training of ANNs with piecewise linear target functions. In the case of linear target functions, a complete characterization of the non-global local minima and the saddle points of the risk function has been obtained in [11]. In this article we establish two basic results for GF differential equations in the training of fully-connected feedforward ANNs with one hidden layer and ReLU activation. Specifically, in the first main result of this article, see Theorem 1.1 below, we establish in the training of such ANNs under the assumption that the probability distribution of the input data of the considered supervised learning problem is absolutely continuous with a bounded density function that every GF differential equation possesses for every initial value a solution which is also unique among a suitable class of solutions (see (1.4) in Theorem 1.1 for details). In the second main result of this article, see Theorem 1.2 below, we prove in the training of such ANNs under the assumption that the target function and the density function are piecewise polynomial (see (1.6) below for details) that every non-divergent GF trajectory converges with an appropriate speed of convergence (see (1.9) below) to a critical point. In Theorems 1.1 and 1.2 we consider ANNs with d ∈ N = {1, 2, 3,... } neurons on the input layer (d-dimensional input), H ∈ N neurons on the hidden layer (H-dimensional hidden layer), and 1 neuron on the output layer (1-dimensional output). There are thus Hd scalar real weight parameters and H scalar real bias parameters to describe the affine linear transformation between d-dimensional input layer and the H-dimensional hidden layer and there are thus H scalar real weight parameters and 1 scalar real bias parameter to describe the affine linear transformation between the H-dimensional hidden layer and the 1-dimensional output layer. Altogether there are thus d = Hd + H + H + 1 = Hd + 2H + 1 real numbers to describe the ANNs in Theorems 1.1 and 1.2. The real numbers a ∈ R, b ∈ (a, ∞) in Theorems 1.1 and 1.2 are used to specify the set [a, b]d in which the input data of the considered supervised learning problem takes values in d and the function f :[a, b] → R in Theorem 1.1 specifies the target function of the considered supervised learning problem. d In Theorem 1.1 we assume that the target function is an element of the set C([a, b] , R) of d continuous functions from [a, b] to R but beside this continuity hypothesis we do not impose further regularity assumptions on the target function. The function p:[a, b]d → [0, ∞) in Theorems 1.1 and 1.2 is an unnormalized density func- tion of the probability distribution of the input data of the considered supervised learning problem and in Theorem 1.1 we impose that this unnormalized density function is bounded and measurable.

3 In Theorems 1.1 and 1.2 we consider ANNs with the ReLU activation function R 3 x 7→ max{x, 0} ∈ R. The ReLU activation function fails to be differentiable and this lack of regu- larity also transfers to the risk function of the considered supervised learning problem; cf. (1.3) below. We thus need to employ appropriately generalized gradients of the risk function to specify the dynamics of the gradient flows. As in [25, Setting 2.1 and Proposition 2.3] (cf. also [10, 24]), we accomplish this, first, by approximating the ReLU activation function through continuously differentiable functions which converge pointwise to the ReLU activation function and whose derivatives converge pointwise to the left derivative of the ReLU activation function and, thereafter, by specifying the generalized gradient function as the limit of the gradients of the approximated risk functions; see (1.1) and (1.3) in Theorem 1.1 and (1.7) and (1.8) in Theorem 1.2 for details. We now present the precise statement of Theorem 1.1 and, thereafter, provide further com- ments regarding Theorem 1.2.

d Theorem 1.1. Let d, H, d ∈ N, a ∈ R, b ∈ (a, ∞), f ∈ C([a, b] , R) satisfy d = dH + 2H + 1, d let p:[a, b] → [0, ∞) be bounded and measurable, let Rr ∈ C(R, R), r ∈ N∪{∞}, satisfy for all x ∈ that (S {R }) ⊆ C1( , ), R (x) = max{x, 0}, sup sup |(R )0(y)| < ∞, R r∈N r R R ∞ r∈N y∈[−|x|,|x|] r and 0 1  lim supr→∞ |Rr(x) − R∞(x)| + |(Rr) (x) − (0,∞)(x)| = 0, (1.1) d θ for every θ = (θ1, . . . , θd) ∈ R let D ⊆ N satisfy θ  Pd D = i ∈ {1, 2,...,H}: |θHd+i| + j=1|θ(i−1)d+j| = 0 , (1.2)

d d let Lr : R → R, r ∈ N ∪ {∞}, satisfy for all r ∈ N ∪ {∞}, θ = (θ1, . . . , θd) ∈ R that Z Lr(θ) = f(x1, . . . , xd) [a,b]d PH  Pd 2 − θd − i=1 θH(d+1)+i Rr(θHd+i + j=1 θ(i−1)d+jxj) p(x) d(x1, . . . , xd), (1.3)

d d d d let θ ∈ R , and let G : R → R satisfy for all ϑ ∈ {v ∈ R : ((∇Lr)(v))r∈N is convergent} that G(ϑ) = limr→∞(∇Lr)(ϑ). Then (i) it holds that G is locally bounded and measurable and

d (ii) there exists a unique Θ ∈ C([0, ∞), R ) which satisfies for all t ∈ [0, ∞), s ∈ [t, ∞) that DΘt ⊆ DΘs and Z t Θt = θ − G(Θu) du. (1.4) 0 Theorem 1.1 is a direct consequence of Theorem 3.3 below. In Theorem 1.2 we also assume d that the target function f :[a, b] → R is continuous but additionally assume that, roughly d speaking, both the target function f :[a, b] → R and the unnormalized density function p:[a, b]d → [0, ∞) coincide with polynomial functions on suitable subsets of their domain of d k n×d definition [a, b] . In Theorem 1.2 the (n × d)-matrices αi ∈ R , i ∈ {1, 2, . . . , n}, k ∈ {0, 1}, k n and the n-dimensional vectors βi ∈ R , i ∈ {1, 2, . . . , n}, k ∈ {0, 1}, are used to describe these k d subsets and the functions Pi : R → R, i ∈ {1, 2, . . . , n}, k ∈ {0, 1}, constitute the with which the target function and the unnormalized density function should partially coincide. More formally, in (1.6) in Theorem 1.2 we assume that for every x ∈ [a, b]d we have that

P 0 P 1 p(x) = 0 0 n P (x) and f(x) = 1 1 n P (x). (1.5) i∈{1,2,...,n}, αi x+βi ∈[0,∞) i i∈{1,2,...,n}, αi x+βi ∈[0,∞) i In (1.9) in Theorem 1.2 we prove that there exists a strictly positive real number β ∈ (0, ∞) d such that for every GF trajectory Θ: [0, ∞) → R which does not diverge to infinity in the

4 1 d sense that lim inft→∞kΘtk < ∞ we have that Θt ∈ R , t ∈ [0, ∞), converges with order β −1 d to a critical point ϑ ∈ G ({0}) = {θ ∈ R : G(θ) = 0} and we have that the risk L(Θt) ∈ R, t ∈ [0, ∞), converges with order 1 to the risk L(ϑ) of the critical point ϑ. We now present the precise statement of Theorem 1.2.

d Theorem 1.2. Let d, H, d, n ∈ N, a ∈ R, b ∈ (a, ∞), f ∈ C([a, b] , R) satisfy d = dH+2H+1, k n×d k n k d for every i ∈ {1, 2, . . . , n}, k ∈ {0, 1} let αi ∈ R , let βi ∈ R , and let Pi : R → R be a polynomial, let p:[a, b]d → [0, ∞) satisfy for all k ∈ {0, 1}, x ∈ [a, b]d that

n P  k 1 n k k  kf(x) + (1 − k)p(x) = i=1 Pi (x) [0,∞) (αi x + βi ) , (1.6) let R ∈ C( , ), r ∈ ∪ {∞}, satisfy for all x ∈ that (S {R }) ⊆ C1( , ), R (x) = r R R N R r∈N r R R ∞ 0 max{x, 0}, supr∈N supy∈[−|x|,|x|]|(Rr) (y)| < ∞, and 0 1  lim supr→∞ |Rr(x) − R∞(x)| + |(Rr) (x) − (0,∞)(x)| = 0, (1.7)

d d let Lr : R → R, r ∈ N ∪ {∞}, satisfy for all r ∈ N ∪ {∞}, θ = (θ1, . . . , θd) ∈ R that Z Lr(θ) = f(x1, . . . , xd) [a,b]d PH  Pd 2 − θd − i=1 θH(d+1)+i Rr(θHd+i + j=1 θ(i−1)d+jxj) p(x) d(x1, . . . , xd), (1.8)

d d d let G : R → R satisfy for all θ ∈ {ϑ ∈ R : ((∇Lr)(ϑ))r∈N is convergent} that G(θ) = limr→∞ d (∇Lr)(θ), and let Θ ∈ C([0, ∞), R ) satisfy lim inft→∞kΘtk < ∞ and ∀ t ∈ [0, ∞):Θt = R t −1 Θ0 − 0 G(Θs) ds. Then there exist ϑ ∈ G ({0}), C, β ∈ (0, ∞) which satisfy for all t ∈ [0, ∞) that −β −1 kΘt − ϑk ≤ C(1 + t) and |L∞(Θt) − L∞(ϑ)| ≤ C(1 + t) . (1.9) Theorem 1.2 above is an immediate consequence of Theorem 5.4 in Subsection 5.3 below. Theorem 1.2 is related to Theorem 1.1 in our previous article [24]. In particular, [24, Theorem 1.1] uses weaker assumptions than Theorem 1.2 above but Theorem 1.2 above establishes a stronger statement when compared to [24, Theorem 1.1]. Specifically, on the one hand in [24, Theorem 1.1] the target function is only assumed to be a continuous function and the unnormalized density is only assumed to be measurable and integrable while in Theorem 1.2 it is additionally assumed that both the target function and the unnormalized density are piecewise polynomial in the sense of (1.6) above. On the other hand [24, Theorem 1.1] only asserts that the risk of every bounded GF trajectory converges to the risk of critical point while Theorem 1.2 assures that every non-divergent GF trajectory converges with a polynomial rate of convergence to a critical point and also assures that the risk of the non-divergent GF trajectory converges with rate 1 to the risk of the critical point. The remainder of this article is organized in the following way. In Section2 we establish several regularity properties for the risk function of the considered supervised learning problem and its generalized gradient function. In Section3 we employ the findings from Section2 to establish existence and uniqueness properties for solutions of GF differential equations. In par- ticular, in Section3 we present the proof of Theorem 1.1 above. In Section4 we establish under d the assumption that both the target function f :[a, b] → R and the unnormalized density function p:[a, b]d → [0, ∞) are piecewise polynomial that the risk function is semialgebraic in the sense of Definition 4.3 in Section4 (see Corollary 4.10 in Section4 for details). In Section5 we engage the results from Sections2 and4 to establish several convergence rate results for solutions of GF differential equations and, thereby, we also prove Theorem 1.2 above.

1 n n n Note that the functions k·k:(∪n∈NR ) → R and h·, ·i:(∪n∈N(R × R )) → R satisfy for all n ∈ N, x = n Pn 2 1/2 Pd (x1, . . . , xn), y = (y1, . . . , yn) ∈ R that kxk = [ i=1|xi| ] and hx, yi = i=1 xiyi.

5 2 Properties of the risk function and its generalized gradient function

d In this section we establish several regularity properties for the risk function L: R → R and its d d generalized gradient function G : R → R . In particular, in Proposition 2.12 in Subsection 2.5 d d dH+2H+1 below we prove for every parameter vector θ ∈ R in the ANN parameter space R = R d that the generalized gradient G(θ) is a limiting subdifferential of the risk function L: R → R at θ. In Definition 2.8 in Subsection 2.5 we recall the notion of subdifferentials (which are sometimes also referred to as Fr´echet subdifferentials in the scientific literature) and in Definition 2.9 in Subsection 2.5 we recall the notion of limiting subdifferentials. In the scientific literature Definitions 2.8 and 2.9 can in a slightly different presentational form, e.g., be found in Rockafellar & Wets [39, Definition 8.3] and Bolte et al. [8, Definition 2.10], respectively. Our proof of Proposition 2.12 uses the continuously differentiability result for the risk func- tion in Proposition 2.3 in Subsection 2.2 and the local Lipschitz continuity result for the gener- alized gradient function in Corollary 2.7 in Subsection 2.4. Corollary 2.7 will also be employed in Section3 below to establish existence and uniqueness results for solutions of GF differential equations. Proposition 2.3 follows directly from [24, Proposition 2.11, Lemma 2.12, and Lemma 2.13]. Our proof of Corollary 2.7, in turn, employs the known representation result for the generalized gradient function in Proposition 2.2 in Subsection 2.2 below and the local Lipschitz continuity result for certain parameter integals in Corollary 2.6 in Subsection 2.4. Statements related to Proposition 2.2 can, e.g., be found in [24, Proposition 2.2], [10, Proposition 2.3], and [25, Proposition 2.3]. Our proof of Corollary 2.6 uses the elementary abstract local Lipschitz continuity result for certain parameter integrals in Lemma 2.5 in Subsection 2.4 and the local Lipschitz continuity result for active neuron regions in Lemma 2.4 in Subsection 2.3 below. Lemma 2.4 is a gen- eralization of [26, Lemma 2.8], Lemma 2.5 is a slight generalization of [26, Lemma 2.7], and Corollary 2.6 is a generalization of [24, Lemma 2.13] and [26, Corollaries 2.10 and 2.11]. Only for completeness we include in this section a detailed proof for Lemma 2.5. In Setting 2.1 in Subsection 2.1 below we present the mathematical setup to describe ANNs with ReLU activa- d d d tion, the risk function L: R → R, and its generalized gradient function G : R → R . Moreover, d in (2.6) in Setting 2.1 we define for a given parameter vector θ ∈ R the set of hidden neurons which have all input parameters equal to zero. Such neurons are sometimes called degenerate (cf. [11]) and can cause problems with the differentiability of the risk function, which is why we exclude degenerate neurons in Proposition 2.3 and Corollary 2.7 below.

2.1 Mathematical description of artificial neural networks (ANNs) d Setting 2.1. Let d, H, d ∈ N, a ∈ R, b ∈ (a, ∞), f ∈ C([a, b] , R) satisfy d = dH + 2H + 1, θ d H×d θ θ d H let w = ((wi,j)(i,j)∈{1,...,H}×{1,...,d})θ∈Rd : R → R , b = ((b1,..., bH ))θ∈Rd : R → R , b = θ θ d H θ d d ((v1,..., vH ))θ∈Rd : R → R , and c = (c )θ∈Rd : R → R satisfy for all θ = (θ1, . . . , θd) ∈ R , i ∈ {1, 2,...,H}, j ∈ {1, 2, . . . , d} that

θ θ θ θ wi,j = θ(i−1)d+j, bi = θHd+i, vi = θH(d+1)+i, and c = θd, (2.1)

1 let Rr ∈ C (R, R), r ∈ N, satisfy for all x ∈ R that 0 1  lim supr→∞ |Rr(x) − max{x, 0}| + |(Rr) (x) − (0,∞)(x)| = 0 (2.2)

0 d and supr∈N supy∈[−|x|,|x|]|(Rr) (y)| < ∞, let λ: B(R ) → [0, ∞] be the Lebesgue–Borel measure d d θ d d on R , let p:[a, b] → [0, ∞) be bounded and measurable, let N = (N )θ∈Rd : R → C(R , R) d d d and L: R → R satisfy for all θ ∈ R , x = (x1, . . . , xd) ∈ R that θ θ PH θ  θ Pd θ N (x) = c + i=1 vi max bi + j=1 wi,jxj, 0 (2.3)

6 R θ 2 d and L(θ) = [a,b]d (f(y) − N (y)) p(y) λ(dy), let Lr : R → R, r ∈ N, satisfy for all r ∈ N, d θ ∈ R that Z 2  θ PH θ θ Pd θ  Lr(θ) = f(y) − c − i=1 vi Rr bi + j=1 wi,jyj p(y) λ(dy), (2.4) [a,b]d

d d d for every ε ∈ (0, ∞), θ ∈ R let Bε(θ) ⊆ R satisfy Bε(θ) = {ϑ ∈ R : kθ − ϑk < ε}, for every d θ d θ ∈ R , i ∈ {1, 2,...,H} let Ii ⊆ R satisfy

θ  d θ Pd θ Ii = x = (x1, . . . , xd) ∈ [a, b] : bi + j=1 wi,jxd > 0 , (2.5)

d θ for every θ ∈ R let D ⊆ N satisfy

θ  θ Pd θ D = i ∈ {1, 2,...,H}: |bi | + j=1|wi,j| = 0 , (2.6)

d d d and let G = (G1,..., Gd): R → R satisfy for all θ ∈ {ϑ ∈ R : ((∇Lr)(ϑ))r∈N is convergent} that G(θ) = limr→∞(∇Lr)(θ).

2.2 Differentiability properties of the risk function d Proposition 2.2. Assume Setting 2.1. Then it holds for all θ ∈ R , i ∈ {1, 2,...,H}, j ∈ {1, 2, . . . , d} that Z θ θ G(i−1)d+j(θ) = 2vi xj(N (x) − f(x))p(x) λ(dx), θ Ii Z θ θ GHd+i(θ) = 2vi (N (x) − f(x))p(x) λ(dx), θ Ii Z (2.7)   θ Pd θ  θ GH(d+1)+i(θ) = 2 max bi + j=1 wi,jxj, 0 (N (x) − f(x))p(x) λ(dx), [a,b]d Z θ and Gd(θ) = 2 (N (x) − f(x))p(x) λ(dx). [a,b]d

Proof of Proposition 2.2. Observe that, e.g., [24, Proposition 2.2] establishes (2.7). The proof of Proposition 2.2 is thus complete.

d  d θ Proposition 2.3. Assume Setting 2.1 and let U ⊆ R satisfy U = θ ∈ R : D = ∅ . Then d (i) it holds that U ⊆ R is open, 1 (ii) it holds that L|U ∈ C (U, R), and

(iii) it holds that ∇(L|U ) = G|U . Proof of Proposition 2.3. Note that [24, Proposition 2.11, Lemma 2.12, and Lemma 2.13] es- tablish items (i)–(iii). The proof of Proposition 2.3 is thus complete.

2.3 Local Lipschitz continuity of active neuron regions d+1 v d Lemma 2.4. Let d ∈ N, a ∈ R, b ∈ (a, ∞), for every v = (v1, . . . , vd+1) ∈ R let I ⊆ [a, b] v d Pd n satisfy I = {x ∈ [a, b] : vd+1 + i=1 vixi > 0}, for every n ∈ N let λn : B(R ) → [0, ∞] n d be the Lebesgue–Borel measure on R , let p:[a, b] → [0, ∞) be bounded and measurable, d+1 d+1 and let u ∈ R \{0}. Then there exist ε, C ∈ (0, ∞) such that for all v, w ∈ R with max{ku − vk, ku − wk} ≤ ε it holds that R Iv∆Iw p(x) λd(dx) ≤ Ckv − wk. (2.8)

7 d+1 Proof of Lemma 2.4. Observe that for all v, w ∈ R we have that

R  v w Iv∆Iw p(x) λd(dx) ≤ supx∈[a,b]d p(x) λd(I ∆I ). (2.9)

Moreover, note that the fact that for all y ∈ R it holds that y ≥ −|y| ensures that for all d+1 v = (v1, . . . , vd+1) ∈ R , i ∈ {1, 2, . . . , d + 1} with ku − vk < |ui| it holds that

2 2 2 uivi = (ui) + (vi − ui)ui ≥ |ui| − |ui − vi||ui| ≥ |ui| − ku − vk|ui| > 0. (2.10)

Next observe that for all v1, v2, w1, w2 ∈ R with min{|v1|, |w1|} > 0 it holds that h i v2 w2 |v2w1−w2v1| |v2(w1−v1)+v1(v2−w2)| |v2|+|v1|   − = = ≤ |v1 − w1| + |v2 − w2| . (2.11) v1 w1 |v1w1| |v1w1| |v1w1|

d+1 Combining this and (2.10) demonstrates for all v = (v1, . . . , vd+1), w = (w1, . . . , wd+1) ∈ R , i ∈ {1, 2, . . . , d} with max{kv − uk, kw − uk} < |u1| that v1w1 > 0 and h i h i vi − wi ≤ 2kvk [2kv − wk] ≤ 4kv−uk+4kuk kv − wk. (2.12) v1 w1 |v1w1| |v1w1|

d+1 Hence, we obtain for all v = (v1, . . . , vd+1), w = (w1, . . . , wd+1) ∈ R , i ∈ {1, 2, . . . , d} with |u1| max{kv − uk, kw − uk} ≤ 2 and |u1| > 0 that v1w1 > 0 and h i vi wi (2|u1|+4kuk)kv−wk 6kukkv−wk 24kuk − ≤ ≤ ≤ 2 kv − wk. (2.13) v1 w1 |u1+(v1−u1)||u1+(w1−u1)| (|u1|−kv−uk)(|u1|−kw−uk) |u1|

In the following we distinguish between the case maxi∈{1,2,...,d}|ui| = 0, the case (maxi∈{1,2,...,d}|ui|, d) ∈ (0, ∞) × [2, ∞), and the case (maxi∈{1,2,...,d}|ui|, d) ∈ (0, ∞) × {1}. We first prove (2.8) in the case maxi∈{1,2,...,d}|ui| = 0. (2.14) d+1 Note that (2.14) and the assumption that u ∈ R \{0} imply that |ud+1| > 0. Moreover, d+1 u v observe that (2.14) shows that for all v = (v1, . . . , vd+1) ∈ R , x = (x1, . . . , xd) ∈ I ∆I we have that Pd   Pd   vixi + vd+1 − uixi + ud+1 i=1 i=1 (2.15) Pd  Pd  Pd  = i=1 vixi + vd+1 + i=1 uixi + ud+1 ≥ i=1 uixi + ud+1 = |ud+1|.

d+1 d In addition, note that for all v = (v1, . . . , vd+1) ∈ R , x = (x1, . . . , xd) ∈ [a, b] it holds that Pd   Pd   Pd  vixi + vd+1 − uixi + ud+1 ≤ |vi − ui||xi| + |vd+1 − ud+1| i=1 i=1 i=1 (2.16) Pd  ≤ max{|a|, |b|} i=1|vi − ui| + |vd+1 − ud+1| ≤ (1 + d max{|a, b|})kv − uk.

d+1 |ud+1| u v This and (2.15) prove that for all v ∈ R with ku−vk ≤ 2+d max{|a,b|} we have that I ∆I = ∅, u v d+1 |ud+1| i.e., I = I . Therefore, we get for all v, w ∈ R with max{ku − vk, ku − wk} ≤ 2+d max{|a,b|} v w u d+1 that I = I = I . Hence, we obtain for all v, w ∈ R with max{ku − vk, ku − wk} ≤ |ud+1| v w 2+d max{|a,b|} that λd(I ∆I ) = 0. This establishes (2.8) in the case maxi∈{1,2,...,d}|ui| = 0. In the next step we prove (2.8) in the case

(maxi∈{1,2,...,d}|ui|, d) ∈ (0, ∞) × [2, ∞). (2.17)

v,w For this we assume without loss of generality that |u1| > 0. In the following let Jx ⊆ R, d−1 d+1 d−1 d+1 x ∈ [a, b] , v, w ∈ R , satisfy for all x = (x2, . . . , xd) ∈ [a, b] , v, w ∈ R that

8 v,w v w Jx = {y ∈ [a, b]:(y, x2, . . . , xd) ∈ I \I }. Next observe that Fubini’s theorem and the fact d+1 v d+1 that for all v ∈ R it holds that I is measurable show that for all v, w ∈ R we have that Z Z v w  λd(I ∆I ) = 1Iv∆Iw (x) λd(dx) = 1Iv\Iw (x) + 1Iw\Iv (x) λd(dx) [a,b]d [a,b]d Z Z  = 1Iv\Iw (y, x2, . . . , xd) + 1Iw\Iv (y, x2, . . . , xd) λ1(dy) λd−1(d(x2, . . . , xd)) [a,b]d−1 [a,b] Z Z 1 v,w 1 w,v  = Jx (y) + Jx (y) λ1(dy) λd−1(dx) [a,b]d−1 [a,b] Z v,w w,v = (λ1(Jx ) + λ1(Jx )) λd−1(dx). [a,b]d−1 (2.18)

d−1 Furthermore, note that for all x = (x2, . . . , xd) ∈ [a, b] , v = (v1, . . . , vd+1), w = (w1, . . . , wd+1) ∈ d+1 R , s ∈ {−1, 1} with min{sv1, sw1} > 0 it holds that v,w v w Jx = {y ∈ [a, b]:(y, x2, . . . , xd) ∈ I \I } n Pd  Pd  o = y ∈ [a, b]: v1y + i=2 vixi + vd+1 > 0 ≥ w1y + i=2 wixi + wd+1 (2.19) n o = y ∈ [a, b]: − s Pd v x  + v  < sy ≤ − s Pd w x  + w  . v1 i=2 i i d+1 w1 i=2 i i d+1

d−1 Hence, we obtain for all x = (x2, . . . , xd) ∈ [a, b] , v = (v1, . . . , vd+1), w = (w1, . . . , wd+1) ∈ d+1 R , s ∈ {−1, 1} with min{sv1, sw1} > 0 that

v,w s Pd   s Pd   λ1(J ) ≤ vixi + vd+1 − wixi + wd+1 x v1 i=2 w1 i=2 h i Pd vi wi vd+1 wd+1 ≤ − |xi| + − (2.20) i=2 v1 w1 v1 w1 h i ≤ max{|a|, |b|} Pd vi − wi + vd+1 − wd+1 . i=2 v1 w1 v1 w1

d+1 Furthermore, observe that (2.10) demonstrates for all v = (v1, . . . , vd+1) ∈ R with ku − vk < d+1 |u1| that u1v1 > 0. This implies that for all v = (v1, . . . , vd+1), w = (w1, . . . , wd+1) ∈ R with max{ku − vk, ku − wk} < |u1| there exists s ∈ {−1, 1} such that min{sv1, sw1} > 0. Combining d−1 this and (2.13) with (2.20) proves that there exists C ∈ R such that for all x ∈ [a, b] , d+1 |u1| v,w w,v v, w ∈ R with max{ku − vk, ku − wk} ≤ 2 we have that λ1(Jx ) + λ1(Jx ) ≤ Ckv − wk. This, (2.18), and (2.9) establish (2.8) in the case (maxi∈{1,2,...,d}|ui|, d) ∈ (0, ∞)×[2, ∞). Finally, we prove (2.8) in the case

(maxi∈{1,2,...,d}|ui|, d) ∈ (0, ∞) × {1}. (2.21)

Note that (2.21) demonstrates that |u1| > 0. In addition, observe that for all v = (v1, v2), 2 w = (w1, w2) ∈ R , s ∈ {−1, 1} with min{sv1, sw1} > 0 it holds that n o Iv\Iw = {y ∈ [a, b]: v y + v > 0 ≥ w y + w } = y ∈ [a, b]: − sv2 < sy ≤ − sw2 1 2 1 2 v1 w1 n o (2.22) ⊆ y ∈ : − sv2 < sy ≤ − sw2 . R v1 w1

2 Therefore, we get for all v = (v1, v2), w = (w1, w2) ∈ R , s ∈ {−1, 1} with min{sv1, sw1} > 0 that     v w sv2 sw2 v2 w2 λ1(I \I ) ≤ − − − = − . (2.23) v1 w1 v1 w1 2 Furthermore, note that (2.10) ensures for all v = (v1, v2) ∈ R with ku−vk < |u1| that u1v1 > 0. 2 This proves that for all v = (v1, v2), w = (w1, w2) ∈ R with max{ku−vk, ku−wk} < |u1| there

9 exists s ∈ {−1, 1} such that min{sv1, sw1} > 0. Combining this with (2.23) demonstrates for 2 all v = (v1, v2), w = (w1, w2) ∈ R with max{ku − vk, ku − wk} < |u1| that min{|v1|, |w1|} > 0 and v w v w w v v2 w2 λ1(I ∆I ) = λ1(I \I ) + λ1(I \I ) ≤ 2 − . (2.24) v1 w1

This, (2.13), and (2.9) establish (2.8) in the case (maxi∈{1,2,...,d}|ui|, d) ∈ (0, ∞) × {1}. The proof of Lemma 2.4 is thus complete.

2.4 Local Lipschitz continuity properties for the generalized gradient func- tion n n d Lemma 2.5. Let d, n ∈ N, a ∈ R, b ∈ (a, ∞), x ∈ R , C, ε ∈ (0, ∞), let φ: R × [a, b] → R be locally bounded and measurable, assume for all r ∈ (0, ∞) that

|φ(y,s)−φ(z,s)| supy,z∈Rn, kyk+kzk≤r, y6=z sups∈[a,b]d ky−zk < ∞, (2.25)

d y d n let µ: B([a, b] ) → [0, ∞) be a finite measure, let I ∈ B([a, b] ), y ∈ R , satisfy for all n y z n y, z ∈ {v ∈ R : kx − vk ≤ ε} that µ(I ∆I ) ≤ Cky − zk, and let Φ: R → R satisfy for all y ∈ n that R Z Φ(y) = φ(y, s) µ(ds). (2.26) Iy n Then there exists C ∈ R such that for all y, z ∈ {v ∈ R : kx − vk ≤ ε} it holds that |Φ(y) − Φ(z)| ≤ C ky − zk. Proof of Lemma 2.5. Observe that (2.25) and the assumption that φ is locally bounded ensure n d that there exists C ∈ R which satisfies for all y, z ∈ {v ∈ R : kx − vk ≤ ε}, s ∈ [a, b] with y 6= z that |φ(y,s)−φ(z,s)| ky−zk + |φ(y, s)| + |φ(z, s)| ≤ C . (2.27) n Next note that (2.26) shows for all y, z ∈ R that Z Z Z |Φ(y)−Φ(z)| ≤ |φ(y, s)−φ(z, s)| µ(ds)+ |φ(y, s)| µ(ds)+ |φ(z, s)| µ(ds). (2.28) Iy∩Iz Iy\Iz Iz\Iy

n Moreover, observe that (2.27) assures for all y, z ∈ {v ∈ R : kx − vk ≤ ε} that Z |φ(y, s) − φ(z, s)| µ(ds) ≤ C ky − zkµ([a, b]d). (2.29) Iy∩Iz n In the next step we combine (2.27) with the assumption that for all y, z ∈ {v ∈ R : kx−vk ≤ ε} y z n it holds that µ(I ∆I ) ≤ Cky − zk to obtain that for all y, z ∈ {v ∈ R : kx − vk ≤ ε} it holds that Z Z |φ(y, s)| µ(ds) + |φ(z, s)| µ(ds) ≤ CC ky − zk. (2.30) Iy\Iz Iz\Iy n This, (2.28), and (2.29) demonstrate for all y, z ∈ {v ∈ R : kx − vk ≤ ε} that |Φ(y) − Φ(z)| ≤ C (C + µ([a, b]d))ky − zk. (2.31)

The proof of Lemma 2.5 is thus complete.

d d Corollary 2.6. Assume Setting 2.1, let φ: R ×[a, b] → R be locally bounded and measurable, and assume for all r ∈ (0, ∞) that

|φ(θ,x)−φ(ϑ,x)| supθ,ϑ∈Rd, kθk+kϑk≤r, θ6=ϑ supx∈[a,b]d kθ−ϑk < ∞. (2.32) Then

10 (i) it holds that Z d R 3 θ 7→ φ(θ, x)p(x) λ(dx) ∈ R (2.33) [a,b]d is locally Lipschitz continuous and

(ii) it holds for all i ∈ {1, 2,...,H} that Z  d ϑ ϑ ∈ R : i∈ / D 3 θ 7→ φ(θ, x)p(x) λ(dx) ∈ R (2.34) θ Ii is locally Lipschitz continuous.

d Proof of Corollary 2.6. First note that Lemma 2.5 (applied for every θ ∈ R with n x d, x x θ, d R y d µ x (B([a, b] ) 3 A 7→ A p(x) λ( dx) ∈ [0, ∞)), (I )y∈Rn x ([a, b] )y∈Rd in the notation of d ϑ Lemma 2.5) establishes item (i). In the following let i ∈ {1, 2,...,H}, θ ∈ {ϑ ∈ R : i∈ / D }. d Observe that Lemma 2.4 shows that there exist ε, C ∈ (0, ∞) which satisfy for all ϑ1, ϑ2 ∈ R with max{kθ − ϑ1k, kθ − ϑ2k} ≤ ε that R ϑ1 ϑ2 p(x) λ(dx) ≤ Ckϑ1 − ϑ2k. (2.35) Ii ∆Ii

d d Combining this with Lemma 2.5 (applied for every θ ∈ R with n x d, x x θ, µ x (B([a, b] ) 3 R y y A 7→ A p(x) λ( dx) ∈ [0, ∞)), (I )y∈Rn x (Ii )y∈Rd in the notation of Lemma 2.5) demonstrates d that there exists C ∈ R such that for all ϑ1, ϑ2 ∈ R with max{kθ − ϑ1k, kθ − ϑ2k} ≤ ε it holds that Z Z φ(ϑ1, x)p(x) λ(dx) − φ(ϑ2, x)p(x) λ(dx) ≤ C kϑ1 − ϑ2k. (2.36) ϑ1 ϑ2 Ii Ii This establishes item (ii). The proof of Corollary 2.6 is thus complete.

Corollary 2.7. Assume Setting 2.1. Then

(i) it holds for all k ∈ N ∩ (Hd + H, d] that

d R 3 θ 7→ Gk(θ) ∈ R (2.37)

is locally Lipschitz continuous,

(ii) it holds for all i ∈ {1, 2,...,H}, j ∈ {1, 2, . . . , d} that

 d ϑ ϑ ∈ R : i∈ / D 3 θ 7→ G(i−1)d+j(θ) ∈ R (2.38)

is locally Lipschitz continuous, and

(iii) it holds for all i ∈ {1, 2,...,H} that

 d ϑ ϑ ∈ R : i∈ / D 3 θ 7→ GHd+i(θ) ∈ R (2.39)

is locally Lipschitz continuous.

Proof of Corollary 2.7. Note that (2.7) and Corollary 2.6 establish items (i)–(iii). The proof of Corollary 2.7 is thus complete.

11 2.5 Subdifferentials n n Definition 2.8 (Subdifferential). Let n ∈ N, f ∈ C(R , R), x ∈ R . Then we denote by n ∂fˆ (x) ⊆ R the set given by     n f(x + h) − f(x) − hy, hi ∂fˆ (x) = y ∈ R : lim inf ≥ 0 . (2.40) Rn\{0}3h→0 khk

n n Definition 2.9 (Limiting subdifferential). Let n ∈ N, f ∈ C(R , R), x ∈ R . Then we denote n by ∂f(x) ⊆ R the set given by

T hS i ∂f(x) = n ∂fˆ (y) (2.41) ε∈(0,∞) y∈{z∈R : kx−zk<ε}

(cf. Definition 2.8).

n n Lemma 2.10. Let n ∈ N, f ∈ C(R , R), x ∈ R . Then

 n n n  ˆ  ∂f(x) = y ∈ R : ∃ z = (z1, z2): N → R × R : ∀ k ∈ N: z2(k) ∈ ∂f(z1(k)) ,   lim supk→∞(kz1(k) − xk + kz2(k) − yk) = 0 (2.42)

(cf. Definitions 2.8 and 2.9).

Proof of Lemma 2.10. Observe that (2.41) establishes (2.42). The proof of Lemma 2.10 is thus complete.

n n 1 Lemma 2.11. Let n ∈ N, f ∈ C(R , R), let U ⊆ R be open, assume f|U ∈ C (U, R), and let x ∈ U. Then ∂fˆ (x) = ∂f(x) = {(∇f)(x)} (cf. Definitions 2.8 and 2.9).

Proof of Lemma 2.11. This is a direct consequence of, e.g., Rockafellar & Wets [39, Exercise 8.8]. The proof of Lemma 2.11 is thus complete.

d Proposition 2.12. Assume Setting 2.1 and let θ ∈ R . Then G(θ) ∈ ∂L(θ) (cf. Definition 2.9). d Proof of Proposition 2.12. Throughout this proof let ϑ = (ϑn)n∈N : N → R satisfy for all n ∈ N, ϑ ϑ ϑ n θ n θ 1 1 θ n θ ϑn θ i ∈ {1, 2,...,H}, j ∈ {1, 2, . . . , d} that wi,j = wi,j, bi = bi − n D (i), vi = vi , and c = c . We prove Proposition 2.12 through an application of Lemma 2.10. Note that for all n ∈ N, θ ϑn θ θ i ∈ {1, 2,...,H}\D it holds that bi = bi . This implies for all n ∈ N, i ∈ {1, 2,...,H}\D that i∈ / Dϑn . (2.43)

θ ϑn 1 In addition, observe that for all n ∈ N, i ∈ D it holds that bi = − n < 0. This shows for all θ n ∈ N, i ∈ D that i∈ / Dϑn . (2.44) ϑ Hence, we obtain for all n ∈ N that D n = ∅. Combining this with Proposition 2.3 and ˆ Lemma 2.11 demonstrates that for all n ∈ N it holds that ∂L(ϑn) = {(∇L)(ϑn)} = {G(ϑn)} (cf. Definition 2.8). Moreover, note that limn→∞ ϑn = θ. It thus remains to show that G(ϑn), n ∈ N, converges to G(θ). Observe that Corollary 2.7 ensures that for all k ∈ N ∩ (Hd + H, d] it holds that limn→∞ Gk(ϑn) = Gk(θ). (2.45) Furthermore, note that Corollary 2.7,(2.43), and (2.44) assure that for all i ∈ {1, 2,...,H}\Dθ, j ∈ {1, 2, . . . , d} it holds that

limn→∞ G(i−1)d+j(ϑn) = G(i−1)d+j(θ) and limn→∞ GHd+i(ϑn) = GHd+i(θ). (2.46)

12 θ ϑn θ In addition, observe that for all n ∈ N, i ∈ D we have that Ii = Ii = ∅. Hence, we obtain for all i ∈ Dθ, j ∈ {1, 2, . . . , d} that

limn→∞ G(i−1)d+j(ϑn) = 0 = G(i−1)d+j(θ) and limn→∞ GHd+i(ϑn) = 0 = GHd+i(θ). (2.47) Combining this, (2.45), and (2.46) demonstrates that limn→∞ G(ϑn) = G(θ). This and Lemma 2.10 assure that G(θ) ∈ ∂L(θ). The proof of Proposition 2.12 is thus complete.

3 Existence and uniqueness properties for solutions of gradient flows (GFs)

In this section we employ the local Lipschitz continuity result for the generalized gradient func- tion in Corollary 2.7 from Section2 to establish existence and uniqueness results for solutions of GF differential equations. Specifically, in Proposition 3.1 in Subsection 3.1 below we prove the existence of solutions GF differential equations, in Lemma 3.2 in Subsection 3.2 below we establish the uniqueness of solutions of GF differential equations among a suitable class of GF so- lutions, and in Theorem 3.3 in Subsection 3.3 below we combine Proposition 3.1 and Lemma 3.2 to establish the unique existence of solutions of GF differential equations among a suitable class of GF solutions. Theorem 1.1 in the introduction is an immediate consequence of Theorem 3.3. Roughly speaking, we show in Theorem 3.3 the unique existence of solutions of GF differ- ential equations among the class of GF solutions which satisfy that the set of all degenerate neurons of the GF solution at time t ∈ [0, ∞) is non-decreasing in the time variable t ∈ [0, ∞). In other words, in Theorem 3.3 we prove the unique existence of GF solutions with the property that once a neuron has become degenerate it will remain degenerate for subsequent times. Our strategy of the proof of Theorem 3.3 and Proposition 3.1, respectively, can, loosely speaking, be described as follows. Corollary 2.7 above implies that the components of the d d generalized gradient function G : R → R corresponding to non-degenerate neurons are locally Lipschitz continuous so that the classical Picard-Lindel¨oflocal existence and uniqueness theorem for ordinary differential equations can be brought into play for those components. On the other hand, if at some time t ∈ [0, ∞) the i-th neuron is degenerate, then Proposition 2.2 above d d shows that the corresponding components of the generalized gradient function G : R → R vanish. The GF differential equation is thus satisfied if the neuron remains degenerate at all subsequent times s ∈ [t, ∞). Using these arguments we prove in Proposition 3.1 the existence of GF solutions by induction on the number of non-degenerate neurons of the initial value.

3.1 Existence properties for solutions of GF differential equations d d Proposition 3.1. Assume Setting 2.1 and let θ ∈ R . Then there exists Θ ∈ C([0, ∞), R ) which satisfies for all t ∈ [0, ∞), s ∈ [0, ∞) that Z t Θt Θs Θt = θ − G(Θu) du and D ⊆ D . (3.1) 0 Proof of Proposition 3.1. We prove the statement by induction on the quantity H − #(Dθ) ∈ θ θ N ∩ [0,H]. Assume first that H − #(D ) = 0, i.e., D = {1, 2,...,H}. Note that this implies θ θ that w = 0 and b = 0. In the following let κ ∈ R satisfy Z κ = f(x)p(x) λ(dx). (3.2) [a,b]d Observe that the Picard–Lindel¨ofTheorem shows that there exists a unique c ∈ C([0, ∞), R) which satisfies for all t ∈ [0, ∞) that Z !Z t  c(0) = cθ and c(t) = c(0) + 2κt − 2 p(x) λ(dx) c(s) ds . (3.3) [a,b]d 0

13 d Next let Θ ∈ C([0, ∞), R ) satisfy for all t ∈ [0, ∞), i ∈ {1, 2,...,H}, j ∈ {1, 2, . . . , d} that

Θt θ Θt θ Θt θ Θt wi,j = wi,j = bi = bi = 0, vi = vi , and c = c(t). (3.4) Note that (2.7), (3.3), and (3.4) ensure for all t ∈ [0, ∞) that

Z !Z t  cΘt = cθ + 2κt − 2 p(x) λ(dx) cΘs ds [a,b]d 0 Z t Z ! = cθ − 2 −κ + cΘs p(x) λ(dx) ds 0 [a,b]d Z t Z   θ Θs PH  Θs  Θs Pd Θs  = c − 2 c + i=1 vi max bi + j=1 wi,j xj, 0 − f(x) p(x) λ(dx) ds 0 [a,b]d Z t Z Z t θ Θs θ = c − 2 (N (x) − f(x))p(x) λ(dx) ds = c − Gd(Θs) ds. 0 [a,b]d 0 (3.5)

Θ Next observe that (3.4) and (2.7) show for all t ∈ [0, ∞), i ∈ N ∩ [1, d) that D t = {1, 2,...,H} and Gi(Θt) = 0. Combining this with (3.4) and (3.5) proves that Θ satisfies (3.1). This establishes the claim in the case #(Dθ) = H. θ d For the induction step assume that #(D ) < H and assume that for all ϑ ∈ R with ϑ θ d #(D ) > #(D ) there exists Θ ∈ C([0, ∞), R ) which satisfies for all t ∈ [0, ∞), s ∈ [0, ∞) that R t Θt Θs d Θt = ϑ − 0 G(Θu) du and D ⊆ D . In the following let U ⊆ R satisfy  d ϑ θ U = ϑ ∈ R : D ⊆ D (3.6)

d and let G: U → R satisfy for all ϑ ∈ U, i ∈ {1, 2,..., d} that

( θ θ 0 : i ∈ {(` − 1)d + j : ` ∈ D , j ∈ N ∩ [1, d]} ∪ {Hd + `: ` ∈ D } Gi(ϑ) = (3.7) Gi(ϑ) : else.

d Note that (3.6) assures that U ⊆ R is open. In addition, observe that Corollary 2.7 implies that G is locally Lipschitz continuous. Combining this with the Picard–Lindel¨ofTheorem demonstrates that there exist a unique maximal τ ∈ (0, ∞] and Ψ ∈ C([0, τ),U) which satisfy for all t ∈ [0, τ) that Z t Ψt = θ − G(Ψu) du. (3.8) 0 θ Next note that the fact that for all ϑ ∈ U, i ∈ {(`−1)d+j : ` ∈ D , j ∈ N∩[1, d]}∪{Hd+`: ` ∈ θ θ D } it holds that Gi(ϑ) = 0 ensures that for all t ∈ [0, τ), i ∈ D , j ∈ {1, 2, . . . , d} we have that

Ψt θ Ψt θ Ψt θ wi,j = wi,j = bi = bi = 0 and vi = vi . (3.9)

This, (3.7), and (2.7) demonstrate for all t ∈ [0, τ) that G(Ψt) = G(Ψt). In addition, observe that (3.6) and (3.9) imply for all t ∈ [0, τ) that DΨt = Dθ. Hence, if τ = ∞ then Ψ satisfies (3.1). Next assume that τ < ∞. Note that the Cauchy-Schwarz inequality and [24, Lemma 3.1] prove for all s, t ∈ [0, τ) with s ≤ t that

Z t Z t 1/2 1/2 2 kΨt − Ψsk ≤ kG(Ψu)k du ≤ (t − s) kG(Ψu)k du s s Z t 1/2 1/2 2 1/2 1/2 (3.10) ≤ (t − s) kG(Ψu)k du = (t − s) L(Ψ0) − L(Ψt) 0 1/2 1/2 ≤ (t − s) L(Ψ0) .

14 Hence, we obtain for all (tn)n∈N ⊆ [0, τ) with lim infn→∞ tn = τ that (Ψtn ) is a Cauchy . d This implies that ϑ := limt↑τ Ψt ∈ R exists. Furthermore, observe that the fact that τ is ϑ θ maximal proves that ϑ∈ / U. Therefore, we have that D \D 6= ∅. Moreover, note that (3.9) θ ϑ ϑ ϑ shows that for all i ∈ D , j ∈ {1, 2, . . . , d} it holds that wi,j = bi = 0 and, therefore, i ∈ D . This demonstrates that #(Dϑ) > #(Dθ). Combining this with the induction hypothesis ensures d that there exists Φ ∈ C([0, ∞), R ) which satisfies for all t ∈ [0, ∞), s ∈ [0, ∞) that Z t Φt Φs Φt = ϑ − G(Φu) du and D ⊆ D . (3.11) 0 d In the following let Θ: [0, ∞) → R satisfy for all t ∈ [0, ∞) that ( Ψt : t ∈ [0, τ) Θt = (3.12) Φt−τ : t ∈ [τ, ∞).

Observe that the fact that ϑ = limt↑τ Ψt and the fact that Φ0 = ϑ imply that Θ is continuous. Furthermore, note that the fact that G is locally bounded and (3.8) ensure that  Z t  Z τ Z τ Θτ = ϑ = lim Ψt = lim θ − G(Ψs) ds = θ − G(Ψs) ds = θ − G(Θs) ds. (3.13) t↑τ t↑τ 0 0 0 Hence, we obtain for all t ∈ [τ, ∞) that Z t−τ Z τ Θt = (Θt − Θτ ) + Θτ = (Φt−τ − Φ0) + Θτ = − G(Φs) ds + θ − G(Θs) ds 0 0 (3.14) Z τ Z τ Z t = − G(Θs) + θ − G(Θs) ds = θ − G(Θs) ds. t 0 0 This shows that Θ satisfies (3.1). The proof of Proposition 3.1 is thus complete.

3.2 Uniqueness properties for solutions of GF differential equations d 1 2 d Lemma 3.2. Assume Setting 2.1 and let θ ∈ R , Θ , Θ ∈ C([0, ∞), R ) satisfy for all t ∈ [0, ∞), s ∈ [t, ∞), k ∈ {1, 2} that

t Z k k k k Θt Θs Θt = θ − G(Θu) du and D ⊆ D . (3.15) 0 1 2 Then it holds for all t ∈ [0, ∞) that Θt = Θt . Proof of Lemma 3.2. Assume for the sake of contradiction that there exists t ∈ [0, ∞) such that 1 2 Θt 6= Θt . By translating the variable t if necessary, we may assume without loss of generality  1 2 1 2 that inf t ∈ [0, ∞):Θt 6= Θt = 0. Next observe that the fact that Θ and Θ are continuous k implies that there exists δ ∈ (0, ∞) which satisfies for all t ∈ [0, δ], k ∈ {1, 2} that DΘt ⊆ Dθ. k Furthermore, note that (3.15) ensures for all t ∈ [0, ∞), i ∈ Dθ, k ∈ {1, 2} that i ∈ DΘt . Hence, we obtain for all t ∈ [0, ∞), i ∈ Dθ, j ∈ {1, 2, . . . , d}, k ∈ {1, 2} that

k k k G(i−1)d+j(Θt ) = GHd+i(Θt ) = GH(d+1)+i(Θt ) = 0. (3.16) In addition, observe that the fact that Θ1 and Θ2 are continuous implies that there exists a d ϑ θ k compact K ⊆ {ϑ ∈ R : D ⊆ D } which satisfies for all t ∈ [0, δ], k ∈ {1, 2} that Θt ∈ K. Moreover, note that Corollary 2.7 proves that for all i ∈ {1, 2,...,H}\Dθ, j ∈ {1, 2, . . . , d} it holds that G(i−1)d+j, GHd+i, GH(d+1)+i, Gd : K → R are Lipschitz continuous. This and (3.16) show that there exists L ∈ (0, ∞) such that for all t ∈ [0, δ] we have that

1 2 1 2 kG(Θt ) − G(Θt )k ≤ LkΘt − Θt k. (3.17)

15 1 2 In the following let M : [0, ∞) → [0, ∞) satisfy for all t ∈ [0, ∞) that Mt = sups∈(0,t]kΘs − Θsk.  1 2 Observe that the fact that inf t ∈ [0, ∞):Θt 6= Θt = 0 proves for all t ∈ (0, ∞) that Mt > 0. Moreover, note that (3.17) ensures for all t ∈ (0, δ) that

Z t Z t Z t 1 2 1 2 1 2 kΘ − Θ k = G(Θ ) du − G(Θ ) du ≤ kG(Θ ) − G(Θ )k du t t u u u u 0 0 0 (3.18) Z t 1 2 ≤ L kΘu − Θuk du ≤ LtMt. 0 Combining this with the fact that M is non-decreasing shows for all t ∈ (0, δ), s ∈ (0, t] that

1 2 kΘs − Θsk ≤ LsMs ≤ LtMt. (3.19)

This demonstrates for all t ∈ (0, min{L−1, δ}) that

0 < Mt ≤ LtMt < Mt, (3.20) which is a contradiction. The proof of Lemma 3.2 is thus complete.

3.3 Existence and uniqueness properties for solutions of GF differential equa- tions d d Theorem 3.3. Assume Setting 2.1 and let θ ∈ R . Then there exists a unique Θ ∈ C([0, ∞), R ) which satisfies for all t ∈ [0, ∞), s ∈ [t, ∞) that

Z t Θt Θs Θt = θ − G(Θu) du and D ⊆ D . (3.21) 0 Proof of Theorem 3.3. Proposition 3.1 establishes the existence and Lemma 3.2 establishes the uniqueness. The proof of Theorem 3.3 is thus complete.

4 Semialgebraic sets and functions

In this section we establish in Corollary 4.10 in Subsection 4.3 below that under the assump- d tion that both the target function f :[a, b] → R and the unnormalized density function p:[a, b]d → [0, ∞) are piecewise polynomial in the sense of Definition 4.9 in Subsection 4.3 d we have that the risk function L: R → R is a semialgebraic function in the sense of Defini- tion 4.3 in Subsection 4.1. In Definition 4.9 we specify precisely what we mean by a piecewise polynomial function, in Definition 4.2 in Subsection 4.1 we recall the notion of a semialgebraic set, and in Definition 4.3 we recall the notion of a semialgebraic function. In the scientific literature Definitions 4.2 and 4.3 can in a slightly different presentational form, e.g., be found in Bierstone & Milman [7, Definitions 1.1 and 1.2] and Attouch et al. [4, Definition 2.1]. d Note that the risk function L: R → R is given through a parametric in the sense d R θ 2 that for all θ ∈ R we have that L(θ) = [a,b]d (f(y)−N (y)) p(y) λ(dy). In general, parametric integrals of semialgebraic functions are no longer semialgebraic functions and the characteri- zation of functions that can occur as such integrals is quite involved (cf. Kaiser [27]). This is the reason why we introduce in Definition 4.6 in Subsection 4.2 below a suitable subclass of the class of semialgebraic functions which is rich enough to contain the realization functions of ANNs with ReLU activation (cf. (4.28) in Subsection 4.2 below) and which can be shown to be closed under integration (cf. Proposition 4.8 in Subsection 4.2 below for the precise statement).

16 4.1 Semialgebraic sets and functions n 2 Definition 4.1 (Set of polynomials). Let n ∈ N0. Then we denote by Pn ⊆ C(R , R) the set n of all polynomials from R to R. n Definition 4.2 (Semialgebraic sets). Let n ∈ N and let A ⊆ R be a set. Then we say that A is a semialgebraic set if and only if there exist k ∈ N,(Pi,j,`)(i,j,`)∈{1,2,...,k}2×{0,1} ⊆ Pn such that

k k [ \ n A = {x ∈ R : Pi,j,0(x) = 0 < Pi,j,1(x)} (4.1) i=1 j=1

(cf. Definition 4.1).

n m Definition 4.3 (Semialgebraic functions). Let m, n ∈ N and let f : R → R be a function. n Then we say that f is a semialgebraic function if and only if it holds that {(x, f(x)): x ∈ R } ⊆ m+n R is a semialgebraic set (cf. Definition 4.2). n Lemma 4.4. Let n ∈ N and let f, g : R → R be semialgebraic functions (cf. Definition 4.3). Then

n (i) it holds that R 3 x 7→ f(x) + g(x) ∈ R is semialgebraic and n (ii) it holds that R 3 x 7→ f(x)g(x) ∈ R is semialgebraic. Proof of Lemma 4.4. Observe that, e.g., Coste [15, Corollary 2.9] (see, e.g., also Bierstone & Milman [7, Section 1]) establishes items (i) and (ii). The proof of Lemma 4.4 is thus complete.

4.2 On the semialgebraic property of certain parametric integrals

Definition 4.5 (Set of rational functions). Let n ∈ N. Then we denote by ℛn the set given by

( " ( P (x) #) n n Q(x) : Q(x) 6= 0 ℛn = R: R → R: ∃ P,Q ∈ Pn : ∀ x ∈ R : R(x) = (4.2) 0 : Q(x) = 0

(cf. Definition 4.1).

Definition 4.6. Let m ∈ N, n ∈ N0. Then we denote by Am,n the R-vector space given by

n m n h Am,n = span f : R × R → R: ∃ r ∈ N,A1,A2,...,Ar ∈ {{0}, [0, ∞), (0, ∞)}, m n R ∈ ℛm,Q ∈ Pn,P = (Pi,j)(i,j)∈{1,2,...,r}×{0,1,...,n} ⊆ Pm : ∀ θ ∈ R , x = (x1, . . . , xn) ∈ R : io Qr 1 Pn  f(θ, x) = R(θ)Q(x) i=1 Ai Pi,0(θ) + j=1 Pi,j(θ)xj (4.3)

(cf. Definitions 4.1 and 4.5).

Lemma 4.7. Let m ∈ N, f ∈ Am,0 (cf. Definition 4.6). Then f is semialgebraic (cf. Defini- tion 4.3).

Proof of Lemma 4.7. Throughout this proof let r ∈ N, A1,A2,...,Ar ∈ {{0}, [0, ∞), (0, ∞)}, m m R ∈ ℛm, P = (Pi)i∈{1,2,...,r} ⊆ Pm, and let g : R → R satisfy for all θ ∈ R that Qr 1 g(θ) = R(θ) i=1 Ai (Pi(θ)) (4.4)

2 0 0 0 Note that R = {0}, C(R , R) = C({0}, R), and #(C(R , R)) = #(C({0}, R)) = ∞. In particular, this shows n n for all n ∈ N0 that dim(R ) = n and #(C(R , R)) = ∞.

17 (cf. Definitions 4.1 and 4.5). Since sums of semialgebraic functions are again semialgebraic (cf. Lemma 4.4), it suffices to show that g is semialgebraic. Furthermore, note that for all y ∈ R it holds that 1(0,∞)(y) = 1 − 1[0,∞)(−y) and 1{0}(y) = 1[0,∞)(y)1[0,∞)(−y). Hence, by linearity we may assume for all i ∈ {1, 2, . . . , r} that Ai = [0, ∞). Next let Q1,Q2 ∈ Pm satisfy for all m x ∈ R that ( Q1(x) : Q2(x) 6= 0 R(x) = Q2(x) (4.5) 0 : Q2(x) = 0. m Observe that the graph of R 3 θ 7→ R(θ) ∈ R is given by

m {(θ, y) ∈ R × R: Q2(θ) = 0, y = 0} m ∪ {(θ, y) ∈ R × R: Q2(θ) 6= 0,Q2(θ)y − Q1(θ) = 0}. (4.6)

Since both of these sets are described by polynomial equations and inequalities, it follows that m R 3 θ 7→ R(θ) ∈ R is semialgebraic. In addition, note that for all i ∈ {1, 2, . . . , r} the graph m of R 3 θ 7→ 1[0,∞)(Pi(θ)) ∈ R is given by

m m {(θ, y) ∈ R × R: Pi(θ) < 0, y = 0} ∪ {(θ, y) ∈ R × R: Pi(θ) ≥ 0, y = 1}. (4.7)

m This demonstrates for all i ∈ {1, 2, . . . , r} that R 3 θ 7→ 1[0,∞)(Pi(θ)) ∈ R is semialgebraic. Combining this and (4.4) with Lemma 4.4 demonstrates that g is semialgebraic. The proof of Lemma 4.7 is thus complete.

Proposition 4.8. Let m, n ∈ N, a ∈ R, b ∈ (a, ∞), f ∈ Am,n (cf. Definition 4.6). Then

" Z b # m n−1 R × R 3 (θ, x1, . . . , xn−1) 7→ f(θ, x1, . . . , xn) dxn ∈ R ∈ Am,n−1. (4.8) a

Proof of Proposition 4.8. By linearity of the integral it suffices to consider a function f of the form r Y 1 Pn  f(θ, x) = R(θ)Q(x) Ai Pi,0(θ) + j=1 Pi,j(θ)xj (4.9) i=1 where r ∈ N,(Pi,j)(i,j)∈{1,2,...,r}×{0,1,...,n} ⊆ Pm, A1,A2,...,Ar ∈ {{0}, (0, ∞), [0, ∞)}, Q ∈ Pn, and R ∈ ℛm (cf. Definitions 4.1 and 4.5). Moreover, observe that for all y ∈ R it holds that 1(0,∞)(y) = 1 − 1[0,∞)(−y) and 1{0}(y) = 1[0,∞)(y)1[0,∞)(−y). Hence, by linearity we may assume that Ai = [0, ∞) for all i ∈ {1, 2, . . . , r}. Furthermore, by linearity we may assume that Q is of the form Qn i` Q(x1, . . . , xn) = `=1(x`) (4.10) with i1, i2, . . . , in ∈ N0. In the following let s: R → R satisfy for all x ∈ R that s(x) = 1 1 m θ θ (0,∞)(x) − (0,∞)(−x), for every θ ∈ R , k ∈ {−1, 0, 1} let Sk ⊆ {1, 2, . . . , r} satisfy Sk = {i ∈ m n {1, 2, . . . , r}: s(Pi,n(θ)) = k}, and for every i ∈ {1, 2, . . . , r} let Zi : R × R → R satisfy for all m n (θ, x) ∈ R × R that Pn−1 Zi(θ, x) = −Pi,0(θ) − j=1 Pi,j(θ)xj. (4.11) m n Note that for all θ ∈ R , x = (x1, . . . , xn) ∈ R with xn ∈ [a, b], f(θ, x) can only be nonzero if

θ Zi(θ, x) ∀ i ∈ S1 : xn ≥ , Pi,n(θ)

θ Zi(θ, x) (4.12) ∀ i ∈ S−1 : xn ≤ , Pi,n(θ) θ ∀ i ∈ S0 : − Zi(θ, x) ≥ 0.

18 m n−1 Hence, if for given θ ∈ R ,(x1, . . . , xn−1) ∈ R there exists xn ∈ [a, b] which satisfies these conditions, we have

Z b f(θ, x1, . . . , xn) dxn a  ( )!in+1 ( )!in+1 R(θ)   Z (θ, x) Z (θ, x) = Qn−1 xi` min b, min j − max a, max j . `=1 `  θ θ  in + 1 j∈S−1 Pj,n(θ) j∈S1 Pj,n(θ) (4.13)

R b Otherwise, we have that a f(θ, x1, . . . , xn) dxn = 0. It remains to write these expressions in the different cases as a sum of functions of the required form by introducing suitable indicator functions. Observe that there are four possible cases where the integral is nonzero:

Zj (θ,x) Zj (θ,x) • It holds that a < max θ < min θ < b. In this case, we have j∈S1 Pj,n(θ) j∈S−1 Pj,n(θ)

Z b f(θ, x1, . . . , xn) dxn a  !in+1 !in+1 (4.14) R(θ)   Z (θ, x) Z (θ, x) = Qn−1 xi` min j − max j . `=1 `  θ θ  in + 1 j∈S−1 Pj,n(θ) j∈S1 Pj,n(θ)

Zj (θ,x) Zj (θ,x) • It holds that a < max θ < b ≤ min θ . In this case, we have j∈S1 Pj,n(θ) j∈S−1 Pj,n(θ)

 !in+1 Z b R(θ)   Z (θ, x) Qn−1 i` in+1 j f(θ, x1, . . . , xn) dxn = x b − max . (4.15) `=1 `  θ  a in + 1 j∈S1 Pj,n(θ)

Zj (θ,x) Zj (θ,x) • It holds that max θ ≤ a < min θ < b. In this case, we have j∈S1 Pj,n(θ) j∈S−1 Pj,n(θ)

 !in+1  Z b R(θ)   Z (θ, x) Qn−1 i` j in+1 f(θ, x1, . . . , xn) dxn = x min − a . (4.16) `=1 `  θ  a in + 1 j∈S−1 Pj,n(θ)

Zj (θ,x) Zj (θ,x) • It holds that max θ ≤ a < b ≤ min θ . In this case, we have j∈S1 Pj,n(θ) j∈S−1 Pj,n(θ)

Z b R(θ)   Qn−1 i`  in+1 in+1 f(θ, x1, . . . , xn) dxn = `=1 x` b − a . (4.17) a in + 1

θ Since these four cases are disjoint, by summing over all possible choices of the sets Sk , k ∈ θ θ {−1, 0, 1}, and all choices of subsets of S1 , S−1 where the maximal/minimal values are achieved, we can write

Z b R(θ)   Qn−1 i` f(θ, x1, . . . , xn) dxn = `=1 x` [(I) + (II) + (III) + (IV )], (4.18) a in + 1

19 where " # X Y Y Y (I) = 1(0,∞)(Pj,n(θ)) 1(0,∞)(−Pj,n(θ)) (1{0}(Pj,n(θ))1[0,∞)(−Zj(θ, x)) A∪˙ B∪˙ C={1,...,r} j∈A j∈B j∈C ""      X X Y Zi(θ, x) Zi(θ, x) Zmin I (θ, x) 1(a,b) 1{0} − Pi,n(θ) Pi,n(θ) Pmin I,n(θ) ∅6=I⊆A ∅6=J ⊆B i∈I        Y Zmin I (θ, x) Zj(θ, x) Y Zi(θ, x) Zi(θ, x) Zmin J (θ, x) × 1(0,∞) − 1(a,b) 1{0} − Pmin I,n(θ) Pj,n(θ) Pi,n(θ) Pi,n(θ) Pmin J ,n(θ) j∈A\I i∈J    # Y Zj(θ, x) Zmin J (θ, x) Zmin J (θ, x) Zmin I (θ, x) × 1(0,∞) − 1(0,∞) − Pj,n(θ) Pmin J ,n(θ) Pmin J ,n(θ) Pmin I,n(θ) j∈B\J " ## Z (θ, x)in+1 Z (θ, x)in+1 × min J − min I , Pmin J ,n(θ) Pmin I,n(θ) (4.19) " # X Y Y Y (II) = 1(0,∞)(Pj,n(θ)) 1(0,∞)(−Pj,n(θ)) (1{0}(Pj,n(θ))1[0,∞)(−Zj(θ, x)) A∪˙ B∪˙ C={1,...,r} j∈A j∈B j∈C ""      X Y Zi(θ, x) Zi(θ, x) Zmin I (θ, x) 1(a,b) 1{0} − Pi,n(θ) Pi,n(θ) Pmin I,n(θ) ∅6=I⊆A i∈I      Y Zmin I (θ, x) Zj(θ, x) Y Zi(θ, x) × 1(0,∞) − 1[b,∞) Pmin I,n(θ) Pj,n(θ) Pi,n(θ) j∈A\I i∈B " ## Z (θ, x)in+1 × bin+1 − min I , Pmin I,n(θ) (4.20) " # X Y Y Y (III) = 1(0,∞)(Pj,n(θ)) 1(0,∞)(−Pj,n(θ)) (1{0}(Pj,n(θ))1[0,∞)(−Zj(θ, x)) A∪˙ B∪˙ C={1,...,r} j∈A j∈B j∈C ""         X Y Zi(θ, x) Y Zi(θ, x) Zi(θ, x) Zmin J (θ, x) 1(−∞,a] 1(a,b) 1{0} − Pi,n(θ) Pi,n(θ) Pi,n(θ) Pmin J ,n(θ) ∅6=J ⊆B i∈A i∈J  # " in+1 ## Y Zj(θ, x) Zmin J (θ, x) Zmin J (θ, x) in+1 × 1(0,∞) − × − a , Pj,n(θ) Pmin J ,n(θ) Pmin J ,n(θ) j∈B\J (4.21) and " # X Y Y Y (IV ) = 1(0,∞)(Pj,n(θ)) 1(0,∞)(−Pj,n(θ)) (1{0}(Pj,n(θ))1[0,∞)(−Zj(θ, x)) A∪˙ B∪˙ C={1,...,r} j∈A j∈B j∈C    ! Y Zi(θ, x) Y Zi(θ, x)  in+1 in+1 × 1(−∞,a] 1[b,∞) b − a . Pi,n(θ) Pi,n(θ) i∈A i∈B (4.22) Furthermore, note that, e.g., in (I) we have for all i ∈ I ⊆ A that       Zi(θ, x) Zi(θ, x) Zi(θ, x) 1(a,b) = 1(a,∞) 1(−∞,b) Pi,n(θ) Pi,n(θ) Pi,n(θ) (4.23)

= 1(0,∞)(Zi(θ, x) − aPi,n(θ))1(0,∞)(bPi,n(θ) − Zi(θ, x)).

20 Similarly, the other indicator functions can be brought into the correct form, taking into account the different signs of Pj,n(θ) for j ∈ A and j ∈ B. Moreover, observe that the remaining terms can be written as linear combinations of rational functions in θ and polynomials in x. Hence, we obtain that the expressions (I), (II), (III), (IV ) are elements of Am,n−1. The proof of Proposition 4.8 is thus complete.

4.3 On the semialgebraic property of the risk function d Definition 4.9. Let d ∈ N, let A ⊆ R be a set, and let f : A → R be a function. Then n×d we say that f is piecewise polynomial if and only if there exist n ∈ N, α1, α2, . . . , αn ∈ R , n β1, β2, . . . , βn ∈ R , P1,P2,...,Pn ∈ Pd such that for all x ∈ A it holds that

n P 1 n f(x) = i=1[Pi(x) [0,∞) (αix + βi)] (4.24)

(cf. Definition 4.1).

Corollary 4.10. Assume Setting 2.1 and assume that f and p are piecewise polynomial (cf. Definition 4.9). Then L is semialgebraic (cf. Definition 4.3).

d d Proof of Corollary 4.10. Throughout this proof let F : R → R and P: R → R satisfy for all d x ∈ R that ( ( f(x): x ∈ [a, b]d p(x): x ∈ [a, b]d F (x) = and P(x) = (4.25) 0 : x∈ / [a, b]d 0 : x∈ / [a, b]d.

Note that (4.25) and the assumption that f and p are piecewise polynomial assure that

h d d i h d d i R × R 3 (θ, x) 7→ F (x) ∈ R ∈ Ad,d and R × R 3 (θ, x) 7→ P(x) ∈ R ∈ Ad,d (4.26)

d d (cf. Definition 4.6). In addition, observe that the fact that for all θ ∈ R , x ∈ R we have that

H θ θ X θ nPd θ θ o N (x) = c + vi max `=1 wi,`x` + bi , 0 i=1 (4.27) H     θ X θ Pd θ θ 1 Pd θ θ = c + vi `=1 wi,`x` + bi [0,∞) `=1 wi,`x` + bi i=1 demonstrates that h d d θ i R × R 3 (θ, x) 7→ N (x) ∈ R ∈ Ad,d. (4.28)

Combining this with (4.26) and the fact that Ad,d is an algebra proves that

h d d θ 2 i R × R 3 (θ, x) 7→ (N (x) − F (x)) P(x) ∈ R ∈ Ad,d. (4.29)

This, Proposition 4.8, and induction demonstrate that

" Z b Z b Z b # d θ 2 R 3 θ 7→ ··· (N (x) − F (x)) P(x) dxd ··· dx2 dx1 ∈ R ∈ Ad,0. (4.30) a a a

Fubini’s theorem hence implies that L ∈ Ad,0. Combining this and Lemma 4.7 shows that L is semialgebraic. The proof of Corollary 4.10 is thus complete.

21 5 Convergence rates for solutions of GF differential equations

In this section we employ the findings from Sections2 and4 to establish in Proposition 5.2 in Subsection 5.2 below, in Proposition 5.3 in Subsection 5.2, and in Theorem 5.4 in Subsection 5.3 below several convergence rate results for solutions of GF differential equations. Theorem 1.2 in the introduction is a direct consequence of Theorem 5.4. Our proof of Theorem 5.4 is based on an application of Proposition 5.3 and our proof of Proposition 5.3 uses Proposition 5.2. Our proof of Proposition 5.2, in turn, employs Proposition 5.1 in Subsection 5.1 below. In Propo- d sition 5.1 we establish that under the assumption that the target function f :[a, b] → R and the unnormalized density function p:[a, b]d → [0, ∞) are piecewise polynomial (see Defini- d tion 4.9 in Subsection 4.3) we have that the risk function L: R → R satisfies an appropriately generalizedLojasiewicz inequality. In the proof of Proposition 5.1 the classicalLojasiewicz inequality for semialgebraic or subanalytic functions (cf., e.g., Bierstone & Milman [7]) is not directly applicable since the d d generalized gradient function G : R → R is not continuous. We will employ the more general results from Bolte et al. [8] which also apply to not necessarily continuously differentiable functions. The arguments used in the proof of Proposition 5.2 are slight adaptions of well-known arguments in the literature; see, e.g., Kurdyka et al. [29, Section 1], Bolte et al. [8, Theorem 4.5], or Absil et al. [1, Theorem 2.2]. On the one hand, in Kurdyka et al. [29, Section 1] and Absil et al. [1, Theorem 2.2] it is assumed that the object function of the considered optimization problem is analytic and in Bolte et al. [8, Theorem 4.5] it is assumed that the objective function of the considered optimization problem is convex or lower C2 and Proposition 5.2 does not require these assumptions. On the other hand, Bolte et al. [8, Theorem 4.5] consider more general differential dynamics and the considered gradients are allowed to be more general than d d the specific generalized gradient function G : R → R which is considered in Proposition 5.2.

5.1 GeneralizedLojasiewicz inequality for the risk function Proposition 5.1 (GeneralizedLojasiewicz inequality). Assume Setting 2.1, assume that p and d f are piecewise polynomial, and let ϑ ∈ R (cf. Definition 4.9). Then there exist ε, D ∈ (0, ∞), α ∈ (0, 1) such that for all θ ∈ Bε(ϑ) it holds that

|L(θ) − L(ϑ)|α ≤ DkG(θ)k. (5.1)

d d Proof of Proposition 5.1. Throughout this proof let M: R → [0, ∞] satisfy for all θ ∈ R that

M(θ) = inf({khk: h ∈ ∂L(θ)} ∪ {∞}). (5.2)

d Note that Proposition 2.12 implies for all θ ∈ R that M(θ) ≤ kG(θ)k. Furthermore, observe that Corollary 4.10, the fact that semialgebraic functions are subanalytic, and Bolte et al. [8, Theorem 3.1 and Remark 3.2] ensure that there exist ε, D ∈ (0, ∞), a ∈ [0, 1) which satisfy for all θ ∈ Bε(ϑ) that |L(θ) − L(ϑ)|a ≤ DM(θ). (5.3) d Combining this with the fact that for all θ ∈ R it holds that M(θ) ≤ kG(θ)k and the fact that supθ∈Bε(ϑ)|L(θ) − L(ϑ)| < ∞ demonstrates that for all θ ∈ Bε(ϑ), α ∈ (a, 1) we have that

α a α−a |L(θ) − L(ϑ)| ≤ |L(θ) − L(ϑ)| supψ∈B (ϑ)|L(ψ) − L(ϑ)| ε (5.4) α−a ≤ D supψ∈Bε(ϑ)|L(ψ) − L(ϑ)| kG(θ)k. This completes the proof of Proposition 5.1.

22 5.2 Local convergence for solutions of GF differential equations d Proposition 5.2. Assume Setting 2.1 and let ϑ ∈ R , ε, D ∈ (0, ∞), α ∈ (0, 1) satisfy for all θ ∈ Bε(ϑ) that |L(θ) − L(ϑ)|α ≤ DkG(θ)k. (5.5) d Then there exists δ ∈ (0, ε) such that for all Θ ∈ C([0, ∞), R ) with Θ0 ∈ Bδ(ϑ), ∀ t ∈ R t [0, ∞):Θt = Θ0 − 0 G(Θs) ds, and inft∈{s∈[0,∞):Θs∈Bε(ϑ)} L(Θt) ≥ L(ϑ) there exists ψ ∈ −1 R ∞ L ({L(ϑ)}) such that for all t ∈ [0, ∞) it holds that Θt ∈ Bε(ϑ), 0 kG(Θs)k ds ≤ ε, −2 −1 |L(Θt) − L(ψ)| ≤ (1 + D t) , and

1−α α h i− min{1, α } −1/α  1−α kΘt − ψk ≤ 1 + D (1 − α) t . (5.6)

Proof of Proposition 5.2. Note that the fact that L is continuous implies that there exists δ ∈ (0, ε/3) which satisfies for all θ ∈ Bδ(ϑ) that ε(1 − α) 1 − α  |L(θ) − L(ϑ)|1−α ≤ min , , 1 . (5.7) 3D D

d R t In the following let Θ ∈ C([0, ∞), R ) satisfy ∀ t ∈ [0, ∞):Θt = Θ0 − 0 G(Θs) ds,Θ0 ∈ Bδ(ϑ), and

inft∈{s∈[0,∞):Θs∈Bε(ϑ)} L(Θt) ≥ L(ϑ). (5.8) In the first step we show that for all t ∈ [0, ∞) it holds that

Θt ∈ Bε(ϑ). (5.9)

Observe that, e.g., [24, Lemma 3.1] ensures for all t ∈ [0, ∞) that

Z t 2 L(Θt) = L(Θ0) − kG(Θs)k ds. (5.10) 0

This implies that [0, ∞) 3 t 7→ L(Θt) ∈ [0, ∞) is non-increasing. Next let L: [0, ∞) → R satisfy for all t ∈ [0, ∞) that L(t) = L(Θt) − L(ϑ) (5.11) and let T ∈ [0, ∞] satisfy

T = inf({t ∈ [0, ∞): kΘt − ϑk ≥ ε} ∪ {∞}). (5.12)

We intend to show that T = ∞. Note that (5.8) assures for all t ∈ [0,T ) that L(t) ≥ 0. Moreover, observe that (5.10) and (5.11) ensure that for almost all t ∈ [0,T ) it holds that L is 0 d 2 differentiable at t and satisfies L (t) = dt (L(Θt)) = −kG(Θt)k . In the following let τ ∈ [0,T ] satisfy τ = inf({t ∈ [0,T ): L(t) = 0} ∪ {T }). (5.13) Note that the fact that L is non-increasing implies that for all s ∈ [τ, T ) it holds that L(s) = 0. Combining this with (5.10) demonstrates for almost all s ∈ (τ, T ) that G(Θs) = 0. This proves for all s ∈ [τ, T ) that Θs = Θτ . Next observe that (5.5) ensures that for all t ∈ [0, τ) it holds that α α 0 < [L(t)] = |L(Θt) − L(ϑ)| ≤ DkG(Θt)k. (5.14) Combining this with the chain rule proves for almost all t ∈ [0, τ) that d ([L(t)]1−α) = (1 − α)[L(t)]−α−kG(Θ )k2 dt t (5.15) −1 −1 2 −1 ≤ −(1 − α)D kG(Θt)k kG(Θt)k = −D (1 − α)kG(Θt)k.

23 In addition, note that the fact that [0, ∞) 3 t 7→ L(t) ∈ R is absolutely continuous and the fact 1−α that for all r ∈ (0, ∞) it holds that r, ∞) 3 y 7→ y ∈ R is Lipschitz continuous demonstrate 1−α for all t ∈ [0, τ) that [0, t] 3 s 7→ [L(s)] ∈ R is absolutely continuous. Integrating (5.15) hence shows for all s, t ∈ [0, τ) with t ≤ s that Z s −1 1−α 1−α −1 1−α kG(Θu)k du ≤ −D(1 − α) ([L(s)] − [L(t)] ) ≤ D(1 − α) [L(t)] . (5.16) t

This and the fact that for almost all s ∈ (τ, T ) it holds that G(Θs) = 0 ensure that for all s, t ∈ [0,T ) with t ≤ s we have that Z s −1 1−α kG(Θu)k du ≤ D(1 − α) [L(t)] . (5.17) t Combining this with (5.7) demonstrates for all t ∈ [0,T ) that

Z t Z t 1−α D|L(Θ0) − L(ϑ)| nε o kΘt − Θ0k = G(Θs) ds ≤ kG(Θs)k ds ≤ ≤ min , 1 . (5.18) 0 0 1 − α 3

This, the fact that δ < ε/3, and the triangle inequality assure for all t ∈ [0,T ) that ε ε ε 2ε kΘ − ϑk ≤ kΘ − Θ k + kΘ − ϑk ≤ + δ ≤ + = . (5.19) t t 0 0 3 3 3 3 Combining this with (5.12) proves that T = ∞. This establishes (5.9). Next observe that the fact that T = ∞ and (5.18) prove that Z ∞ nε o kG(Θs)k ds ≤ min , 1 ≤ ε < ∞. (5.20) 0 3 In the following let σ : [0, ∞) → [0, ∞) satisfy for all t ∈ [0, ∞) that Z ∞ σ(t) = kG(Θs)k ds. (5.21) t

Note that (5.20) proves that lim supt→∞ σ(t) = 0. In addition, observe that (5.20) assures that d there exists ψ ∈ R such that lim supt→∞kΘt − ψk = 0. (5.22) In the next step we combine the weak chain rule for the risk function in (5.10) with (5.9) and (5.5) to obtain that for almost all t ∈ [0, ∞) we have that

0 2 −2 2α L (t) = −kG(Θt)k ≤ −D [L(t)] . (5.23)

In addition, note that the fact that L is non-increasing and (5.7) ensure that for all t ∈ [0, ∞) it holds that L(t) ≤ L(0) ≤ 1. Therefore, we get for almost all t ∈ [0, ∞) that

L0(t) ≤ −D−2[L(t)]2. (5.24)

Combining this with the fact that for all t ∈ [0, τ) it holds that L(t) > 0 establishes for almost all t ∈ [0, τ) that d  D2  D2L0(t) = − ≥ 1. (5.25) dt L(t) [L(t)]2 The fact that for all t ∈ [0, τ) it holds that [0, t] 3 s 7→ L(s) ∈ (0, ∞) is absolutely continuous hence demonstrates for all t ∈ [0, τ) that

D2 D2 ≥ + t ≥ D2 + t. (5.26) L(t) L(0)

24 Therefore, we infer for all t ∈ [0, τ) that

−1 −1 L(t) ≤ D2D2 + t = 1 + D−2t . (5.27)

This and the fact that for all t ∈ [τ, ∞) it holds that L(t) = 0 prove that for all t ∈ [0, ∞) we have that −2 −1 |L(Θt) − L(ϑ)| = L(t) ≤ 1 + D t . (5.28)

Furthermore, observe that (5.22) and the fact that L is continuous imply that lim supt→∞|L(Θt)− L(ψ)| = 0. Hence, we obtain that L(ψ) = L(ϑ). This shows for all t ∈ [0, ∞) that

−2 −1 |L(Θt) − L(ψ)| ≤ 1 + D t . (5.29)

In the next step we establish a convergence rate for the quantity kΘt − ψk, t ∈ [0, ∞). We d accomplish this by employing an upper bound for the tail length of the curve Θt ∈ R , t ∈ [0, ∞). More formally, note that (5.17), (5.9), and (5.5) demonstrate for all t ∈ [0, ∞) that

Z ∞ Z s  σ(t) = kG(Θu)k du = lim kG(Θu)k du t s→∞ t (5.30) −1 1−α −1 1−α ≤ D(1 − α) [L(t)] ≤ D(1 − α) (DkG(Θt)k) α . R ∞ Next observe that the fact that for all t ∈ [0, ∞) it holds that σ(t) = 0 kG(Θs)k ds − R t 0 0 kG(Θs)k ds shows that for almost all t ∈ [0, ∞) we have that σ (t) = −kG(Θt)k. This and 1−α 1 (5.30) yield for almost all t ∈ [0, ∞) that σ(t) ≤ D /α(1−α)−1[−σ0(t)] α . Therefore, we obtain for almost all t ∈ [0, ∞) that

α σ0(t) ≤ −(1 − α)D−1/ασ(t) 1−α . (5.31)

Combining this with the fact that σ is absolutely continuous implies for all t ∈ [0, ∞) that

α Z t α −1/α σ(t) − σ(0) ≤ −(1 − α)D  1−α [σ(s)] 1−α ds. (5.32) 0

α α −1/α 1−α In the following let β, C ∈ (0, ∞) satisfy β = max{1, 1−α } and C = (1 − α)D . Note that (5.32) and the fact that for all t ∈ [0, ∞) it holds that σ(t) ≤ σ(0) ≤ 1 ensure that for all t ∈ [0, ∞) it holds that Z t σ(t) ≤ σ(0) − C [σ(s)]β ds. (5.33) 0 This, the fact that σ is non-increasing, and the fact that for all t ∈ [0, ∞) it holds that 0 ≤ σ(t) ≤ 1 prove that for all t ∈ [0, ∞) we have that

[σ(t)]β ≤ σ(t) ≤ σ(0) − C[σ(t)]βt ≤ 1 − Ct[σ(t)]β. (5.34)

− 1 Hence, we obtain for all t ∈ [0, ∞) that σ(t) ≤ (1 + Ct) β . Combining this with the fact that for all t ∈ [0, ∞) it holds that

Z s Z s 

kΘt − ψk ≤ lim supkΘt − Θsk = lim sup G(Θu) du ≤ lim sup kG(Θu)k du s→∞ s→∞ t s→∞ t Z ∞ (5.35) = kG(Θu)k du = σ(t) t

−1/β shows that for all t ∈ [0, ∞) we have that kΘt − ψk ≤ (1 + Ct) . This, (5.9), (5.20), and (5.29) establish (5.6). The proof of Proposition 5.2 is thus complete.

25 5.3 Global convergence for solutions of GF differential equations Proposition 5.3. Assume Setting 2.1, assume that p and f are piecewise polynomial, and let d R t Θ ∈ C([0, ∞), R ) satisfy lim inft→∞kΘtk < ∞ and ∀ t ∈ [0, ∞):Θt = Θ0 − 0 G(Θs) ds (cf. Definition 4.9). Then there exist ϑ ∈ G−1({0}), C, τ, β ∈ (0, ∞) which satisfy for all t ∈ [τ, ∞) that −β −1 kΘt − ϑk ≤ 1 + C(t − τ) and |L(Θt) − L(ϑ)| ≤ 1 + C(t − τ) . (5.36) Proof of Proposition 5.3. First observe that [24, Lemma 3.1] ensures that for all t ∈ [0, ∞) it holds that Z t 2 L(Θt) = L(Θ0) − kG(Θs)k ds. (5.37) 0 This implies that [0, ∞) 3 t 7→ L(Θt) ∈ [0, ∞) is non-increasing. Hence, we obtain that there exists m ∈ [0, ∞) which satisfies that

m = lim supt→∞ L(Θt) = lim inft→∞ L(Θt) = inft∈[0,∞) L(Θt). (5.38) d Moreover, note that the assumption that lim inft→∞kΘtk < ∞ ensures that there exist ϑ ∈ R and τ = (τn)n∈N : N → [0, ∞) which satisfy lim infn→∞ τn = ∞ and

lim supn→∞kΘτn − ϑk = 0. (5.39) Combining this with (5.38) and the fact that L is continuous shows that

L(ϑ) = m and ∀ t ∈ [0, ∞): L(Θt) ≥ L(ϑ). (5.40) Next observe that Proposition 5.1 demonstrates that there exist ε, D ∈ (0, ∞), α ∈ (0, 1) such that for all θ ∈ Bε(ϑ) we have that |L(θ) − L(ϑ)|α ≤ DkG(θ)k. (5.41) Combining this and (5.39) with Proposition 5.2 proves that there exists δ ∈ (0, ε) which d R t satisfies for all Φ ∈ C([0, ∞), R ) with Φ0 ∈ Bδ(ϑ), ∀ t ∈ [0, ∞):Φt = Φ0 − 0 G(Φs) ds, and inft∈{s∈[0,∞):Φs∈Bε(ϑ)} L(Φt) ≥ L(ϑ) that it holds for all t ∈ [0, ∞) that Φt ∈ Bε(ϑ), −2 −1 |L(Φt) − L(ϑ)| ≤ (1 + D t) , and

1−α α h i− min{1, α } −1/α  1−α kΦt − ϑk ≤ 1 + D (1 − α) t . (5.42)

Moreover, note that (5.39) ensures that there exists n ∈ N which satisfies Θτn ∈ Bδ(ϑ). Next d let Φ ∈ C([0, ∞), R ) satisfy for all t ∈ [0, ∞) that

Φt = Θt+τn . (5.43) Observe that (5.40) and (5.43) assure that Z t Φ0 ∈ Bδ(ϑ), inft∈[0,∞) L(Φt) ≥ L(ϑ), and ∀ t ∈ [0, ∞):Φt = Φ0 − G(Φs) ds. (5.44) 0

Combining this with (5.42) proves for all t ∈ [τn, ∞) that −2 −1 |L(Θt) − L(ϑ)| ≤ 1 + D (t − τn) (5.45) and 1−α α h i− min{1, α } −1/α  1−α kΘt − ϑk ≤ 1 + D (1 − α) (t − τn) . (5.46) d Next note that [24, Corollary 2.16] shows that R 3 θ 7→ kG(θ)k ∈ [0, ∞) is lower semicontin- uous. The fact that lim infs→∞kG(Θs)k = 0 and the fact that lim supt→∞kΘt − ϑk = 0 hence imply that G(ϑ) = 0. Combining this with (5.45) and (5.46) establishes (5.36). The proof of Proposition 5.3 is thus complete.

26 Theorem 5.4. Assume Setting 2.1, assume that p and f are piecewise polynomial, and let d R t Θ ∈ C([0, ∞), R ) satisfy lim inft→∞kΘtk < ∞ and ∀ t ∈ [0, ∞):Θt = Θ0 − 0 G(Θs) ds (cf. Definition 4.9). Then there exist ϑ ∈ G−1({0}), C , β ∈ (0, ∞) which satisfy for all t ∈ [0, ∞) that −β −1 kΘt − ϑk ≤ C (1 + t) and |L(Θt) − L(ϑ)| ≤ C (1 + t) . (5.47) Proof of Theorem 5.4. Observe that Proposition 5.3 assures that there exist ϑ ∈ G−1({0}), C, τ, β ∈ (0, ∞) which satisfy for all t ∈ [τ, ∞) that

−β kΘt − ϑk ≤ 1 + C(t − τ) (5.48) and −1 |L(Θt) − L(ϑ)| ≤ 1 + C(t − τ) . (5.49) In the following let C ∈ (0, ∞) satisfy

n −1 −β β β  o C = max C , 1 + τ, C , (1 + τ) , (1 + τ) sups∈[0,τ]kΘs − ϑk , (1 + τ)L(Θ0) . (5.50)

Note that (5.49), (5.50), and the fact that [0, ∞) 3 t 7→ L(Θt) ∈ [0, ∞) is non-increasing show for all t ∈ [0, τ] that

−β −β kΘt − ϑk ≤ sups∈[0,τ]kΘs − ϑk ≤ C (1 + τ) ≤ C (1 + t) (5.51) and

−1 −1 |L(Θt) − L(ϑ)| = L(Θt) − L(ϑ) ≤ L(Θt) ≤ L(Θ0) ≤ C (1 + τ) ≤ C (1 + t) . (5.52)

Moreover, observe that (5.48) and (5.50) imply for all t ∈ [τ, ∞) that

1/β 1/β −β 1/β −β −β kΘt − ϑk ≤ C C + CC (t − τ) ≤ C C − τ + t ≤ C (1 + t) . (5.53)

In addition, note that (5.49) and (5.50) demonstrate for all t ∈ [τ, ∞) that

−1 −1 −1 |L(Θt) − L(ϑ)| ≤ C C + CC (t − τ) ≤ C C − τ + t ≤ C (1 + t) . (5.54)

This completes the proof of Theorem 5.4.

Acknowledgements This work has been funded by the Deutsche Forschungsgemeinschaft (DFG, German Research Foundation) under Germany’s Excellence Strategy EXC 2044-390685587, Mathematics M¨unster: Dynamics-Geometry-Structure.

References

[1] P.-A. Absil, R. Mahony, and B. Andrews. Convergence of the iterates of descent methods for analytic cost functions. SIAM J. Optim., 16(2):531–547, 2005. doi:10.1137/040605266.

[2] Sanjeev Arora, Simon Du, Wei Hu, Zhiyuan Li, and Ruosong Wang. Fine-grained analysis of optimization and generalization for overparameterized two-layer neural net- works. In Kamalika Chaudhuri and Ruslan Salakhutdinov, editors, Proceedings of the 36th International Conference on Machine Learning, volume 97 of Proceedings of Machine Learning Research, pages 322–332, Long Beach, California, USA, 6 2019. PMLR. URL: http://proceedings.mlr.press/v97/arora19a.html.

27 [3] Hedy Attouch and J´erˆomeBolte. On the convergence of the proximal algorithm for nons- mooth functions involving analytic features. Math. Program., 116(1-2, Ser. B):5–16, 2009. doi:10.1007/s10107-007-0133-5.

[4] Hedy Attouch, J´erˆomeBolte, and Benar Fux Svaiter. Convergence of descent methods for semi-algebraic and tame problems: proximal algorithms, forward-backward splitting, and regularized Gauss-Seidel methods. Math. Program., 137(1-2, Ser. A):91–129, 2013. doi:10.1007/s10107-011-0484-9.

[5] Francis Bach and Eric Moulines. Non-strongly-convex smooth stochastic approximation with convergence rate O(1/n). In C. J. C. Burges, L. Bottou, M. Welling, Z. Ghahra- mani, and K. Q. Weinberger, editors, Advances in Neural Information Processing Systems, volume 26, pages 773–781. Curran Associates, Inc., 2013. URL: http://papers.nips. cc/paper/4900-non-strongly-convex-smooth-stochastic-approximation-with- convergence-rate-o1n.pdf.

[6] Dimitri P. Bertsekas and John N. Tsitsiklis. Gradient convergence in gradient methods with errors. SIAM Journal on Optimization, 10(3):627–642, 2000. doi:10.1137/S10526 23497331063.

[7] Edward Bierstone and Pierre D. Milman. Semianalytic and subanalytic sets. Inst. Hautes Etudes´ Sci. Publ. Math., 67:5–42, 1988. URL: http://www.numdam.org/item?id=PMIHES 1988 67 5 0.

[8] J´erˆomeBolte, Aris Daniilidis, and Adrian Lewis. Thelojasiewicz inequality for nonsmooth subanalytic functions with applications to subgradient dynamical systems. SIAM J. Optim., 17(4):1205–1223, 2006. doi:10.1137/050644641.

[9] Zhengdao Chen, Grant Rotskoff, Joan Bruna, and Eric Vanden-Eijnden. A dynamical central limit theorem for shallow neural networks. In H. Larochelle, M. Ranzato, R. Hadsell, M. F. Balcan, and H. Lin, editors, Advances in Neural Information Processing Systems, volume 33, pages 22217–22230. Curran Associates, Inc., 2020. URL: https://proceedi ngs.neurips.cc/paper/2020/file/fc5b3186f1cf0daece964f78259b7ba0-Paper.pdf.

[10] Patrick Cheridito, Arnulf Jentzen, Adrian Riekert, and Florian Rossmannek. A proof of convergence for gradient descent in the training of artificial neural networks for constant target functions, 2021. arXiv:2102.09924.

[11] Patrick Cheridito, Arnulf Jentzen, and Florian Rossmannek. Landscape analysis for shallow ReLU neural networks: complete classification of critical points for affine target functions, 2021. arXiv:2103.10922.

[12] L´ena¨ıcChizat. Sparse optimization on measures with over-parameterized gradient descent. Mathematical Programming, 2021. doi:10.1007/s10107-021-01636-z.

[13] L´ena¨ıcChizat and Francis Bach. On the global convergence of gradient descent for over- parameterized models using optimal transport. In S. Bengio, H. Wallach, H. Larochelle, K. Grauman, N. Cesa-Bianchi, and R. Garnett, editors, Advances in Neural Information Processing Systems, volume 31, pages 3036–3046. Curran Associates, Inc., 2018. URL: https://proceedings.neurips.cc/paper/2018/file/a1afc58c6ca9540d057299ec301 6d726-Paper.pdf.

[14] L´ena¨ıc Chizat, Edouard Oyallon, and Francis Bach. On lazy training in differentiable programming. In H. Wallach, H. Larochelle, A. Beygelzimer, F. d'Alch´e-Buc,E. Fox, and R. Garnett, editors, Advances in Neural Information Processing Systems, volume 32.

28 Curran Associates, Inc., 2019. URL: https://proceedings.neurips.cc/paper/2019/fi le/ae614c557843b1df326cb29c57225459-Paper.pdf. [15] Michel Coste. An introduction to semialgebraic geometry. Istituti editoriali e poligrafici internazionali, Pisa, 2000.

[16] Steffen Dereich and Sebastian Kassing. Convergence of stochastic gradient descent schemes for Lojasiewicz-landscapes, 2021. arXiv:2102.09385. [17] Simon S. Du, Xiyu Zhai, Barnab´asP´oczos,and Aarti Singh. Gradient descent provably optimizes over-parameterized neural networks. In 7th International Conference on Learning Representations, ICLR 2019, New Orleans, LA, USA, May 6-9, 2019. OpenReview.net, 2019. URL: https://openreview.net/forum?id=S1eK3i09YQ. [18] Weinan E, Chao Ma, Stephan Wojtowytsch, and Lei Wu. Towards a mathematical un- derstanding of neural network-based machine learning: what we know and what we don’t, 2020. arXiv:2009.10713. [19] Weinan E, Chao Ma, and Lei Wu. A comparative analysis of optimization and general- ization properties of two-layer neural network and random feature models under gradient descent dynamics. Sci. China Math., 63(7):1235–1258, 2020. doi:10.1007/s11425-019- 1628-5. [20] Benjamin Fehrman, Benjamin Gess, and Arnulf Jentzen. Convergence rates for the stochas- tic gradient descent method for non-convex objective functions. J. Mach. Learn. Res., 21:Paper No. 136, 48, 2020.

[21] Arthur Jacot, Franck Gabriel, and Clement Hongler. Neural tangent kernel: Convergence and generalization in neural networks. In S. Bengio, H. Wallach, H. Larochelle, K. Grau- man, N. Cesa-Bianchi, and R. Garnett, editors, Advances in Neural Information Processing Systems, volume 31. Curran Associates, Inc., 2018. URL: https://proceedings.neurip s.cc/paper/2018/file/5a4be1fa34e62bb8a6ec6b91d2462f5a-Paper.pdf. [22] Arnulf Jentzen and Timo Kr¨oger. Convergence rates for gradient descent in the training of overparameterized artificial neural networks with biases, 2021. arXiv:2102.11840. [23] Arnulf Jentzen, Benno Kuckuck, Ariel Neufeld, and Philippe von Wurstemberger. Strong error analysis for stochastic gradient descent optimization algorithms. IMA Journal of , 41(1):455–492, 2021. doi:10.1093/imanum/drz055. [24] Arnulf Jentzen and Adrian Riekert. Convergence analysis for gradient flows in the training of artificial neural networks with ReLU activation, 2021. arXiv:2107.04479. [25] Arnulf Jentzen and Adrian Riekert. A proof of convergence for stochastic gradient descent in the training of artificial neural networks with ReLU activation for constant target functions, 2021. arXiv:2104.00277. [26] Arnulf Jentzen and Adrian Riekert. A proof of convergence for the gradient descent opti- mization method with random initializations in the training of neural networks with ReLU activation for piecewise linear target functions, 2021. arXiv:2108.04620. [27] Tobias Kaiser. Integration of semialgebraic functions and integrated Nash functions. Math. Z., 275(1-2):349–366, 2013. doi:10.1007/s00209-012-1138-1. [28] Hamed Karimi, Julie Nutini, and Mark Schmidt. Linear convergence of gradient and proximal-gradient methods under the Polyak-Lojasiewicz condition, 2020. arXiv:1608.0 4636.

29 [29] Krzysztof Kurdyka, Tadeusz Mostowski, and Adam Parusi´nski. Proof of the gradient conjecture of R. Thom. Ann. of Math. (2), 152(3):763–792, 2000. doi:10.2307/2661354.

[30] Jason D. Lee, Ioannis Panageas, Georgios Piliouras, Max Simchowitz, Michael I. Jordan, and Benjamin Recht. First-order methods almost always avoid strict saddle points. Math. Program., 176(1–2):311–337, July 2019. doi:10.1007/s10107-019-01374-3.

[31] Jason D. Lee, Max Simchowitz, Michael I. Jordan, and Benjamin Recht. Gradient descent only converges to minimizers. In Vitaly Feldman, Alexander Rakhlin, and Ohad Shamir, editors, 29th Annual Conference on Learning Theory, volume 49 of Proceedings of Machine Learning Research, pages 1246–1257, Columbia University, New York, New York, USA, 23–26 Jun 2016. PMLR. URL: http://proceedings.mlr.press/v49/lee16.html.

[32] Y. Lei, T. Hu, G. Li, and K. Tang. Stochastic gradient descent for nonconvex learning with- out bounded gradient assumptions. IEEE Transactions on Neural Networks and Learning Systems, 31(10):4394–4400, 2020. doi:10.1109/TNNLS.2019.2952219.

[33] S.Lojasiewicz. Sur les trajectoires du gradient d’une fonction analytique. In Geometry seminars, 1982–1983 (Bologna, 1982/1983), pages 115–117. Univ. Stud. Bologna, Bologna, 1984.

[34] Eric Moulines and Francis Bach. Non- of stochastic approximation algorithms for machine learning. In J. Shawe-Taylor, R. Zemel, P. Bartlett, F. Pereira, and K. Q. Weinberger, editors, Advances in Neural Information Processing Systems, volume 24, pages 451–459. Curran Associates, Inc., 2011. URL: https://proceedings.neurips.cc /paper/2011/file/40008b9a5380fcacce3976bf7c08af5b-Paper.pdf.

[35] Yurii Nesterov. Introductory lectures on convex optimization, volume 87 of Applied Op- timization. Kluwer Academic Publishers, Boston, MA, 2004. A basic course. doi: 10.1007/978-1-4419-8853-9.

[36] Peter Ochs. Unifying abstract inexact convergence theorems and block coordinate variable metric iPiano. SIAM J. Optim., 29(1):541–570, 2019. doi:10.1137/17M1124085.

[37] Vivak Patel. Stopping criteria for, and strong convergence of, stochastic gradient descent on Bottou-Curtis-Nocedal functions, 2021. arXiv:2004.00475.

[38] Alexander Rakhlin, Ohad Shamir, and Karthik Sridharan. Making gradient descent opti- mal for strongly convex stochastic optimization. In Proceedings of the 29th International Conference on Machine Learning, page 1571–1578, Madison, WI, USA, 2012. Omnipress.

[39] R. Tyrrell Rockafellar and Roger J.-B. Wets. Variational analysis, volume 317 of Grundlehren der Mathematischen Wissenschaften. Springer-Verlag, Berlin, 1998. doi: 10.1007/978-3-642-02431-3.

[40] Filippo Santambrogio. {Euclidean, metric, and Wasserstein} gradient flows: an overview. Bull. Math. Sci., 7(1):87–154, 2017. doi:10.1007/s13373-017-0101-1.

[41] Guodong Zhang, James Martens, and Roger B Grosse. Fast convergence of natural gradient descent for over-parameterized neural networks. In H. Wallach, H. Larochelle, A. Beygelzimer, F. d'Alch´e-Buc,E. Fox, and R. Garnett, editors, Advances in Neural In- formation Processing Systems 32, pages 8082–8093. Curran Associates, Inc., 2019. URL: http://papers.nips.cc/paper/9020-fast-convergence-of-natural-gradient- descent-for-over-parameterized-neural-networks.pdf.

30