arxiv.org · 2020. 12. 14. · arXiv:2012.06244v1 [cs.LG] 11 Dec 2020 THE IMPLICIT BIAS FOR...

46
arXiv:2012.06244v1 [cs.LG] 11 Dec 2020 T HE I MPLICIT B IAS FOR A DAPTIVE O PTIMIZATION A LGORITHMS ON H OMOGENEOUS N EURAL N ETWORKS APREPRINT Bohan Wang Microsoft Research Asia, Beijing, China [email protected] Qi Meng Microsoft Research Asia, Beijing, China [email protected] Wei Chen Microsoft Research Asia, Beijing, China [email protected] December 14, 2020 ABSTRACT Despite their overwhelming capacity to overfit, deep neural networks trained by specific optimiza- tion algorithms tend to generalize relatively well to unseen data. Recently, researchers explained it by investigating the implicit bias of optimization algorithms. A remarkable progress is the work [18], which proves gradient descent (GD) maximizes the margin of homogeneous deep neural net- works. Except the first-order optimization algorithms like GD, adaptive algorithms such as AdaGrad, RMSProp and Adam are popular owing to its rapid training process. Meanwhile, numerous works have provided empirical evidence that adaptive methods may suffer from poor generalization perfor- mance. However, theoretical explanation for the generalization of adaptive optimization algorithms is still lacking. In this paper, we study the implicit bias of adaptive optimization algorithms on ho- mogeneous neural networks. In particular, we study the convergent direction of parameters when they are optimizing the logistic loss. We prove that the convergent direction of RMSProp is the same with GD, while for AdaGrad, the convergent direction depends on the adaptive conditioner. Tech- nically, we provide a unified framework to analyze convergent direction of adaptive optimization algorithms by constructing novel and nontrivial adaptive gradient flow and surrogate margin. The theoretical findings explain the superiority on generalization of exponential moving average strategy that is adopted by RMSProp and Adam. To the best of knowledge, it is the first work to study the convergent direction of adaptive optimizations on non-linear deep neural networks. Keywords Adaptive Optimizer · Implicit Bias · Margin 1 Introduction Deep learning techniques have been very successful in several domains, like computer vision [26], speech recognition [5] and natural language processing [30]. In practice, deep neural networks (DNN) learned by some optimization algorithms such as gradient descent (GD) and its variants can generalize well to unseen data [28]. However, deep neural networks are non-convex. The non-convex deep neural networks have been found to have large amount of global minima [4], while only few of them can guarantee satisfactory generalization property [3]. Explaining why the highly non-convex model trained by a specific algorithm can generalize has become an important open question in deep learning. Regarding the above question, one plausible explanation is that optimization algorithms implicitly bias training [22]. That is, the optimization algorithm tends to drive parameters to certain kinds of global minima which generalize well, although no explicit regularization is enforced. Recently, exciting results have been shown for vanilla gradient descent. For linear classifier, GD is proved to converge to minimum L 2 norm solution for squared loss and max-margin solution for logistic loss [25]. [10] extend the result for GD to deep linear neural networks. A remarkable progress is the work [18], which proves GD maximizes the margin of homogeneous deep neural networks.

Transcript of arxiv.org · 2020. 12. 14. · arXiv:2012.06244v1 [cs.LG] 11 Dec 2020 THE IMPLICIT BIAS FOR...

Page 1: arxiv.org · 2020. 12. 14. · arXiv:2012.06244v1 [cs.LG] 11 Dec 2020 THE IMPLICIT BIAS FOR ADAPTIVE OPTIMIZATION ALGORITHMS ON HOMOGENEOUS NEURAL NETWORKS A PREPRINT Bohan Wang Microsoft

arX

iv:2

012.

0624

4v1

[cs

.LG

] 1

1 D

ec 2

020

THE IMPLICIT BIAS FOR ADAPTIVE OPTIMIZATION

ALGORITHMS ON HOMOGENEOUS NEURAL NETWORKS

A PREPRINT

Bohan WangMicrosoft Research Asia,

Beijing, [email protected]

Qi MengMicrosoft Research Asia,

Beijing, [email protected]

Wei ChenMicrosoft Research Asia,

Beijing, [email protected]

December 14, 2020

ABSTRACT

Despite their overwhelming capacity to overfit, deep neural networks trained by specific optimiza-tion algorithms tend to generalize relatively well to unseen data. Recently, researchers explainedit by investigating the implicit bias of optimization algorithms. A remarkable progress is the work[18], which proves gradient descent (GD) maximizes the margin of homogeneous deep neural net-works. Except the first-order optimization algorithms like GD, adaptive algorithms such as AdaGrad,RMSProp and Adam are popular owing to its rapid training process. Meanwhile, numerous workshave provided empirical evidence that adaptive methods may suffer from poor generalization perfor-mance. However, theoretical explanation for the generalization of adaptive optimization algorithmsis still lacking. In this paper, we study the implicit bias of adaptive optimization algorithms on ho-mogeneous neural networks. In particular, we study the convergent direction of parameters whenthey are optimizing the logistic loss. We prove that the convergent direction of RMSProp is the samewith GD, while for AdaGrad, the convergent direction depends on the adaptive conditioner. Tech-nically, we provide a unified framework to analyze convergent direction of adaptive optimizationalgorithms by constructing novel and nontrivial adaptive gradient flow and surrogate margin. Thetheoretical findings explain the superiority on generalization of exponential moving average strategythat is adopted by RMSProp and Adam. To the best of knowledge, it is the first work to study theconvergent direction of adaptive optimizations on non-linear deep neural networks.

Keywords Adaptive Optimizer · Implicit Bias · Margin

1 Introduction

Deep learning techniques have been very successful in several domains, like computer vision [26], speech recognition[5] and natural language processing [30]. In practice, deep neural networks (DNN) learned by some optimizationalgorithms such as gradient descent (GD) and its variants can generalize well to unseen data [28]. However, deepneural networks are non-convex. The non-convex deep neural networks have been found to have large amount ofglobal minima [4], while only few of them can guarantee satisfactory generalization property [3]. Explaining why thehighly non-convex model trained by a specific algorithm can generalize has become an important open question indeep learning.

Regarding the above question, one plausible explanation is that optimization algorithms implicitly bias training [22].That is, the optimization algorithm tends to drive parameters to certain kinds of global minima which generalize well,although no explicit regularization is enforced. Recently, exciting results have been shown for vanilla gradient descent.For linear classifier, GD is proved to converge to minimumL2 norm solution for squared loss and max-margin solutionfor logistic loss [25]. [10] extend the result for GD to deep linear neural networks. A remarkable progress is the work[18], which proves GD maximizes the margin of homogeneous deep neural networks.

Page 2: arxiv.org · 2020. 12. 14. · arXiv:2012.06244v1 [cs.LG] 11 Dec 2020 THE IMPLICIT BIAS FOR ADAPTIVE OPTIMIZATION ALGORITHMS ON HOMOGENEOUS NEURAL NETWORKS A PREPRINT Bohan Wang Microsoft

A PREPRINT

On the other hand, adaptive algorithms such as AdaGrad [6], RMSProp [9], and Adam [13] have been in spotlightthese years. These algorithms are proposed to improve the convergence rate of GD (or SGD) by using second-ordermoments of historical gradients and have been widely applied in deep learning [24]. Despite the rapid convergenceof adaptive methods, numerous works have provided empirical evidence that adaptive methods may suffer from poorgeneralization performance [27, 17]. Specifically, it has been observed that first-order method such as SGD canoutperform adaptive methods on the test set, even when the adaptive methods achieve the same or lower training loss.Several works try to improve the performance of adaptive optimization algorithms such as AdamW [16], AdaBound[17], AdaBelief [32]. However, there is little theoretical analysis for generalization of adaptive algorithms. Theseobservations and the research for GD motivate us to study the implicit bias for adaptive algorithms.

The key factor for the success of adaptive algorithms is to design better conditioners of the gradient. There aretwo popular kinds of conditioners: one is the simple average of the squared values of past gradients that is used inAdaGrad, and the other is the exponential moving average that is used in RMSProp, Adam, etc. In this paper, weaim to study the influence of these two kinds of conditioners on convergent direction of parameters on deep neuralnetworks. Specifically, we work on the homogeneous neural networks with separable data under logistic regressionsetting. Unlike the squared loss, the logistic doss does not admit a finite global minimizer on separable data: theiterates will diverge in order to drive the loss to zero. Therefore, researchers use the asymptotic direction of algorithms(i.e., limt→∞

wt

‖wt‖2) for logistic loss instead of characterizing the exact limit of parameters [25, 8].

The main result and our technical contributions are summarized as follows.

Theorem 1. (Informal) We use Φ(w,x) to denote the homogeneous neural network model with parameter w andinput x. (1) For AdaGrad, any limit point of wt/‖wt‖2 is the KKT point of the optimization problem

min ‖h1/2∞ ⊙w‖ subject to yiΦ(w,xi) ≥ 1, ∀i, (1)

where h∞ = limt→∞ h(t) is the limit of the conditioner in AdaGrad. (2) For RMSProp, any limit point ofw(t)/‖w(t)‖2 is the KKT point of the optimization problem

min ‖w‖2 subject to yiΦ(w,xi) ≥ 1, ∀i. (2)

Theorem 1 provides several interesting insights. RMSProp converges to the direction which is independent with theconditioner while AdaGrad does not. The limit direction of RMSProp is the same with gradient descent, which isthe max-margin solution and indicates good generalization. It matches the empirical observations that RMSPropgeneralize better than AdaGrad on deep neural networks [27]. The reason is the exponential weighted average strategy

(i.e., m(t+ 1) = bm(t) + (1 − b)∂L2, where√m(t) + ε = h(t)−1) in RMSProp makes the part m(t) converge to

zero, while in AdaGrad, the update rule is m(t + 1) = m(t) + ∂L2, which makes m(t) keep growing and rely onthe optimization process. Therefore, adaptive algorithms that adopt exponential weighted average strategy achievestradeoff between optimization efficiency and generalization guarantee, i.e., it accelerates the training process at earlystage of optimization by adaptively adjusting the learning rate, but it still converges to low-complexity model (i.e.,max margin solution) because the conditioner converges to constant at later stage.

In order to prove the Theorem 1, we first propose adaptive gradient flow, which is a unified framework to deal withadaptive gradients. It transforms original parameter space to a parameter space normalized by the conditioners. Second,in the normalized parameter space, we construct surrogate margin for the adaptive algorithms. The surrogate marginis defined as the ratio of negative logarithmic loss normalized by parameter norm. We prove that surrogate margin islower bounded, and then the increasing rate of parameter norm can then be bounded by the decreasing rate of loss [8],which together drive the loss to zero. At last, we can prove any limit direction of the normalized parameter flow is theKKT point of the margin maximization problem in normalized parameter space. Moreover, we prove the convergentdirection is unique if the neural network is definable [14].

Technically, although we adopt the framework of [18], the techniques for vanilla GD can no longer be applied toadaptive algorithms. We highlight our technical contributions as: (1) we construct adaptive gradient flow, whichis novel and non trivial. With the adaptive gradient flow, the analysis of convergent direction is transformed fromoriginal parameter space to a normalized parameter space. The adaptive gradient flow can be applied into analysis

of other adaptive algorithms. (2) Previous works [25, 18] prove for vanilla GD, smoothed margin defined aslog 1

L(w)

‖w‖L

monotonously increases, which however, in adaptive methods, no longer holds. Instead, we define the surrogatemargin which take the conditioner into account, and prove the surrogate margin (i) is lower bounded; (ii) has the samelimit with the normalized margin.

2

Page 3: arxiv.org · 2020. 12. 14. · arXiv:2012.06244v1 [cs.LG] 11 Dec 2020 THE IMPLICIT BIAS FOR ADAPTIVE OPTIMIZATION ALGORITHMS ON HOMOGENEOUS NEURAL NETWORKS A PREPRINT Bohan Wang Microsoft

A PREPRINT

2 Related Work

Implicit Bias of First-order Optimization Methods. [25] proved that gradient descent on linear logistic regressionwith separable data converges in the direction of the maxL2 margin solution of the corresponding hard-margin SupportVector Machine, and motivate a line of works on the bias of GD on linear model. For example, [20] studied other losswith different tails; [11] proved similar result without assuming separable data; [15] worked on the problem in theadversarial scenario; [29] analyzed the case when there is only a ReLU neuron.

Then, there are some works study the implicit bias of GD on deep neural networks. The asymptotic direction of theproduct of weights in a deep linear network is studied in [10, 8]. [25] studied the implicit bias of two-layer neuralnetwork with ReLU activation. [19] proved the asymptotic direction is along a KKT point of the L2 max marginproblem for homogeneous deep neural networks. [18] independently proved similar result for homogeneous neuralnetworks with simplified assumptions. Based on [18], [12] further prove that parameters have only one asymptoticdirection.

There are also works considering implicit bias of other first-order optimization algorithms. [21] worked on StochasticGradient Descent for linear logistic regression. [7] studied the implicit bias of mirror descent, natural gradient descent,and steepest descent on linear model. [1] proved gradient descent on Neural Tangent Kernel will converge linearly toa global minimum near the initial point.

Theoretical Evidence of Generalization of Adaptive Algorithms. [27] constructed a linear regression task whereAdaGrad, RMSProp, and ADAM converge to a solution that incorrectly classifies new data with probability arbitrarilyclose to half and SGD finds a solution with zero error on new data. [31] modeled the distribution of stochastic noise inAdam, and they showed that SGD tends to converge to flatter local minima. Another viewpoint is to study the implicitbias of adaptive optimization algorithms. To the best of our knowledge, the only work on the bias of adaptivealgorithms is [23], which proves AdaGrad has different asymptotic direction with GD on linear logistic regression.In this paper, we study the asymptotic direction of adaptive optimization algorithms on deep neural networks whichrequires different techniques due to high non-convex of deep networks.

3 Preliminaries

In this paper, we study the logistic regression problem with homogeneous neural networks. Let training set S definedas S = (xi, yi)Ni=1, where xi ∈ X (i = 1, 2, · · · , N ) are inputs, yi ∈ R (i = 1, 2, · · · , N ) are labels, and N is the

size of S. The empirical loss L with training set S, neural network classifier Φ, individual loss ℓ(x) = e−f(x) andparameters w ∈ Rp can be written as follows:

L(w,S) =N∑

i=1

ℓ (yiΦ(w,xi)) .

In an optimization process, the training set S is fixed. Therefore, without loss of generality, we abbreviate L(w,S) =L(w), and yiΦ(w,xi) = qi(w). In this paper, we consider the exponential loss, i.e., f(qi(w)) = qi(w), and the

logistic loss, i.e., f(qi(w)) = − log log(1 + e−qi(w)). Both of f are monotonously increasing and have an inverse.

As for optimization methods, we consider AdaGrad and RMSProp. The update rule for the two algorithms can bewritten as

w(t+ 1)−w(t) = −η(t)h(t) ⊙ ∂L(w(t)), (3)

where h(t) is called the conditioner which adaptively assigns different learning rates for different coordinates. For

AdaGrad, h(t)−1 =√

ε1p +∑t

τ=0 ∂L(w(τ ))2 where ǫ is a positive constant. Here, ∂L(w(τ))2 = L(w(τ))⊙L(w(τ))

and ⊙ denotes the element-wised product of a vector. Different from AdaGrad, RMSProp adopts exponential weighted

average stategy in h(t), i.e., h(t)−1 =√

ε1p +∑t

τ=0(1− b)bt−τ ∂L(w(τ ))2. For clarity, we use hA(t) and hR(t) to

distinguish the term h(t) in AdaGrad and RMSProp, respectively.

We consider the flow of the two optimization algorithms, i.e.,

dw(t)

dt= −h(t)⊙ ∂L(w(t)), (4)

which is the continuous version of Eq.3. For the flow, the conditioner has the form hA(t)−1 =

ε1p +∫ t

0∂L(w(τ ))2dτ

and hR(t)−1 =

ε1p +∫ t

0(1− b)e−(1−b)(t−τ)∂L(w(τ ))2dτ . Our study will start with the continuous version of the two

algorithms, which is easier to analyze.

3

Page 4: arxiv.org · 2020. 12. 14. · arXiv:2012.06244v1 [cs.LG] 11 Dec 2020 THE IMPLICIT BIAS FOR ADAPTIVE OPTIMIZATION ALGORITHMS ON HOMOGENEOUS NEURAL NETWORKS A PREPRINT Bohan Wang Microsoft

A PREPRINT

Specifically, for the continuous case, we focus on the following scenario:

Assumption 1. The empirical loss is defined as L(w) =∑N

i=1 e−f(qi(w)). The following propositions hold:

I (Regularity). For any i, qi is locally Lipchitz and admits a chain rule;

II (Homogeneity). There exists L > 0 such that ∀α > 0 and i, qi(αw) = αLqi(w);

III (Separability). There exists a time t0 such that f−1(log 1L(t0)

) > 0.

I, II in Assumption 1 holds for a board class of networks allowing for ReLU, max pooling, and convolutional lay-ers; Assumption 1.III holds generally for over-parameterized neural networks, which can achieve complete correctclassification in training set.

At the end of this section, we give a brief introduction to KKT conditions and KKT points. For a constrained optimiza-tion problem defined as

min f(v) subject to: gi(v) ≤ 0, ∀i ∈ [N ],

KKT conditions are necessary conditions for a feasible point v0 to be an optimal point. Concretely, KKT conditionsrequire that there exists non-negative reals λi, such that

∂f(v0) +

N∑

i=1

λi∂gi(v0) = 0; (5)

N∑

i=1

λigi(v0) = 0. (6)

A feasible point is said to be a KKT point if it satisfies KKT condition.

A weaker notion of KKT condition is (ε, δ) KKT condition, which only requires left sides of eq. (5) and (6) to berespectively smaller than ε and δ. We will formally define (ε, δ) KKT points and give some of their properties inAppendix A.2.

Notations. In this paper, we use O, Θ, and Ω to hide the absolute multiplicative factors. We also use ∂f to denote thesubgradient of a function f .

4 Adaptive Gradient Flow: Definition and Verification

In this section, we will prove some common properties of AdaGrad flow and RMSProp flow. Based on these properties,we will define the general adaptive gradient flow, and present our convergent direction results on adaptive gradientflows. Then, the convergent direction of AdaGrad and RMSProp can be derived as a special case of adaptive gradientflow. In the end of this section, we will show that under additional assumptions, the convergent direction is unique.

We first prove the following properties of AdaGrad flow.

Theorem 2. For AdaGrad flow defined as eq. (4) with h(t) = hA(t), we have that

• hA(t) converges as t→ ∞. Furthermore, limt→∞ hA(t) has no zero component.

• If we let h∞ = limt→∞ hA(t), and define

vA(t) = h−1/2∞ ⊙w(t),

βA(t) = h−1∞ ⊙ h(t),

LA(vA) = L(h12∞ ⊙ vA),

we have

dvA(t)

dt= −βA(t)⊙ ∂LA(vA(t)).

•d logβA(t)

dt is Lebesgue integrable with respect to t.

4

Page 5: arxiv.org · 2020. 12. 14. · arXiv:2012.06244v1 [cs.LG] 11 Dec 2020 THE IMPLICIT BIAS FOR ADAPTIVE OPTIMIZATION ALGORITHMS ON HOMOGENEOUS NEURAL NETWORKS A PREPRINT Bohan Wang Microsoft

A PREPRINT

We make some explanations to above properties: The former part of the first property is natural, since hA(t) isnon-increasing with respect to t. However, the latter part yields that integration of square of the gradient convergesto a positive real, which is non-trivial; the second property is obtained by component-wisely scaling w and directverification; the last property can be obtained by Newton-Leibnz formula for absolutely continuous function sincedβA(t)

dt is non-negative, and guarantees that the change of βA will not be too severe. We defer the detailed proof toAppendix B.1.

Similar properties also hold for RMSProp as the following lemma.

Theorem 3. For RMSProp flow defined as eq. (4) with h(t) = hR(t), we have that

• hR(t) converges as t→ ∞. Furthermore, limt→∞ hR(t) = ε−12 1

Tp .

• If we let h∞ = limt→∞ hR(t), and define

vR(t) = ε14w(t),

βR(t) = ε12h(t),

LR(vR) = L(ε− 14 vR),

we havedvR(t)

dt= −βR(t)⊙ ∂LR(vR(t)).

•d logβR(t)

dt is Lebesgue integrable with respect to t.

The conditioner hR has a decay term e−(t−τ)(1−b), which drives∫ t

τ=0(1 − b)e−(1−b)(t−τ)∂L(w(τ))2dτ to zero, andthe conditioner to isotropy. The detailed proof requires a more careful analysis in measure than the AdaGrad flow. Wedefer them to Section B.1.

Based on the properties they share, we can define the following adaptive gradient flow for more general adaptiveoptimizers.

Definition 1. A function w(t) is called to obey an adaptive gradient flow F with loss L and component learning rateβ(t), if it can be written as the following form

dw(t)

dt= −β(t)⊙ ∂L(w(t)),

where β(t) satisfies that limt→∞ β(t) = 1p, andd logβ(t)

dt is Lebesgue Integrable.

By Theorems 2 and 3, AdaGrad flow and RMSProp can both be transferred into adaptive gradient flows: vA obeys

an adaptive gradient flow with loss LA and conditioner βA, while vR obeys an adaptive gradient flow with loss LR

and conditioner βR. Furthermore, Assumption 1 also holds for LA and LR: for both AdaGrad and RMSProp, we can

uniformly represent LA(vA) and LR(vR) as L(h 12 ⊙ v), where h is a component-wisely positive constant vector. By

Assumption 1, L(v) can be further written as

L(v) =L(h 12 ⊙ v)

=N∑

i=1

e−f(qi(h12 ⊙v)).

If we denote qi(v) = qi(h12 ⊙ v), we have qi is also an L homogeneous function, and L(v) =

∑Ni=1 e

−f(qi(v)).

Therefore, we only consider adaptive gradient flow satisfying Assumption 1. Our main result is the following theorem:

Theorem 4. Let v satisfy an adaptive gradient flow F which satisfies Assumption 1. Let v be any limit point of

v(t)∞t=0 (where v(t) = v(t)‖v(t)‖ is normalized parameter). Then v is along the direction of a KKT point of the

following L2 max-margin problem (P ):

min1

2‖v‖2

subject to qi(v) ≥ 1, ∀i ∈ [N ].

5

Page 6: arxiv.org · 2020. 12. 14. · arXiv:2012.06244v1 [cs.LG] 11 Dec 2020 THE IMPLICIT BIAS FOR ADAPTIVE OPTIMIZATION ALGORITHMS ON HOMOGENEOUS NEURAL NETWORKS A PREPRINT Bohan Wang Microsoft

A PREPRINT

(P ) is equivalent to the L2 max-margin problem: suppose v0 is an optimal point of (P ). Then there exists an i ∈ [N ],

such that, qi(v0) = 1 (otherwise, we can let v′0= v0/qmin(v0)

1L . Then v′

0is also a fixed point of (P ) and have

a smaller L2 norm than v′0

, which leads to contradictory). Therefore, qmin(v0) = 1 (qmin(v)= miniqi(v)), and

maximizing the normalized marginqmin(v)‖v‖L is equivalent to minimize ‖v‖L.

Theorem 4 shows that the adaptive gradient flow actually drives the parameters to solutions of L2 max-margin problem.We will give the proof skeleton of Theorem 4 in Section 5.

Combining Theorem 4 with Theorems 2 and 3, one can obtain convergent directions of AdaGrad flow and RMSPropflow by simple parameter substitution of (P ).

Theorem 5. Let w satisfy AdaGrad flow defined as eq. (4) with h(t) = hA(t). Then, any limit point of w(t)∞t=0

(where w(t) = w(t)‖w(t)‖ is normalized parameter) is along the direction of a KKT point of the following optimization

problem (PA):

min1

2‖h− 1

2∞ ⊙w‖Subject to: qi(w) ≥ 1.

Theorem 6. Let w satisfy RMSProp flow defined as eq. (4) with h(t) = hR(t). Then, any limit point of w(t)∞t=0

(where w(t) = w(t)‖w(t)‖ is normalized parameter) is along the direction of a KKT point of the following optimization

problem (PR):

min1

2‖w‖

Subject to: qi(w) ≥ 1.

Intuitively, (PR) is the L2 max-margin problem, which means RMSProp flow biases parameters to a local minimumwith good generalization property; on the other hand, the target of (PA) has a reliance of h∞, which can be influencedby the optimization process and initialization, and may further lead to worse generalization. We will discuss thedifference between convergent directions of AdaGrad and RMSProp in detail in Section 6.1.

5 Proof Sketch of Theorem 4

In this section, we present the proof sketch of Theorem 4. The proof can be divided into three stages: (I). An surrogatemargin is constructed, and we prove it is lower bounded and equivalent to normalized margin as time tends to infinity;

(II). We use this surrogate margin to lower bound the decreasing rate of empirical loss L, and prove limt→∞ L(w(t)) =0; (III). For every convergent direction v, a series of (εi, δi) KKT point which converges to v with limi→∞ εi =limi→∞ δi = 0 is constructed. We then show every convergent direction is a KKT point of optimization problem (P ).

5.1 surrogate margin on adaptive gradient flow

By homogeneity of qi, for any parameter v, 〈∂qi(v),v〉 = Lqi(v). Therefore, the inner product between ∂L(v) =

−∑i∈[N ] e−f(qi(v))f ′(qi(v))∂qi(v) and v has a clear form that −〈∂L(v),v〉 = L

∑i∈[N ] e

−f(qi(v))f ′(qi(v))qi(v).

However, for adaptive gradient flow, ∂L(v(t)) = −β(t) ⊙ dv(t)dt , and we can no longer adopt the above method to

calculate the change of ‖v‖. To solve the problem, we study the surrogate norm ρ(t)= ‖β(t)− 1

2 ⊙ v(t)‖. Sincelimt→∞ β = 1, ρ(t) = Θ(‖v(t)‖).

The normalized margin γ(t) = qmin(t)‖v(t)‖L connects margin qmin(t) with parameter norm ‖v(t)‖. However, since normal-

ized margin is derived by taking minimum among all qi, it requires huge efforts to calculate and estimate derivative of

γ(t). Fortunately, the next lemma admits us to define an surrogate margin defined using L.

Lemma 1. If limt→∞ L = 0, we have thatg(log 1

˜L(v(t)))

ρ(t) = Θ(γ(t)), where g is the inverse of function f in Assumption

1.

Based Lemma 1, we can defined our surrogate margin as follows:

6

Page 7: arxiv.org · 2020. 12. 14. · arXiv:2012.06244v1 [cs.LG] 11 Dec 2020 THE IMPLICIT BIAS FOR ADAPTIVE OPTIMIZATION ALGORITHMS ON HOMOGENEOUS NEURAL NETWORKS A PREPRINT Bohan Wang Microsoft

A PREPRINT

Definition 2. Let a function v(t) satisfies an adaptive gradient flow F with loss L and component learning rate β(t),

where L satisfies Assumption 1. Then the surrogate margin γ is defined as

γ(t) =g(log 1

L(v(t)))

ρ(t)L.

Since ρ(t) = Θ(‖v(t)‖), γ actually bridge the norm of parameters with empirical loss. A desired property for γ isto have a positive lower bound, since with this property, one can further bound parameter norm using empirical loss.

The following lemma shows that γ is lower bounded for adaptive gradient flow F with empirical loss L satisfyingAssumption 1.

Lemma 2. Let a function v(t) satisfies an adaptive gradient flow F with loss L and component learning rate β(t),

where L satisfies Assumption 1. Then there exists a time t1 ≥ t0, such that, for any time t ≥ t1, γ(t) ≥ e−12 γ(t1).

Here we briefly give a road map of the proof. The derivative of norm ρ(t) can be split into two parts: one is the increas-

ing of parameter v along the direction of β− 12 (t), and another is the change of component learning rate β− 1

2 (t). Ap-

plying homogeneity of qi and Cauchy Inequality, we bound the first term using the increasing rate of g(log(

1L(t)

)),

which is exactly derivative of numerator of γ; while for the second term, we show that it can be lower bounded by∑p

i=1

(d logβ

−12

i (t)

dt

)

+

. By the definition of adaptive gradient flow, integration of the second term tend to zero as time

approaches infinity. The proof is completed by putting two parts together.

By the discussion above, one can observe that derivative of γ(t) can be calculated by subtracting a small enough term

∑pi=1

(d logβ

−12

i (t)

dt

)

+

from a non-negative term. This observation leads to the the convergence of γ(t).

Lemma 3. Suppose a function v obey an adaptive gradient flow F , which satisfies Assumption 1. Then the surrogatemargin γ(t) converges.

5.2 Convergence of Empirical Loss and Parameters

By Lemma 2, we have that for an adaptive gradient flow F with Assumption 1, the norm ρ(t) can be bounded as

ρ(t) ≤ e12L g(log( 1

L(v(t))))

1L . On the other hand, by chain rule, the derivative of empirical loss with respect to time

can be calculated as

dL(v(t))dt

= 〈∂L(v(t)), dv(t)dt

〉 = −‖β 12 (t)⊙ ∂L(v(t))‖2 ≤ −〈∂L(v(t)),v(t)〉2

ρ(t)2,

where the last inequality is derived by the Cauchy inequality applying to 〈β 12 ⊙ ∂L,β− 1

2 ⊙ v〉. By the homogeneity

of qi, we can further lower bound〈∂L,v〉2

ρ2 using L. In other words, Lemma 2 ensures that the decreasing rate of the

empirical loss L can be lower bounded by a function of itself. Based on the above methodology, we can prove thefollowing lemma and show that empirical loss will decrease to zero, while parameter norm will converge to infinity.

Lemma 4. Let a function v(t) satisfies an adaptive gradient flow F with loss L and component learning rate β(t),

where L satisfies Assumption 1. Then, limt→∞ L(v(t)) = 0, and consequently, limt→∞ ‖v(t)‖ = ∞.

Here the proposition that parameter norm goes to infinity can be derived naturally by empirical loss going to zero,since f(qi(v)) is upper bounded within any bounded set of v.

5.3 Convergence to KKT point

We start by proving for any t ≥ t1, v(t) = v(t)‖v(t)‖ is an approximate KKT point. Based on the surrogate margin that

we construct in Section 5.1, we can provide a construction of coefficients λi as follows:

Lemma 5. Let v and∂L(v) be v and ∂L(v) respectively normalized by their L2 norms. Then v(t) is a(

O(1 − 〈v(t), −∂L(t)〉),O(

1log 1

L(t)

))KKT point of optimization problem (P ) in Theorem 4.

7

Page 8: arxiv.org · 2020. 12. 14. · arXiv:2012.06244v1 [cs.LG] 11 Dec 2020 THE IMPLICIT BIAS FOR ADAPTIVE OPTIMIZATION ALGORITHMS ON HOMOGENEOUS NEURAL NETWORKS A PREPRINT Bohan Wang Microsoft

A PREPRINT

We made some explanations to Lemma 5: by the results in Section 5.2, we have limt→∞ O(

1log 1

L(t)

)= 0. Therefore,

we only need to find a convergent series v(t) with 1 + 〈v(t),∂L(t)〉 goes to zero.

For this purpose, we construct another approximate norm

ρ(t) =

√√√√∥∥∥β− 1

2 (t)⊙ v(t)∥∥∥2

− 2

∫ t

t1

⟨v(τ),β− 1

2 (τ) ⊙ dβ− 12

dt(τ)⊙ v(τ)

⟩dτ,

which measures the increasing of v(t). Recall that we separate the change of ρ into two parts: change of v and changeof β in order to prove that the surrogate margin is lower bounded. While keeping the second part in ρ(t), we com-

bine the change of v (which is −Θ(d log ρdt )) with the increasing of g(log( 1

L )) (which is Θ(d log ρdt 〈v(t),−∂L(t)〉−2)).

Combining the methodology above together, we have the following lemma:

Lemma 6. For any t3 > t2 ≥ t1, there exists a ξ ∈ [t2, t3], such that(〈v(ξ),−∂L(ξ)〉−2 − 1

)(log ρ(t2)− log ρ(t1)) ≤C1 log

γ (t3)

γ (t2)

+C2

p∑

i=1

∫ t3

t2

p∑

i=1

(d logβ

− 12

i (t)

dt

)

+

dt, (7)

where C1, C2 are positive constants which is independent of t2 and t3.

The right side of inequality 7 is bounded and goes to zero as t2 → ∞ although log ρ(t3)− log ρ(t2) can be arbitrarily

large, which indicates that there exists ξ ∈ [t2, t3] that makes

(〈v(ξ),−∂L(ξ)〉−2 − 1

)is small enough, which is

equivalent to

(1− 〈v(ξ),−∂L(ξ)〉

)is small enough.

However, ξ is guaranteed by First Mean Value Theorem, and does not have a close form. Therefore, we furtherintroduce another lemma to bound ‖v(ξ)− v(t2)‖.

Lemma 7. For any time t ≥ t1, we have that

∥∥∥∥dv(t)

dt

∥∥∥∥ ≤ 3 · 2L−1

(1− e12

2 )

e12B1

L

d log ρ

dt

1

γ(t1).

Combining Lemma 5, 6, and 7, for any convergent direction v, we can construct a series of ti∞i=1, such that v(ti) is(εi, δi) KKT point, with limi→∞ v(ti) = v, and limi→∞ εi = limi→∞ δi = 0. On the other hand, constraints of (P )satisfies Mangasarian-Fromovitz constraint qualification (see Appendix A.2), which ensures that v is a KKT point of(P ), and complete the proof.

6 Discussions

6.1 RMSProp generalizes better than AdaGrad

In this section, we make some discussions on the results derived in section 4.

First, Theorem 6 and Theorem 5 shows that RMSProp implicitly bias the solution to ones with maximized surrogatemargin, while AdaGrad may drive the parameters to a different direction. The corresponding optimization problemof AdaGrad has a reliance on h∞, which is shown to be sensitive to the optimization path before convergence. Be-cause the normalized margin is used as a complexity norm in generalization literature (i.e., larger normalized marginindicating better generalization performance) [2], we conclude that RMSProp generalizes better than AdaGrad. Thisconclusion matches the empirical observations [17].

Second, two key factors that guarantee generalization of RMSProp are exponential weighted average strategy and theadded constant ǫ in h(t). Our results show the benefit of the two factors: it accelerates the training process at earlystage of optimization by adaptively adjusting the learning rate, but it still converges to max margin solution becausethe denominator of conditioner tends to constant ǫ at later stage. Most of previous works explain ǫ to ensure positivity

8

Page 9: arxiv.org · 2020. 12. 14. · arXiv:2012.06244v1 [cs.LG] 11 Dec 2020 THE IMPLICIT BIAS FOR ADAPTIVE OPTIMIZATION ALGORITHMS ON HOMOGENEOUS NEURAL NETWORKS A PREPRINT Bohan Wang Microsoft

A PREPRINT

of h(t). Our results show that ǫ is important for the convergent direction of the parameters and the generalizationability. Our results also indicates generalization guarantee for other adaptive optimization algorithms such as Adam,AdamW which adopts the same design of h(t) with RMSProp.

6.2 Gap between continuous flow and discrete update rule

Our main results are proved for the continuous flow. In this section, we show that the results can be extended to thediscrete update rules.

As for the discrete case, two additional assumptions are needed as follows (For brevity, we put the complete assumptionto the appendix):

Assumption 2. I (smooth). For any fixed x, Φ(·;x) is M smooth;

II (Learning Rate). For t > t0, ηt ≤ C(t), where C(t) is a non-decreasing function. Also, ηt is lower boundedby a positive real, that is, there exists a constant η > 0, such that, for any t > t0, ηt ≥ η.

We make the following explanations for Assumption 2. Assumption 2(I) is needed technically because we need toconsider second order Taylor expansion around each point along the training w(t); t = 1, · · · , . Assumption 2(II)guarantees that the second order Taylor expansion is upper bounded and the step size is not too small. With Assumption2, we can get the main results in section 4.5 for discrete update rules of AdaGrad and RMSProp. The proof of discretecase bears great similarity with the continuous case, and we defer them to Appendix D.

6.3 Definability can guarantee convergence of direction

Note that Theorems 4, 5, and 6 do not guarantee the uniqueness of convergent direction. To solve this problem, weintroduce one additional assumption as follows:

Assumption 3. For any i, qi is definable in some o-minimal structure including the exponential function;

We defer the formal definition of definable to Appendix C, but point out here that Assumption 3(I) allows for linear,ReLU, polynomial activations, max pooling and convolutional layers, and skip connections. Furthermore, for locallyLipchitz definable function, chain rule holds almost everywhere (Lemma 13). With Assumption 1 and Assumption 1,we can prove direction of parameters does converge:

Theorem 7. Let v satisfy an adaptive gradient flow F which satisfies Assumption 3. Then v(t) converges.

The proof can be derived by bounding the curve length of v(t) using γ(t) and Kurdyka-Lojasiewicz inequality, andwe defer the details to Appendix C.

7 Conclusion

In this paper, we study the convergent direction of AdaGrad and RMSProp on homogeneous deep neural networks withthe logistic loss. We prove RMSProp will converge to the KKT points of the L2 max margin problem, while AdaGraddoes not. The main technical contribution of this paper is to propose a general framework for analyses of adaptiveoptimization algorithms’ convergent direction. In future, the following topics regarding implicit bias of optimizationalgorithms are deserved to investigate.

• How does momentum influence the convergent direction of parameters? With momentum term, optimizationalgorithms may have an update with direction quite different from negative gradient. Therefore, whether theempirical loss is decreasing with respect to iterations for optimization algorithms with momentum is not beenstudied yet.

• How does regularization techniques such as weight decay influence the convergent direction of parameters?In practice, regularized losses are more popular, which may have an optimal point with finite norm. It will beinteresting to investigate implicit bias of optimizers in this circumstance.

References

[1] S. Arora, S. S. Du, W. Hu, Z. Li, R. R. Salakhutdinov, and R. Wang. On exact computation with an infinitelywide neural net. In Advances in Neural Information Processing Systems, pages 8141–8150, 2019.

9

Page 10: arxiv.org · 2020. 12. 14. · arXiv:2012.06244v1 [cs.LG] 11 Dec 2020 THE IMPLICIT BIAS FOR ADAPTIVE OPTIMIZATION ALGORITHMS ON HOMOGENEOUS NEURAL NETWORKS A PREPRINT Bohan Wang Microsoft

A PREPRINT

[2] P. Bartlett and J. Shawe-Taylor. Generalization performance of support vector machines and other pattern classi-fiers. Advances in Kernel methods—support vector learning, pages 43–54, 1999.

[3] A. Brutzkus, A. Globerson, E. Malach, and S. Shalev-Shwartz. Sgd learns over-parameterized networks thatprovably generalize on linearly separable data. arXiv preprint arXiv:1710.10174, 2017.

[4] A. Choromanska, M. Henaff, M. Mathieu, G. B. Arous, and Y. LeCun. The loss surfaces of multilayer networks.In Artificial intelligence and statistics, pages 192–204, 2015.

[5] L. Deng, G. Hinton, and B. Kingsbury. New types of deep neural network learning for speech recognitionand related applications: An overview. In 2013 IEEE international conference on acoustics, speech and signalprocessing, pages 8599–8603. IEEE, 2013.

[6] J. Duchi, E. Hazan, and Y. Singer. Adaptive subgradient methods for online learning and stochastic optimization.Journal of machine learning research, 12(7), 2011.

[7] S. Gunasekar, J. Lee, D. Soudry, and N. Srebro. Characterizing implicit bias in terms of optimization geometry.arXiv preprint arXiv:1802.08246, 2018.

[8] S. Gunasekar, J. D. Lee, D. Soudry, and N. Srebro. Implicit bias of gradient descent on linear convolutionalnetworks. In Advances in Neural Information Processing Systems, pages 9461–9471, 2018.

[9] G. Hinton, N. Srivastava, and K. Swersky. Neural networks for machine learning lecture 6a overview of mini-batch gradient descent. Cited on, 14(8), 2012.

[10] Z. Ji and M. Telgarsky. Gradient descent aligns the layers of deep linear networks. arXiv preprintarXiv:1810.02032, 2018.

[11] Z. Ji and M. Telgarsky. The implicit bias of gradient descent on nonseparable data. In Conference on LearningTheory, pages 1772–1798, 2019.

[12] Z. Ji and M. Telgarsky. Directional convergence and alignment in deep learning. arXiv preprintarXiv:2006.06657, 2020.

[13] D. P. Kingma and J. Ba. Adam: A method for stochastic optimization. arXiv preprint arXiv:1412.6980, 2014.

[14] K. Kurdyka. On gradients of functions definable in o-minimal structures. In Annales de l’institut Fourier,volume 48, pages 769–783, 1998.

[15] Y. Li, E. X. Fang, H. Xu, and T. Zhao. Implicit bias of gradient descent based adversarial training on separabledata. In International Conference on Learning Representations, 2019.

[16] I. Loshchilov and F. Hutter. Decoupled weight decay regularization. arXiv preprint arXiv:1711.05101, 2017.

[17] L. Luo, Y. Xiong, Y. Liu, and X. Sun. Adaptive gradient methods with dynamic bound of learning rate. arXivpreprint arXiv:1902.09843, 2019.

[18] K. Lyu and J. Li. Gradient descent maximizes the margin of homogeneous neural networks. arXiv preprintarXiv:1906.05890, 2019.

[19] M. S. Nacson, S. Gunasekar, J. D. Lee, N. Srebro, and D. Soudry. Lexicographic and depth-sensitive margins inhomogeneous and non-homogeneous deep models. arXiv preprint arXiv:1905.07325, 2019.

[20] M. S. Nacson, J. Lee, S. Gunasekar, P. H. P. Savarese, N. Srebro, and D. Soudry. Convergence of gradientdescent on separable data. In The 22nd International Conference on Artificial Intelligence and Statistics, pages3420–3428. PMLR, 2019.

[21] M. S. Nacson, N. Srebro, and D. Soudry. Stochastic gradient descent on separable data: Exact convergencewith a fixed learning rate. In The 22nd International Conference on Artificial Intelligence and Statistics, pages3051–3059. PMLR, 2019.

[22] B. Neyshabur, R. R. Salakhutdinov, and N. Srebro. Path-sgd: Path-normalized optimization in deep neuralnetworks. In Advances in Neural Information Processing Systems, pages 2422–2430, 2015.

[23] Q. Qian and X. Qian. The implicit bias of adagrad on separable data. In Advances in Neural InformationProcessing Systems, pages 7761–7769, 2019.

[24] S. Ruder. An overview of gradient descent optimization algorithms. arXiv preprint arXiv:1609.04747, 2016.

[25] D. Soudry, E. Hoffer, M. S. Nacson, S. Gunasekar, and N. Srebro. The implicit bias of gradient descent onseparable data. The Journal of Machine Learning Research, 19(1):2822–2878, 2018.

[26] A. Voulodimos, N. Doulamis, A. Doulamis, and E. Protopapadakis. Deep learning for computer vision: A briefreview. Computational intelligence and neuroscience, 2018, 2018.

10

Page 11: arxiv.org · 2020. 12. 14. · arXiv:2012.06244v1 [cs.LG] 11 Dec 2020 THE IMPLICIT BIAS FOR ADAPTIVE OPTIMIZATION ALGORITHMS ON HOMOGENEOUS NEURAL NETWORKS A PREPRINT Bohan Wang Microsoft

A PREPRINT

[27] A. C. Wilson, R. Roelofs, M. Stern, N. Srebro, and B. Recht. The marginal value of adaptive gradient methodsin machine learning. In Advances in neural information processing systems, pages 4148–4158, 2017.

[28] I. H. Witten, E. Frank, M. A. Hall, and C. J. Pal. Data Mining: Practical machine learning tools and techniques.Morgan Kaufmann, 2016.

[29] T. Xu, Y. Zhou, K. Ji, and Y. Liang. When will gradient methods converge to max-margin classifier under relumodels? arXiv preprint arXiv:1806.04339, 2018.

[30] T. Young, D. Hazarika, S. Poria, and E. Cambria. Recent trends in deep learning based natural language process-ing. ieee Computational intelligenCe magazine, 13(3):55–75, 2018.

[31] P. Zhou, J. Feng, C. Ma, C. Xiong, S. HOI, et al. Towards theoretically understanding why sgd generalizes betterthan adam in deep learning. arXiv preprint arXiv:2010.05627, 2020.

[32] J. Zhuang, T. Tang, S. Tatikonda, N. Dvornek, Y. Ding, X. Papademetris, and J. S. Duncan. Adabelief optimizer:Adapting stepsizes by the belief in observed gradients. arXiv preprint arXiv:2010.07468, 2020.

A Preliminaries

A.1 Property of Exponential and Logistic Loss

In this subsection, we prove several properties which both exponential and logistic loss possess. The properties ofexponential and logistic loss can be described as the following Lemma.

Lemma 8. For ℓ ∈ ℓexp, ℓlog:

• There exists a C1 function f , such that ℓ = e−f ;

• For any x ∈ R, f ′ > 0. Therefore, f is reversible, and f−1 ∈ C1;

• f ′(x)x is non-decreasing for x ∈ (0,∞), and limx→∞ f ′(x)x = ∞;

• There exists a large enough xf and a constant K ≥ 1 ,such that,

– ∀θ ∈ [ 12 , 1), ∀x ∈ (xf ,∞), and ∀y ∈ f−1(xf ,∞): (f−1)′(x) ≤ K(f−1)′(θx) and f ′(y) ≤ Kf ′(θy);

– For all y ∈ [xf ,∞), f(x)f ′(x) ∈ [ 1

2K x, 2Kx];

– For all x ∈ [f−1(xf ),∞), f−1(x)(f−1)′(x) ∈ [ 1

2K x, 2Kx].

– f(x) = Θ(x) as x→ ∞.

All the properties are easy to verify in Lemma 8 and we omit it here. For brevity, we will use g(x) = f−1(x) in thefollowing proofs.

A.2 KKT Condition

Being a KKT point is a first order necessary condition for being an optimal point. We first give the definition ofapproximate KKT point for general optimization problem (Q).

Definition 3. Consider the following optimization problem (Q) for x ∈ Rd:

min f(x)

subject to gi(x) ≤ 0, ∀i ∈ [N ],

where f, gi (i = 1, · · · , N ): Rd → R are locally Lipschitz functions. We say that x ∈ Rd is a feasible point of (P ) ifx satisfies gi(x) ≤ 0 for all i ∈ [N ].

For any ε, δ > 0, a feasible point of (Q) is an (ε, δ)-KKT point if there exists λi ≥ 0, k ∈ ∂f(x), and hi ∈ ∂gi(x)for all i ∈ [N ] such that

1.

∥∥∥k +∑

i∈[N ] λihi(x)∥∥∥2≤ ε;

2. ∀i ∈ [N ] : λigi(x) ≥ −δ.

Specifically, when ε = δ = 0, we call x a KKT point of (Q).

11

Page 12: arxiv.org · 2020. 12. 14. · arXiv:2012.06244v1 [cs.LG] 11 Dec 2020 THE IMPLICIT BIAS FOR ADAPTIVE OPTIMIZATION ALGORITHMS ON HOMOGENEOUS NEURAL NETWORKS A PREPRINT Bohan Wang Microsoft

A PREPRINT

The following Mangasarian-Fromovitz constraint qualification (MFCQ) bridges (ε, δ) KKT points with KKT points.

Definition 4. A feasible point x of (Q) is said to satisfy MFCQ if there exists a ∈ Rd such that for every i ∈ [N ] withgi(x) = 0,

∀h ∈ ∂gi(x) : 〈h,a〉 > 0.

The MFCQ guarantees the limit of approximate KKT point with convergent ε and δ is a KKT point.

Lemma 9. Suppose for any k ∈ N xk is a (εk, δk)-KKT point of (Q) defined in Definition 3. If limk→∞ εk = 0,limk→∞ δk = 0, and limk→∞ xk = x, where the limit point x satisfies MFCQ, then x is a KKT point of (Q).

In our setting, the constraints gi is 1− qi, and by the homogeneity of qi, we can easily verify that every feasible pointof (P ) satisfies MFCQ.

A.3 How is the Continuous Form of Conditioner Formulated?

In this subsection, we explain the relationship between conditioners of the discrete case and the continuous case. Asfor AdaGrad,

h(t+ 1)−2 − h(t)−2 = ∂L(w(t))2,

which is apparently a discretization of

dh(t)−2

dt= ∂L(w(t))2.

As for RMSProp, the discrete update can be written as

h(t+ 1)−2 − h(t)−2 = (1− b)(∂L(w(t))2 − h(t)−2),

which is a discretization ofdh(t)−2

dt= (1− b)(∂L(w(t))2 − h(t)−2).

By solving the above differential equation, we have

de(1−b)th(t)−2

dt= e(1−b)t(1− b)∂L(w(t))2,

which implies by integration

h(t)−2 = ε+

∫ t

0

e−(1−b)(t−τ)(1− b)∂L(w(τ))2dτ,

which is exactly hR.

A.4 o-Minimal Structure and Definable functions

Here we define o-Minimal structure and definable functions which we omit in Assumption 3.

Definition 5 (Appendix B, [12]). An o-minimal structure is a collection S = Sn∞n=1, where each Sn is a set ofsubsets of Rn satisfying the following conditions: 1. S1 is the collection of all finite unions of open intervals andpoints;2. Sn includes the zero sets of all polynomials on Rn;3. Sn is closed under finite union, finite intersection, and complement;4. S is closed under Cartesian products: if A ∈ Sm and B ∈ Sn, then A×B ∈ Sm+n;5. S is closed under projection Πn onto the first n coordinates: if A ∈ Sn+1, then Πn(A) ∈ Sn.

A definable function on above o-Minimal Structure can be defined as follows:

Definition 6 (Appendix B, [12]). A function f : D → Rm with D ⊂ Rn is definable if the graph of f is in Sn+m.

A natural question is: which function is definable? The next Lemma helps to solve this question.

Lemma 10 (Lemma B.2, [12]).

• All polynomials are definable, therefore, linear or other polynomial activation is definable;

12

Page 13: arxiv.org · 2020. 12. 14. · arXiv:2012.06244v1 [cs.LG] 11 Dec 2020 THE IMPLICIT BIAS FOR ADAPTIVE OPTIMIZATION ALGORITHMS ON HOMOGENEOUS NEURAL NETWORKS A PREPRINT Bohan Wang Microsoft

A PREPRINT

• If both f(x) and g(x) are definable, min f(x), g(x) and max f(x), g(x) are definable, therefore, ReLUactivation is definable;

• If fi(x) : D → R (i = 1, 2, · · · , n) is definable, then f(x) = (f1(x), f2(x), · · · , fn(x)) is definable.

• Suppose there exists k, d0, d1, · · · , dL > 0, and L definable functions (g1, g2, · · · , gL), where gj :→ Rd0 ×· · · × Rdj−1 × Rk → Rdj . Let h1(x,W )

= g1(x,W ), and for 2 ≤ j ≤ L,

hj(x,W ) := gj (x, h1(x,W ), . . . , hj−1(x,W ),W ) ,

then all hj are definable. Therefore, neural networks with polynomial and ReLU activation, convolutionaland max-pooling layers, and skip connections are definable.

An important property for definable function is Kurdyka-Lojasiewicz inequality, which can bound gradient of definablefunction in a small region. Here we present two Kurdyka-Lojasiewicz inequalities given by [12]:

Lemma 11 (Lemma 3.6, [12]). Given a locally Lipschitz definable function f with an open domainD ∈ x|‖x‖ > 1,for any c, η > 0, there exists a > 0 and a definable desingularizing function Ψ on [0, a) (that is, Ψ(x) ∈ C1((0, a)) ∩C0([0, a)) with Ψ(0) = 0), such that,

Ψ′(f(x))‖x‖‖∂f(x)‖ ≥ 1, if f(x) ∈ (0, a), and∥∥∂⊥f(x)

∥∥ ≥ c‖x‖η∥∥∂\\f(x)

∥∥ .Lemma 12 (Lemma 3.7, [12]). Given a locally Lipschitz definable function f with an open domainD ⊂ x|‖x‖ > 1,for any λ > 0, there exists a > 0, and a definable desingularizing function Ψ on [0, a) such that

max

1,

2

λ

Ψ′(f(x))‖x‖1+λ‖∂f(x)‖ ≥ 1, if f(x) ∈ (0, a).

A.5 Locally Lipchitz Definable Functions Admit Chain Rule

Assumption 3 actually guarantees that Φ admits a chain rule, which is formally stated as following:

Lemma 13 (Lemma B.9, [12]). Given a locally Lipschitz definable f : D → R with an open domain D, for anyinterval I and any arc z : I → D, it holds for a.e. t ∈ I that

df (zt)

dt=

⟨z∗t ,

dztdt

⟩, for all z∗t ∈ ∂f (zt) .

B Proof of AdaGrad Flow and RMSProp Flow

B.1 Proof of Lemma 2 and 3

We first prove Lemma 2. The claim of Lemma 2 can be split into two parts: convergence of h1(t) and construction of

vA, βA, and LA.

Lemma 14. For AdaGrad flow defined as eq. (4) with h = hA,∫ ∞

0

(∂L(w(t)))2i dτ <∞, i = 1, · · · , p.

Proof. We use reduction of absurdity. If there exists an i, such that,∫∞0gi(τ)

2dτ diverges, by equivalence of integralconvergence, then ∫ ∞

0

(∂L(w(t)))2i

ε+∫ t

0(∂L(w(τ)))2i dτ

dt = ∞.

Since∫∞0

(∂L(w(τ)))2i dτ = ∞, when t is large enough,

ε+

∫ t

0

(∂L(w(τ)))2i dτ >

√ε+

∫ t

0

(∂L(w(τ)))2i dτ.

Therefore, ∫ ∞

0

(∂L(w(t)))2i√ε+

∫ t

0 (∂L(w(τ)))2i dτdt = ∞.

13

Page 14: arxiv.org · 2020. 12. 14. · arXiv:2012.06244v1 [cs.LG] 11 Dec 2020 THE IMPLICIT BIAS FOR ADAPTIVE OPTIMIZATION ALGORITHMS ON HOMOGENEOUS NEURAL NETWORKS A PREPRINT Bohan Wang Microsoft

A PREPRINT

However,

L(w(0)) − L(w(t)) =

∫ t

0

〈∂L(w(t)),hA(t)⊙ ∂L(w(t))〉dt

=

p∑

i=1

∫ t

0

(∂L(w(t)))2i√ε+

∫ t

0(∂L(w(τ)))2i dτ

dt

= ∞,

which leads to a contradictory, since L(ω(0))− L(ω(t)) is upper bounded by L(ω(0)).

The proof is completed.

Lemma 15. Define h∞ = limt→∞ hA(t). Let

vA(t) = h−1/2∞ ⊙w(t),

βA(t) = h−1∞ ⊙ hA(t),

LA(v) = L(h12∞ ⊙ v).

We have

dvA(t)

dt= −βA(t)⊙ ∂LA(vA(t)), (8)

limt→∞ βA(t) = 1, andd logβA(t)

dt is Lebesgue integrable.

Proof. We prove 8 by direct calculation.

dvA(t)

dt= h−1/2

∞ ⊙ dw(t)

dt

= −h−1/2∞ ⊙ hA(t)⊙ ∂L(w(t))

= −h−1/2∞ ⊙ hA(t)⊙ h−1/2

∞ ⊙ h1/2∞ ⊙ ∂wLA(h−1/2

∞ ⊙w(t))

= −βA(t)⊙ ∂h

−1/2∞ ⊙w(t)

LA(h−1/2∞ ⊙w(t))

= −βA(t)⊙ ∂v(t)LA(v(t)),

which completes the proof 8.

The last two claims can be derived directly by convergence of hA(t) and non-increasing monotonicity of βA(t).

As for RMSProp flow, the convergence of hR requires more effort. We start from the following lemma, which bound

Fi(t)=∫ t

0(1− b)bt−τ (∂L(w(t)))2i dτ .

Lemma 16. For RMSProp flow defined as eq. (4) with h = hR , Fi(t) is bounded (i = 1, 2, · · · , p), that is,

limt→∞Fi(t) <∞.

Proof. When b = 1, Fi(t) = 0 for all t and i, which trivially yields the claim. When b 6= 1, we use reduction of

absurdity. If there exists an i, such that, limt→∞Fi(t) = ∞, then tk= inft : Fi(t) ≥ k < ∞ holds. Furthermore,

since L is locally lipchitz with respect to w, gi(t) is locally bounded for any t, which leads to the absolute continuityof Fi. Therefore, since Fi(0) = 0, tk monotonously increases.

14

Page 15: arxiv.org · 2020. 12. 14. · arXiv:2012.06244v1 [cs.LG] 11 Dec 2020 THE IMPLICIT BIAS FOR ADAPTIVE OPTIMIZATION ALGORITHMS ON HOMOGENEOUS NEURAL NETWORKS A PREPRINT Bohan Wang Microsoft

A PREPRINT

Therefore, we have that∫ tk+1

tk

(∂L(w(t)))2i√ε+ Fi(t)

dt ≥∫ tk+1

tk

(∂L(w(t)))2i√ε+ k + 1

dt

=1√

ε+ k + 1

∫ tk+1

tk

(∂L(w(t)))2i dt

≥ 1√ε+ k + 1

∫ tk+1

tk

(1− b)e−(1−b)(tk+1−t)(∂L(w(t)))2i dt

=1− b√ε+ k + 1

(Fi(tk+1)− e−(1−b)(tk+1−tk)Fi(tk))

≥ 1− b√ε+ k + 1

(k − e−(1−b)(tk+1−tk)(k − 1))

≥ 1− b√ε+ k + 1

.

Adding all k ≥ K , we then have

L(tK) ≥∫ ∞

tK

(∂L(w(t)))2i√ε+ Fi(t)

dt = ∞,

which leads to a contradictory.

The proof is completed.

The next lemma shows that∫∞0 gi(t)

2dt converges, which indicates that limt→∞ Fi(t) = 0.

Lemma 17. For RMSProp flow defined as eq. (4) with h = hR ,∫∞0 gi(t)

2dt converges (i = 1, 2, · · · , p). Conse-

quently, limt→∞ Fi(t) = 0 and limt→∞ hR(t) = 1√ε1p.

Proof. Similar to Lemma 16, when b = 1, the claim trivially holds. When b 6= 1, by Lemma 16, there exist Mi > 0(i = 1, 2, · · · , p), such that, Fi(t) ≤Mi for any t > 0. Therefore,

L(0) ≥p∑

i=1

∫ ∞

0

(∂L(w(t)))2i√ε+ Fi(t)

dt ≥p∑

i=1

∫ ∞

0

(∂L(w(t)))2i√Mi + ε

dt,

which proves∫∞0 (∂L(w(t)))2i dt <∞.

Therefore, for any positive real ε > 0 and a fixed index i ∈ [p], there exists a time T , such that,∫ ∞

t=T

(∂L(w(t)))2i dt ≤ ε.

Thus, for any t ≥ T ,

Fi(t) ≤ e−(t−T )(1−b)Fi(T ) + ε,

which leads tolimt→∞Fi(t) ≤ ε.

Since ε and i can be picked arbitrarily, the proof is completed.

Similar to the proof of Lemma 15, we can rewrite the RMSProp flow as

dvR(t)

dt= −βR(t)⊙ ∂LR(vR(t)),

where

vR(t) = 4√εw(t),

βR(t) =√εh(t),

LR(v) = L( 4√ε−1v),

and limt→∞ βR(t) = 1p. We only need to proved logβR(t)

dt is Lebesgue integrable in the rest of this subsection.

15

Page 16: arxiv.org · 2020. 12. 14. · arXiv:2012.06244v1 [cs.LG] 11 Dec 2020 THE IMPLICIT BIAS FOR ADAPTIVE OPTIMIZATION ALGORITHMS ON HOMOGENEOUS NEURAL NETWORKS A PREPRINT Bohan Wang Microsoft

A PREPRINT

Lemma 18. Let (x)+ = xIx>0, and (x)− = xIx<0. Then, for any i = 1, 2, · · · , p, we have that

∫ ∞

0

d log

√ε+Fi(t)

ε

dt

+

dt <∞,

∫ ∞

0

d log

√ε+Fi(t)

ε

dt

dt > −∞.

Proof. For any fixed i = 1, 2, · · · , p, by Lemma 17,

∫ ∞

0

d log√

ε+Fi(t)ε

dtdt = 0.

Since

∫ ∞

0

d log√

ε+Fi(t)ε

dtdt =

∫ ∞

0

d log

√ε+Fi(t)

ε

dt

+

dt+

∫ ∞

0

d log

√ε+Fi(t)

ε

dt

dt,

we only need to prove the convergence of∫∞0

(d log

ε+Fi(t)

ε

dt

)

+

dt.

For any k ∈ Z+, denote Bk = (k, k + 1) ∩ t : d log

ε+Fi(t)

ε

dt > 0. Sinced log

ε+Fi(t)

ε

dt is measurable, we have

that Bk is a measurable set. Denote Mk as an upper bound ford log

ε+Fi(t)

ε

dt on [k, k + 1) (which is guaranteed

since L is locally Lipchitz ). Since Bk is a measurable set, there exists an open set Bk ⊂ (k, k + 1), such that

m(Bk/Bk) ≤ 1k2Mk

, where m is Lebesgue measure on R.

Therefore, we have

t∈[k,k+1)

d log

√ε+Fi(t)

ε

dt

+

dt =

t∈Bk

d log

√ε+Fi(t)

ε

dt

dt

≤∫

t∈Bk

d log

√ε+Fi(t)

ε

dt

dt+Mk ·

1

k2Mk

=

t∈Bk

d log

√ε+Fi(t)

ε

dt

dt+

1

k2.

16

Page 17: arxiv.org · 2020. 12. 14. · arXiv:2012.06244v1 [cs.LG] 11 Dec 2020 THE IMPLICIT BIAS FOR ADAPTIVE OPTIMIZATION ALGORITHMS ON HOMOGENEOUS NEURAL NETWORKS A PREPRINT Bohan Wang Microsoft

A PREPRINT

Furthermore, let Bk = ∪∞j=1(tk,j , t

′k,j). Then we have

t∈Bk

d log

√ε+Fi(t)

ε

dt

dt

=

∞∑

j=1

∫ t′k,j

tk,j

d log√

ε+lFi(t)ε

dtdt

=

∞∑

j=1

log

√ε+ Fi(t′k,j)

ε− log

√ε+ Fi(tk,j)

ε

=1

2

∞∑

j=1

logε+ Fi(t

′k,j)

ε+ Fi(tk,j)

=1

2

∞∑

j=1

logε+ e−(t′k,j−tk,j)(1−b)Fi(tk,j) +

∫ t′k,j

tk,j(1 − b)e−(t′k,j−t)(1−b)(∂L(w(t)))2i dt

ε+ Fi(tk,j)

≤1

2

∞∑

j=1

logε+ Fi(tk,j) +

∫ t′k,j

tk,j(1 − b)e−(t′k,j−t)(1−b)(∂L(w(t)))2i dt

ε+ Fi(tk,j)

≤1

2

∞∑

j=1

∫ t′k,j

tk,j(1− b)e−(t′k,j−t)(1−b)(∂L(w(t)))2i dt

ε+ Fi(tk,j)

≤1

2

∞∑

j=1

∫ t′k,j

tk,j(∂L(w(t)))2i dt

ε

≤1

2

∫t∈Bk

(∂L(w(t)))2i dt

ε<∞.

The proof is completed.

B.2 Proof of surrogate margin Lemmas: Lemma 1 and 2

In the beginning, we first prove that the normalized margin γ is upper bounded:

Lemma 19. For any v ∈ Rp/0, the normalized margin γ = qmin(v)‖v‖L and normalized gradients

∂qi(v)‖v‖L−1 (i =

1, 2, · · · , p) is upper bounded universally.

Proof. By homogeneity of qi (i = 1, 2, · · · , p), only parameters with unit norm needed to be considered. That is,

γ(v) : v ∈ Rp/0

= γ(v) : ‖v‖ = 1 .

Since qi is continuous and v : ‖v‖ = 1 is a compact set, normalized margin is upper bounded.

Normalized gradients∂qi(v)‖v‖L−1 are also bounded following similar routine since qi is locally Lipchitz.

We restate Lemma 1 as follows:

Lemma 20. Let a function v(t) satisfy an adaptive gradient flow F which satisfies Assumption 1, with loss L and

component learning rate β(t). Then we have limt→∞ρ(t)

‖v(t)‖ = 1. Furthermore, if further limt→∞ L(t) = ∞, we

have γ(t) = Θ(γ(t)) and limt→∞γ(t)

‖γ(t)‖ = 1

17

Page 18: arxiv.org · 2020. 12. 14. · arXiv:2012.06244v1 [cs.LG] 11 Dec 2020 THE IMPLICIT BIAS FOR ADAPTIVE OPTIMIZATION ALGORITHMS ON HOMOGENEOUS NEURAL NETWORKS A PREPRINT Bohan Wang Microsoft

A PREPRINT

Proof. By the definition of approximate norm ρ(t) = ‖β(t)− 12 ⊙ v(t)‖ and limt→∞ β(t) = 1p, we have that

1 = limt→∞ mini

β− 1

2

i (t) ≤ limt→∞ρ(t)

‖v(t)‖ ≤ limt→∞ρ(t)

‖v(t)‖ ≤ limt→∞ maxi

β− 1

2

i (t) = 1,

which leads to limt→∞ρ(t)

‖v(t)‖ = 1.

By the definition of L(v), we have

e−f(qmin(v)) ≤ L(v) =N∑

i=1

e−f(qi(v)) ≤ Ne−f(qmin(v)).

Rearranging the above equation, we have

qmin(v) ≥ g

(log

1

L(v)

)≥ g(f(qmin(v))− logN)

=g(f(qmin(v))) − logNg′(ξ) = qmin(v)− logNg′(ξ),

where ξ ∈ (f(qmin(v))− logN, f(qmin(v))).

Therefore, the surrogate margin can be bounded as

qmin(v(t)) − logNg′(ξ)

‖v(t)‖L‖v(t)‖Lρ(t)L

=qmin(v(t)) − logNg′(ξ)

ρ(t)L

≤γ(t) ≤ qmin

ρ(t)L= γ(t)

(‖v‖ρ(t)

)L

. (9)

By the assumption that limt→∞ L = 0, there exists a large enough time t, such that, any time t ≥ t, g(ξ) > 0. Theleft side of the above equation 9 can be further rearranged as

qmin(v(t)) − logNg′(ξ)

‖v(t)‖L‖v(t)‖Lρ(t)L

=

(γ(t)− logNg′(ξ)

‖v(t)‖L) ‖v(t)‖L

ρ(t)L

=

(γ(t)− logNg′(ξ)g(ξ)

‖v(t)‖Lg(ξ)

) ‖v(t)‖Lρ(t)L

≥(γ(t)− logNg′(ξ)γ(t)

g(ξ)

) ‖v(t)‖Lρ(t)L

=

(γ(t)− logNγ(t)

g(ξ)f ′(g(ξ))

) ‖v(t)‖Lρ(t)L

.

By the assumption that limt→∞ L(t) = 0, limt→∞ξ ≥ limt→∞ log 1L(t)

− logN = ∞, which further indicates

limt→∞ g(ξ)f ′(g(ξ)) = 0 by Lemma 8. The proof is completed by taking t to infinity of eq. (9).

By Lemmas 19 and 20, we have that if limt→∞ L(t) → 0, γ(t) is upper bounded. We then lower bound γ(t) byproving Lemma 2. As a warm-up, we first calculate the derivative of ρ2.

Lemma 21. The derivative of ρ2 is as follows:

1

2

dρ(t)2

dt= Lν(t) +

⟨v(t),β− 1

2 (t)⊙ dβ− 12

dt(t)⊙ v(t)

⟩,

where ν(t) is defined as∑N

n=1 e−f(qi(v(t)))f ′ (qi(v(t))) qi(v(t)). Furthermore, we have that ν(t) >

g(

log 1L(v(t))

)

g′

(

log 1L(v(t))

) L(v(t)).

18

Page 19: arxiv.org · 2020. 12. 14. · arXiv:2012.06244v1 [cs.LG] 11 Dec 2020 THE IMPLICIT BIAS FOR ADAPTIVE OPTIMIZATION ALGORITHMS ON HOMOGENEOUS NEURAL NETWORKS A PREPRINT Bohan Wang Microsoft

A PREPRINT

Proof. By taking derivative directly, we have that

1

2

dρ(t)2

dt=

⟨β− 1

2 (t)⊙ v(t),dβ(t)−

12 ⊙ v(t)

dt

=

⟨β(t)−

12 ⊙ v(t),v(t) ⊙ dβ(t)−

12

dt+ β(t)−

12 ⊙ dv(t)

dt

= −⟨v(t), ∂L(v(t))

⟩+

⟨β− 1

2 (t)⊙ v(t),v(t) ⊙ dβ(t)−12

dt

(∗)= L

N∑

i=1

e−f(qi(v(t)))f ′ (qi(v(t))) qi(v(t)) +

⟨β(t)−

12 ⊙ v(t),v(t) ⊙ dβ(t)−

12

dt

⟩,

where eq. (∗) comes from Homogeneity Assumption 1. I.

Furthermore, since qi(v(t)) ≥ qmin(v(t)) ≥ g(log 1L(v(t))

) for all i ∈ [N ] and f ′(x)x keeps increasing on (0,∞) by

Lemma 8,

f ′ (qi(v(t))) qi(v(t)) ≥ f ′(g

(log

1

L(v(t))

))· g(log

1

L(v(t))

)=

g(log 1

L(v(t))

)

g′(log 1

L(v(t))

) .

Therefore,

N∑

i=1

e−f(qi(v(t))(v(t)))f ′ (qi(v(t))) qi(v(t))

≥g(log 1

L(v(t))

)

g′(log 1

L(v(t))

)N∑

i=1

e−f(qi(v(t))) =g(log 1

L(v(t))

)

g′(log 1

L(v(t))

) L(v(t)).

The proof is completed.

With the estimation of dρ2

dt above, we come to the proof of Lemma 2.

Proof of Lemma 2. We first construct time t1 as follows: by properties of β(t) in Definition 1, there exists some largeenough time t1 > t0, such that for any t > t1,

p∑

i=1

∫ ∞

t1

d log

(β− 1

2

i (t))

dt

+

≤ min 1

2L,1

4,

p∑

i=1

∫ ∞

t1

d log

(β− 1

2i (t)

)

dt

≥ −min 1

2L,1

4,

and1

2≤ ‖β− 1

2 (t)‖∞ ≤ 3

2.

Taking logarithmic derivative to γ(v(t)), we have

d

dtlog γ(t)

=d

dt

(log

(g

(log

1

L(v(t))

))− L log ρ(t)

)

=g′(log 1

L(v(t))

)

g(log 1

L(v(t))

) · 1

L(v(t))·(−dL(v(t))

dt

)− L2 · ν(t)

ρ(t)2−L

⟨v(t),β− 1

2 (t)⊙ dβ−12

dt (t)⊙ v(t)

ρ(t)2.

19

Page 20: arxiv.org · 2020. 12. 14. · arXiv:2012.06244v1 [cs.LG] 11 Dec 2020 THE IMPLICIT BIAS FOR ADAPTIVE OPTIMIZATION ALGORITHMS ON HOMOGENEOUS NEURAL NETWORKS A PREPRINT Bohan Wang Microsoft

A PREPRINT

Let A =g′

(

log 1L(v(t))

)

g(

log 1L(v(t))

) · 1L(v(t))

·(−dL(v(t))

dt

)− L2 · ν(t)

ρ(t)2 and B =L

v(t),β−12 (t)⊙ dβ

−12

dt (t)⊙v(t)

ρ(t)2 . We then have

A ≥ 1

ν(t)·(−dL(v(t))

dt

)− L2 · ν(t)

ρ(t)2

≥ 1

ν(t)·(−dL(v(t))

dt− L2ν(t)2

ρ(t)2

)

=1

ν(t)

(⟨dv(t)

dt,β(t)−1 ⊙ dv(t)

dt

⟩−⟨β(t)−

12 ⊙ dv(t)

dt, β(t)−

12 ⊙ v(t)

⟩2)

=1

ν(t)

∥∥∥∥∥

(I −

(

β(t)−12 ⊙ v(t)

)(

β(t)−12 ⊙ v(t)

)⊤)β(t)−

12 ⊙ dv(t)

dt

∥∥∥∥∥

2

2

> 0.

As for B, we have that

B = −L

⟨v(t),β− 1

2 ⊙ dβ−12 (t)

dt (t)⊙ v(t)

ρ2

= −L∑p

i=1 v2i (t)β

− 12

i (t)dβ

−12

i (t)

dt∑pi=1 v

2i (t)β

−1i (t)

≥ −Lp∑

i=1

(d logβ

− 12

i (t)

dt

)

+

.

Combining the estimation of A and B, we then have

d

dtlog γ ≥ A+B ≥ −L

p∑

i=1

(d logβ

− 12

i (t)

)

+

, (10)

and integrating both sides leads to

log γ(t)− log γ(t1) ≥ −1

2.

The proof is completed.

By the proof of Lemma 3, we can then prove convergence of surrogate margin γ.

Proof of 3. By eq. (10), log γ(t) +∫ t

t1L∑p

i=1

(d log β

−12

i (τ)

)

+

dτ is non-decreasing. Furthermore, since

g(log 1L(v(t))

) ≤ qmin(v(t)), we have that

limt→∞γ(t) ≤ limt→∞qmin(v(t))

ρ(t)L= limt→∞

qmin(v(t))

‖v(t)‖L ,

which by Lemma 19, the last term is bounded.

Therefore, log γ(t) +∫ t

t1L∑p

i=1

(d logβ

−12

i (τ)

dt

)

+

dτ is bounded, which further indicates that it converges due to its

monotony. The proof is completed by the convergence of∫ t

t1L∑p

i=1

(d logβ

−12

i (τ)

)

+

dτ .

20

Page 21: arxiv.org · 2020. 12. 14. · arXiv:2012.06244v1 [cs.LG] 11 Dec 2020 THE IMPLICIT BIAS FOR ADAPTIVE OPTIMIZATION ALGORITHMS ON HOMOGENEOUS NEURAL NETWORKS A PREPRINT Bohan Wang Microsoft

A PREPRINT

B.3 convergent Property of L and ρ: Proof of Lemma 4

Here we restate the complete Lemma 4.

Lemma 22. Let v obey an adaptive gradient flow which satisfies Assumption 1. Let t1 be constructed as Lemma 2.Then, for any t ≥ t1, define

G(t) =

∫ 1L(t)

1L(t1)

g′ (log x)2

g (log x)2−2/L

· dx.

Then, for any t ≥ t1, the following inequality holds:

G(t) −G(t1) ≥∫ 1

L(t)

1L(t1)

g′ (log x)2

g (log x)2−2/L· dx ≥ (t− t1)e

− 1LL2γ (t1)

2/L.

Consequently, limt→∞ L = 0 and limt→∞ ρ = ∞.

Proof.

−dL(v(t))dt

=

∥∥∥∥β− 12 (t)⊙ dv(t)

dt

∥∥∥∥2

2

≥ 〈β(t)− 12 ⊙ dv(t)

dt, β− 1

2 (t)⊙ v(t)〉2 = L2 · ν(t)2

ρ2.

Furthermore, we have that

L2 ν(t)2

ρ(t)2≥ L2 ·

g(log 1

L(v(t))

)

g′(log 1

L(v(t))

) L(v(t))

2

·

γ(t)

g(log 1

L(v(t))

)

2/L

≥ L2 ·

g(log 1

L(v(t))

)

g′(log 1

L(v(t))

) L(v(t))

2

·

e−1/2γ(t1)

g(log 1

L(v(t))

)

2/L

.

By simple calculation, we have that

g′(log 1

L(v(t))

)2

g(log 1

L(v(t))

)2−2/L· d

dt

1

L(v(t)) ≥ e−1LL2γ (t1)

2/L.

Taking integration to both sides, we have∫ 1

L(t)

1L(t1)

g′ (log x)2

g (log x)2−2/L· dx ≥ (t− t1)e

− 1LL2γ (t1)

2/L.

Since limt→∞(t− t1)e− 1

LL2γ (t1)2/L

= ∞, we have that limt→∞1

L(t)= ∞.

The proof is completed.

B.4 Verification of KKT Condition

In Lemma 5, we omit the construction of coefficients λi to highlight the key factors of coefficients (O(1 −〈v(t), −∂L(t)〉),O

(log 1

L(v(t))

)). We restate Lemma 5 and provide the detailed construction as follows:

Lemma 23. Let v obey adaptive gradient flow F with empirical loss L satisfying Assumption 1. Let timet1 be constructed as Lemma 2. Then, define coefficients in Definition 3 as λi(t) = qmin(v(t))

1−2/L‖v(t)‖ ·e−f(qi(v(t)))f ′ (qi(v(t))) /‖∂L(v(t))‖2. Then, for any time t ≥ t1, v(t) = qmin(v(t))

− 1L v(t) is an (ε(t), δ(t))

KKT point of (P ), where ε(t), δ(t) are defined as follows:

ε(t) = 8e1

γ(t1)2/L(1− cos(θ))

δ(t) =2e1+

1LKN

LK

log2

(

2L+1e12 B1/γ(t1)

)

γ(t1)−2/L 1

log 1L(v(t))

,

21

Page 22: arxiv.org · 2020. 12. 14. · arXiv:2012.06244v1 [cs.LG] 11 Dec 2020 THE IMPLICIT BIAS FOR ADAPTIVE OPTIMIZATION ALGORITHMS ON HOMOGENEOUS NEURAL NETWORKS A PREPRINT Bohan Wang Microsoft

A PREPRINT

where cos(θ(t)) is defined as inner product of v(t) and − ∂L(v(t)).

Proof. We verify the definition of approximate KKT point directly.

∥∥∥∥∥v(t)−N∑

i=1

λi(t)∂qi (v(t))

∥∥∥∥∥

2

2

=

∥∥∥∥qmin(v(t))−1/L‖v(t)‖v(t)− qmin(v(t))

−1/L‖v(t)‖e−f(qi(v(t)))f ′ (qi(v(t))) ∂qi (v(t))

‖∂L(v(t))‖2

∥∥∥∥2

2

= qmin(v(t))−2/L‖v(t)‖2

∥∥∥∥∥v(t) +∂L(v(t))

‖∂L(v(t))‖2

∥∥∥∥∥

2

2

= qmin(v(t))−2/L‖v(t)‖2

(2− 2

⟨v(t), −∂L(v(t))

⟩)

≤ 2g

(log

(1

L(v(t))

))−2/L

‖v(t)‖2(1− cos(θ(t)))

≤ 8g

(log

(1

L(v(t))

))−2/L

ρ(t)2(1− cos(θ(t)))

= 81

γ(t)2/L(1− cos(θ(t))) ≤ 8

1

γ(t)2/L(1− cos(θ(t)))

≤ 8e1

γ(t)2/L(1 − cos(θ(t))).

As for δ, we have that

N∑

i=1

λi (qi(v(t))− 1)

=

∑Ni=1 q

1−2/Lmin (v(t))‖v(t)‖ · e−f(qi(v(t))f ′ (qi(v(t))

(qi(v(t))

qmin(v(t))− 1)

‖∂L(v(t)‖2

=

∑Ni=1 qmin(v(t))

−2/L‖v(t)‖ · e−f(qi(v(t)))f ′ (qi(v(t))) (qi(v(t))− qmin(v(t)))

‖∂L(v(t))‖2(∗)≤

N∑

i=1

2K

Lqmin(v(t))

−2/L‖v(t)‖2ef(qmin(v(t)))−f(qi(v(t)))f ′ (qi(v(t)))

· (qi(v(t))− qmin(v(t)))1

log 1L(v(t))

,

where inequality (∗) is because

‖∂L(v(t))‖ ≥⟨∂L(v(t)), v(t)

⟩=

Lν(t)

‖v(t)‖

≥ L

‖v(t)‖g(log 1

L(v(t))

)

g′(log 1

L(v(t))

) L(v(t)) ≥ L

‖v(t)‖1

2Klog

1

L(v(t))· L(v(t))

≥ L

‖v(t)‖1

2Ke−f(qmin(v(t))) log

1

L(v(t)).

Bounding ‖v‖ using ρ and applying the definition of surrogate margin γ, we further have

22

Page 23: arxiv.org · 2020. 12. 14. · arXiv:2012.06244v1 [cs.LG] 11 Dec 2020 THE IMPLICIT BIAS FOR ADAPTIVE OPTIMIZATION ALGORITHMS ON HOMOGENEOUS NEURAL NETWORKS A PREPRINT Bohan Wang Microsoft

A PREPRINT

N∑

i=1

λi (qi(v(t))− 1)

≤N∑

i=1

8K

Lq−2/Lmin (v(t))ρ(t)2ef(qmin(v(t)))−f(qi(v(t)))f ′ (qi(v(t)))

(qi(v(t)) − qmin(v(t))1

log 1L(v(t))

≤N∑

i=1

8K

Lγ(t)−2/L · ef(qmin(v(t)))−f(qi(v(t)))f ′ (qi(v(t)))

(qi(v(t)) − qmin(v(t)))1

log 1L(v(t))

≤N∑

i=1

8e1LK

Lγ(t1)

−2/L · ef(qmin(v(t)))−f(qi(v(t)))f ′ (qi(v(t)))

(qi(v(t)) − qmin(v(t)))1

log 1L(v(t))

.

Since f(qmin(v(t))(v(t))) − f(qi(v(t))) = (qmin(v(t)) − qi(v(t)))f′(ξi(t)) (ξi(t) ∈ [qmin(v(t)), qi(v(t))] is guar-

anteed by Mean Value Theorem), we then have

N∑

i=1

λi (qi(v(t))− 1)

≤N∑

i=1

2e1LK

Lγ(t1)

−2/L · ef(qmin(v(t)))−f(qi(v(t)))f ′ (qi(v(t)))

(qi(v(t))− qmin(v(t)))1

log 1L(v(t))

=

N∑

i=1

2e1LK

Lγ(t1)

−2/L · e(qmin(v(t))−qi(v(t)))f′(ξi(t))f ′ (qi(v(t)))

(qi(v(t))− qmin(v(t)))1

log 1L(v(t))

≤N∑

i=1

2e1LK

Lγ(t1)

−2/L · e(qmin(v(t))−qi(v(t)))f′(ξi(t))K⌈log2(qi(v(t))/ξi(t))⌉f ′ (ξi(t))

(qi(v(t))− qmin(v(t)))1

log 1L(v(t))

(∗)≤

N∑

i=1

2e1LK

Lγ(t1)

−2/L · e(qmin(v(t))−qi(v(t)))f′(ξi(t))K

log2

(

2e12 B1/γ(t1)

)

f ′ (ξi(t))

(qi(v(t))− qmin(v(t)))1

log 1L(v(t))

(∗∗)≤ 2e1+

1LKN

LK

log2

(

2L+1e12 B1/γ(t1)

)

γ(t1)−2/L 1

log 1L(v(t))

,

23

Page 24: arxiv.org · 2020. 12. 14. · arXiv:2012.06244v1 [cs.LG] 11 Dec 2020 THE IMPLICIT BIAS FOR ADAPTIVE OPTIMIZATION ALGORITHMS ON HOMOGENEOUS NEURAL NETWORKS A PREPRINT Bohan Wang Microsoft

A PREPRINT

where eq. (∗) is because

⌈log2 (qi(v(t))/ξi(t))⌉ ≤ log2 (qi(v(t))/qmin(v(t))) + 1

≤ log2(qi(v(t))‖v(t)‖L/‖v(t)‖Lqmin(v(t))

)+ 1

≤ log2(2L+1B1ρ(t)

L/qmin

)≤ log2

(2L+1B1/γ(t)

)

≤ log2

(2L+1e

12B1/γ(t1)

),

and eq. (∗∗) is because e−xx ≤ e.

The proof is completed.

By Lemma 4, we have proved that limt→∞ L(t) = 0, which leads to limt→∞ δ(t) = 0. As stated in the main text, weonly need to bound ε(t), or equivalently (1 − cos(θ)).

Before moving forward, we introduce an equivalent proposition of that (1− cos(θ)) goes to zero.

Lemma 24. If there exists a time sequence ti∞i=1, such that, limi→∞ ti = ∞ and limi→∞

β− 12 ⊙ v,

β12 ⊙ ∂L

⟩=

−1, then

limi→∞

cos(θ(ti)) = 1.

Proof.

− cos(θ(t))

=

β(t)−12 ⊙ v(t) +

(v(t)−

β(t)−12 ⊙ v(t)

), β(t)

12 ⊙ ∂L(t) +

(∂L(t)−

β(t)12 ⊙ ∂L(t)

)⟩

=

β(t)−12 ⊙ v(t),

β(t)12 ⊙ ∂L(t)

⟩+

β(t)−12 ⊙ v(t),

(∂L(t)−

β(t)12 ⊙ ∂L(t)

)⟩

+

⟨(v(t)−

β(t)−12 ⊙ v(t)

), β(t)

12 ⊙ ∂L(t)

+

⟨(v(t)−

β(t)−12 ⊙ v(t)

),

(∂L(t)−

β(t)12 ⊙ ∂L(t)

)⟩.

Furthermore,

limi→∞

∥∥∥∥∂L(t)−

β(t)12 ⊙ ∂L(t)

∥∥∥∥ = 2− limi→∞

⟨∂L(t),

β(t)12 ⊙ ∂L(t)

⟩= 0.

Following the same routine, we have limi→∞

∥∥∥∥v(t)−

β(t)−12 ⊙ v(t)

∥∥∥∥ = 0.

The proof is completed.

Let cos(θ(t)) = −⟨

β(t)−

12 ⊙ v(t),

β(t)12 ⊙ ∂L(t)

⟩=

β(t)−12 ⊙ v(t),

β(t)−

12 ⊙ dv(t)

dt

⟩. Then we only need to

bound 1− cos(θ(t)).

24

Page 25: arxiv.org · 2020. 12. 14. · arXiv:2012.06244v1 [cs.LG] 11 Dec 2020 THE IMPLICIT BIAS FOR ADAPTIVE OPTIMIZATION ALGORITHMS ON HOMOGENEOUS NEURAL NETWORKS A PREPRINT Bohan Wang Microsoft

A PREPRINT

In Section 5.3, we briefly state the methodology of proving the convergence of ε(t). We will make it more clear here:In the proof Lemma 2, we show that sum of the derivative of g(log 1

L(t)) and 1

ρ(t)L with β(t) fixed can be bounded as

d

dt

(log

(g

(log

1

L(t)

)))− L

(∂v log ρ(t)

) dv(t)dt

=g′(log 1

L(t)

)

g(log 1

L(t)

) · 1

L(t)·(−dL(t)

dt

)− L2 · ν(t)

ρ(t)2

≥ 1

ν(t)

(⟨dv(t)

dt,β(t)−1 ⊙ dv(t)

dt

⟩−⟨β(t)−

12 ⊙ dv(t)

dt, β(t)−

12 ⊙ v(t)

⟩2)

=

⟨β(t)−

12 ⊙ dv(t)

dt , β(t)−

12 ⊙ v(t)

⟩2

ν(t)(cos(θ(t))−2 − 1)

=L2ν(t)

ρ(t)2(cos(θ(t))−2 − 1)

=L

⟨∂v log ρ(t),

dv

dt

⟩(cos(θ(t))−2 − 1). (11)

Eq. (11) indicates that, intuitively, (cos(θ(t))−2 − 1) can be bounded by the division of change of γ to change ofparameter part v in ρ. For this purpose, we define ρ in Section 5.3 to describe the accumulated change of v in ρ. Thefollowing lemma describe the basic property of ρ and its relationship with ρ.

Lemma 25. (1). The derivative of ρ2 is as follows:

1

2

dρ(t)2

dt= 〈∂L(v(t)),v(t)〉 = Lν(t);

(2). ρ(t) satisfies that, for any t2 > t1 > t1,

ρ(t2) ≥ e−12 ρ(t1);

(3). For any t > t1,

√1− e

12

2 ≤ ρ(t)ρ(t) ≤

√1 + e

12

2 .

Proof. (1). can be directly verified similar to Lemma 21. As for (2)., sinced log ρ(t)

dt = 12

dρ(t)2

dt

ρ(t)2 , we have that

d log ρ(t)

dt=

L∑N

i=1 e−f(qi(v(t)))f ′ (qi(v(t))) qi(v(t)) +

⟨β(t)−

12 ⊙ v(t),v(t) ⊙ dβ(t)−

12

dt

ρ(t)2

⟨β(t)−

12 ⊙ v(t),v(t) ⊙ dβ(t)−

12

dt

ρ(t)2

≥∑p

i=1 βi(t)− 1

2 (t)dβ

−12

i (t)

dt vi(t)2

∑pi=1 β

−1i (t)vi(t)2

=

∑pi=1 β

12

i (t)dβ

−12

i (t)

dt β−1i (t)vi(t)

2

∑pi=1 β

−1i (t)vi(t)2

=

∑pi=1

d logβ−

12

i (t)

dt β−1i (t)vi(t)

2

∑pi=1 β

−1i (t)vi(t)2

≥p∑

i=1

(d logβ

− 12

i (t)

dt

)

−.

25

Page 26: arxiv.org · 2020. 12. 14. · arXiv:2012.06244v1 [cs.LG] 11 Dec 2020 THE IMPLICIT BIAS FOR ADAPTIVE OPTIMIZATION ALGORITHMS ON HOMOGENEOUS NEURAL NETWORKS A PREPRINT Bohan Wang Microsoft

A PREPRINT

The proof for (2). is completed by integration.

As for (3)., by expanding 〈v(t),β− 12 (t)⊙ dβ−

12 (t)

dt ⊙ v(t)〉, we have that

∫ t

t1

〈v(τ),β− 12 (τ)⊙ dβ− 1

2 (τ)

dτ⊙ v(τ)〉dτ

=

p∑

i=1

∫ t

t1

v2i (τ)β

−1i (τ)β

12i (τ)

dβ− 1

2

i (τ)

dtdτ

≤p∑

i=1

∫ t

t1

β−1i (τ)v2

i (τ)

12

i (τ)dβ

− 12

i (τ)

dt

)

+

≤∫ t

t1

(p∑

i=1

β−1i v2

i (τ)

)(p∑

i=1

12i (τ)

dβ− 1

2i (τ)

dt

)

+

)dτ

(∗)=‖β(τ0)−

12 ⊙ v(τ0)‖2

∫ t

t1

(p∑

i=1

12i (τ)

dβ− 1

2

i (τ)

dt

)

+

)dτ

≤1

4‖β(τ0)−

12 ⊙ v(τ0)‖2

≤e12

4‖β(t)− 1

2 ⊙ v(t)‖2,

where τ0 in eq. (∗) is in [t1, t] and First Mean Value Theorem guarantees its existence.

The second inequality follows by lower bound∫ t

t1〈v(τ),β− 1

2 (τ) ⊙ dβ−12 (τ)

dτ ⊙ v(τ)〉dτ similarly.

The proof is completed.

With Lemma 25, we integrate the analysis of the change of γ into the following Lemma.

Lemma 26. For any time t3 > t2 ≥ t1,

∫ t3

t2

(cos(θ(τ))−2 − 1

)· d

dτlog ρ(τ)dτ ≤ 1

L

(1− e

12

2

) logγ (t3)

γ (t2)

+L(

1− e12

2

)p∑

i=1

∫ t3

t2

p∑

i=1

(d logβ

− 12

i (t)

dt

)

+

dt.

Proof. Recall that

d log ρ(t)

dt=Lν(t)

ρ(t)2=

⟨v(t),β(t)−1 ⊙ dv(t)

dt

ρ(t)2,

and

d

dtlog γ(t) ≥ 1

ν(t)

(⟨dv(t)

dt,β(t)−1 ⊙ dv(t)

dt

⟩−⟨β(t)−

12 ⊙ dv(t)

dt, β(t)−

12 ⊙ v(t)

⟩)

−Lp∑

i=1

(d logβi(t)

− 12 (t)

dt

)

+

.

26

Page 27: arxiv.org · 2020. 12. 14. · arXiv:2012.06244v1 [cs.LG] 11 Dec 2020 THE IMPLICIT BIAS FOR ADAPTIVE OPTIMIZATION ALGORITHMS ON HOMOGENEOUS NEURAL NETWORKS A PREPRINT Bohan Wang Microsoft

A PREPRINT

We then have

d

dtlog γ(t) ≥ 1

ν(t)

(⟨dv(t)

dt,β(t)−1 ⊙ dv(t)

dt

⟩−⟨β(t)−

12 ⊙ dv(t)

dt, β(t)−

12 ⊙ v(t)

⟩2)

−Lp∑

i=1

(d logβ(t)

− 12

i (t)

dt

)

+

=Lρ(t)2

L2ν(t)2d log ρ(t)

dt

(⟨dv(t)

dt,β(t)−1 ⊙ dv(t)

dt

⟩−⟨β(t)−

12 ⊙ dv(t)

dt, β(t)−

12 ⊙ v(t)

⟩2)

−Lp∑

i=1

(d logβ(t)

− 12

i (t)

dt

)

+

≥L(1− e

12

2

)ρ(t)2

L2ν(t)2d log ρ(t)

dt

(⟨dv(t)

dt,β(t)−1 ⊙ dv(t)

dt

−⟨β(t)−

12 ⊙ dv(t)

dt, β− 1

2 ⊙ v(t)

⟩2)

−Lp∑

i=1

(d logβ

− 12

i (t)

dt

)

+

=L

(1− e

12

2

)d log ρ

dt(cos(θ(t))−2 − 1)− L

p∑

i=1

(d logβ

− 12

i (t)

dt

)

+

.

The proof is completed.

Applying the First Mean Value Theorem together with Lemmas 24 and 26, one can easily obtain Lemma 6. To make

the following proof simpler, we restate Lemma 6 as the following corollary, while using cos(θ(t)) instead of cos(θ(t)).

Corollary 1. For any time t3 > t2 ≥ t1, there exists a time ξ ∈ (t2, t3), such that,

(cos(θ(ξ))−2 − 1

)(log ρ(t2)− log ρ(t1)) ≤

1

L

(1− e

12

2

) logγ (t3)

γ (t2)

+L(

1− e12

2

)p∑

i=1

∫ t3

t2

p∑

i=1

(d logβ

− 12

i (t)

dt

)

+

dt.

There is still a gap that ξ is guaranteed by First Mean Value Theorem and do not apparently converge to limit points.For this purpose, we introduce the following lemma 7, which helps to bound the change of v(t) by ρ(t).

Proof of Lemma 7. By direct calculation, we have that

∥∥∥∥dv(t)

dt

∥∥∥∥ =1

‖v(t)‖

∥∥∥∥(I − v(t)v(t)⊤

) dv(t)dt

∥∥∥∥ ≤ 1

‖v(t)‖

∥∥∥∥dv(t)

dt

∥∥∥∥ ≤ 2

‖v(t)‖

∥∥∥∥β(t)−1 ⊙ dv(t)

dt

∥∥∥∥ .

27

Page 28: arxiv.org · 2020. 12. 14. · arXiv:2012.06244v1 [cs.LG] 11 Dec 2020 THE IMPLICIT BIAS FOR ADAPTIVE OPTIMIZATION ALGORITHMS ON HOMOGENEOUS NEURAL NETWORKS A PREPRINT Bohan Wang Microsoft

A PREPRINT

Furthermore,∥∥∥∥β(t)−1 ⊙ dv(t)

dt

∥∥∥∥ =∥∥∥∂L(t)

∥∥∥

≤∑

i∈[N ]

e−f(qi(v(t)))f ′ (qi(v(t)))∥∥∂qi(v(t))

∥∥

≤∑

i∈[N ]

e−f(qi(v(t)))f ′ (qi(v(t))) qi(v(t))1

qi(v(t))B1‖v(t)‖L−1

≤∑

i∈[N ]

2L−1e−f(qi(v(t)))f ′ (qi(v(t))) qi(v(t))1

qi(v(t))B1ρ(t)

L−1

≤∑

i∈[N ]

2L−1e−f(qi(v(t)))f ′ (qi(v(t))) qi(v(t))1

g(log 1

L(t)

)B1ρ(t)L−1

= 2L−1 ν(t)

ρ(t)

ρ(t)L

g(log 1

L(t)

)B1 = 2L−1 ν(t)

ρ(t)

1

γ(t)B1

Therefore,∥∥∥∥dv(t)

dt

∥∥∥∥ ≤ 2Lν(t)

ρ(t)‖v(t)‖1

γ(t)B1 ≤ 3 · 2L−1 ν(t)

ρ(t)21

γ(t)B1 ≤ 3 · 2L−1

(1− e12

2 )

ν(t)

ρ(t)21

γ(t)B1

=3 · 2L−1

(1− e12

2 )

B1

L

d log ρ(t)

dt

1

γ(t)≤ 3 · 2L−1

(1− e12

2 )

e12B1

L

d log ρ(t)

dt

1

γ(t1).

The proof is completed.

Applying Corollary 1 and Lemma 7, we can then prove Theorem 4.

Proof of Theorem 4. Let v be any limit point of series v(t)∞t=0. We construct a series of approximate KKT pointwhich converges to v by induction.

Let t1 = t1. Now suppose tk−1 has been constructed. By Lemma 3 and that v is a limit point, there exists sk > tk−1

such that, for any t > sk

‖v (sk)− v‖2 ≤ 1

k,2

Llog

γ(t)

γ(sk)≤ 1

2k3, and, L

p∑

i=1

∫ ∞

sm

(d logβi(t)

− 12

dt

)

+

dt ≤ 1

2k3.

Let s′k > sk satisfy log ρ (s′k) = log ρ (sk) +1k (which is guaranteed as limt→∞ ρ = ∞ and Lemma 25). Therefore,

by Corollary 1, there exists tk ∈ (sk, s′k), such that

cos(θ(tk))−2 − 1 ≤ 1

k2(1− e12

2 ). (12)

Furthermore,

∥∥v(tk)− v

∥∥2≤∥∥v(tk)− v (sk)

∥∥2+ ‖v (sk)− v‖2 ≤ 3 · 2L−1

(1 − e12

2 )

e12

L

1

γ(t1)

1

k+

1

k→ 0.

Therefore, limt→∞ v(tk) = v. Furthermore, by eq. (12) and Lemma 23, we have that v(tk)/q1L

min(v(tk)) is an (εi, δi)

KKT point with limi→∞ εi = limi→∞ δi = 0. Since

v(tk)/q1L

min(v(tk)) = v(tk)/γ(tk),

28

Page 29: arxiv.org · 2020. 12. 14. · arXiv:2012.06244v1 [cs.LG] 11 Dec 2020 THE IMPLICIT BIAS FOR ADAPTIVE OPTIMIZATION ALGORITHMS ON HOMOGENEOUS NEURAL NETWORKS A PREPRINT Bohan Wang Microsoft

A PREPRINT

and γ(tk) converges to a positive number, we further have

limk→∞

v(tk)/q1L

min(v(tk)) = v/ lim

k→∞γ(tk)

1L ,

is a KKT point of (P ), and along the same direction of v.

The proof is completed.

B.5 convergent Direction of AdaGrad and RMSProp: Theorems 5 and 6

First of all, we prove Theorem 5 by substitute v in Theorem 4 with h∞ ⊙w.

Proof of Theorem 5. Let w be any limit point of series w(t)∞t=0. Since v(t) = h− 1

2∞ ⊙ w(t),

h− 1

2∞ ⊙ w is a limit

point of v(t). By Theorem 4,

h− 1

2∞ ⊙ w/qmin

(

h− 1

2∞ ⊙ w

)1/L

= h− 1

2∞ ⊙ w/qmin(w)1/L is a KKT point of (P ).

Therefore, there exist non-negative reals λiNi=1, such that

h− 1

2∞ ⊙ w/qmin(w)1/L =

N∑

i=1

λi∂qi(h

− 12∞ ⊙ w/qmin(w)1/L),

N∑

i=1

λi(qi(h− 1

2∞ ⊙ w/qmin(w)1/L)− 1) = 0.

Applying the relationship between qi and qi, we then have

h−1∞ ⊙ w/qmin(w)1/L =

N∑

i=1

λi∂qi(w/qmin(w)1/L),

N∑

i=1

λi(qi(w/qmin(w)1/L)− 1) = 0.

The proof is completed.

Theorem 6 can be obtained in the same way.

Proof of Theorem 6. The claim holds since vR is just w with a positive scaling factor, vR and w share the samedirection.

The proof is completed.

B.6 Convergence Rate of Empirical Loss and Parameter Norm

In this section, we will give a tight bound for the convergence rate of empirical loss and parameter norm, which isderived by estimating G(x) in Lemma 22.

Theorem 8. Let v obey an adaptive gradient flow F with empirical loss L satisfying Assumption 1. Let G(x) bedefined as Lemma 22, i.e.,

G(x) =

∫ x

1L(t1)

g′ (log z)2

g (log z)2−2/L

· dz.

Then G(x) = Θ(x(log x)2L−2), and G−1(x) = Θ(x(log x)2−

2L ). Consequently,

L(t) = Θ

(1

t log t2−2L

), and ‖v(t)‖ = Θ

(1

log t1L

).

29

Page 30: arxiv.org · 2020. 12. 14. · arXiv:2012.06244v1 [cs.LG] 11 Dec 2020 THE IMPLICIT BIAS FOR ADAPTIVE OPTIMIZATION ALGORITHMS ON HOMOGENEOUS NEURAL NETWORKS A PREPRINT Bohan Wang Microsoft

A PREPRINT

Proof. For any large enough x,

G(x) =

∫ x

1L(t1)

g′ (log z)2

g (log z)2−2/L

· dz

(∗)= Θ

(∫ x

1L(t1)

g (log z)2/L

(log z)2· dz)

= Θ

(∫ x

1L(t1)

1

(log z)2−2L

· dz)

= Θ(x(log x)

2L−2

),

where eq. (∗) is due to Lemma 8.

Since G(x) is monotonously increasing, and limx→∞G(x) = ∞, we have x = Θ(G−1(x)(logG−1(x))

2L−2

),

which further leads to G−1(x) = Θ(x(log x)2−2L ).

Since 1L(t)

= G−1(Ω(t)), we have 1L(t)

= Ω(t(log t)2−2L ), which further leads to L(t) = O

(1

t(log t)2−2L

).

On the other hand,

−dL(v(t))dt

=∥∥∥β 1

2 (t)⊙ ∂L(t)∥∥∥2

2≤ 2

∥∥∥∂L(t)∥∥∥2

2= O

(L(v(t))g

(log

1

L(t)

)1− 1L

),

which leads to 1L(t)

= G−1(O(t)), and L(t) = Ω

(1

t(log t)2−2L

). Therefore, L(t) = Θ

(1

t(log t)2−2L

)

By Lemma 19, ‖v(t)‖L = Θ(log 1

L(t)

), which leads to ‖v(t)‖ = Θ

(1

log t1L

).

The proof is completed.

The convergent behavior of ∂L(v(t)) can be derived immediately by the above Theorem.

Corollary 2. Let v obey an adaptive gradient flow F with empirical loss L satisfying Assumption 1. Then,

‖∂L(v(t))‖ = Θ

(1

t(log t)1−1L

).

Proof. Since

Ω(L(v(t))‖v(t)‖L−1

)= 〈∂L(v(t)), v(t)〉 ≤ ‖∂L(v(t))‖ = O

(L(v(t))‖v(t)‖L−1

),

the proof is completed.

C Proof of Theorem 7

In this section, we will prove that direction of parameters converges, that is, limt→∞v(t)

‖v(t)‖ exists. Concretely, define

the length swept byv(t)

‖v(t)‖ as ζ(t), i.e.,

ζ(t) =

∫ t

0

∥∥∥∥dv(τ)

∥∥∥∥ dτ.

We will upper bound ζ(t) in rest of this section.

To begin with, we define another surrogate margin γ as

γ(v) =log 1

L(v)

‖v‖L + e−L(v).

30

Page 31: arxiv.org · 2020. 12. 14. · arXiv:2012.06244v1 [cs.LG] 11 Dec 2020 THE IMPLICIT BIAS FOR ADAPTIVE OPTIMIZATION ALGORITHMS ON HOMOGENEOUS NEURAL NETWORKS A PREPRINT Bohan Wang Microsoft

A PREPRINT

The following Lemma then lower bound the derivative of γ.

Lemma 27. For large enough t, we have

dγ(v(t))

dt≥ 1

2

(∥∥∂\\γ (v(t))∥∥∥∥∥∂\\L (v(t))

∥∥∥+∥∥∂⊥γ (v(t))

∥∥∥∥∥∂⊥L (v(t))

∥∥∥),

and

dζ(t)

dt=

∥∥∥(β(t)⊙ ∂L (v(t)))⊥∥∥∥

‖v(t)‖ .

Proof. To begin with, we calculate the rate of β converging to 1p. For AdaGrad, given a fixed index i ∈ [N ], we havethat,

0 ≤ βi(t)− 1 =(h∞)−1i (hi(t)− (h∞)i)

=Θ(hi(t)− (h∞)i)

(∫ ∞

t

∂L(w(τ))2dτ

)

(∫ ∞

t

1

τ2(log τ)2−2L

)

=O(

1

t(log t)2−2L

). (13)

Similarly, for RMSProp, given a fixed index i ∈ [N ],

0 ≤ 1−√ε√

ε+ (1− b)∫ t

0 e−(1−b)(t−τ)(∂L(τ))2i dτ

= Θ

((1 − b)

∫ t

0

e−(1−b)(t−τ)(∂L(τ))2i dτ)

((1 − b)

∫ t−√t

0

e−(1−b)(t−τ)(∂L(τ))2i dτ)

((1 − b)

∫ t

t−√t

e−(1−b)(t−τ)(∂L(τ))2i dτ)

=O((1 − b)e−(1−b)

√t

∫ ∞

0

(∂L(τ))2i dτ)+O

((1− b)

√t

(t−√t)2

)

=O(

1

t32

)= O

(1

t(log t)2−2L

).

We then directly calculate the derivative of γ:

dγ(v(t))

dt=〈∂γ(v(t)), dv(t)

dt〉

=

⟨− ∂LL‖v(t)‖L

− Llog 1

L‖v(t)‖L+1

v(t)− e−L∂L, dv(t)dt

=

⟨∂L

L‖v(t)‖L+ L

log 1L

‖v(t)‖L+1v(t) + e−L∂L,β(t)⊙ ∂L

=

⟨∂L

L‖v(t)‖L+ L

log 1L

‖v(t)‖L+1v(t), ∂L

⟩+⟨e−L∂L,β(t)⊙ ∂L

+

⟨∂ ˜L(v(t))L‖v(t)‖L

+ Llog 1

L‖v(t)‖L+1

v(t), (1− β(t))⊙ ∂L⟩.

31

Page 32: arxiv.org · 2020. 12. 14. · arXiv:2012.06244v1 [cs.LG] 11 Dec 2020 THE IMPLICIT BIAS FOR ADAPTIVE OPTIMIZATION ALGORITHMS ON HOMOGENEOUS NEURAL NETWORKS A PREPRINT Bohan Wang Microsoft

A PREPRINT

The final term can be further bounded as follows:

∣∣∣∣∣

⟨∂L

L‖v(t)‖L+ L

log 1L

‖v(t)‖L+1v, (1− β(t))⊙ ∂L

⟩∣∣∣∣∣

≤‖1− β(t)‖∞

∥∥∥∥∥∂L

L‖v(t)‖L+ L

log 1L

‖v(t)‖L+1v

∥∥∥∥∥ ‖∂L‖

≤O(

1

t(log t)2−2L

)O(

1

‖v(t)‖

)‖∂L‖

=O(L(v(t))

)O(

1

‖v(t)‖

)‖∂L‖

=O(

1

‖v(t)‖L)‖∂L‖2.

On the other hand,

⟨e−L(v(t))∂L(v(t)),β(t) ⊙ ∂L(v(t))

⟩= Θ(‖∂L(v(t))‖2).

Therefore, there exists a large enough T , such that, any t ≥ T ,

⟨e−L∂L(v(t)),β(t)⊙ ∂L(v(t))

⟩+

⟨∂L

L‖v(t)‖L+ L

log 1L

‖v(t)‖L+1v, (1− β(t))⊙ ∂L(v(t))

≥1

2

⟨e−L(v(t))∂L(v(t)), ∂L(v(t))

⟩,

which further leads to

dγ(v(t))

dt≥ 1

2

⟨e−L(v(t))∂L(v(t)), ∂L(v(t))

⟩+

⟨∂L(v(t))

L(v(t))‖v(t)‖L+ L

log 1L(v(t))

‖v(t)‖L+1v, ∂L(v(t))

⟩.

We then calculate axial component and radial component of e−L(v)∂L(v), ∂L(v)

L(v)‖v‖L+L

log 1L(v)

‖v‖L+1 v, and ∂L(v) respec-

tively as follows:

(e−L(v)∂L(v)

)\\

= 〈e−L(v)∂L(v), v〉v,(e−L(v)∂L(v)

)⊥= e−L(v)(∂L(v)− 〈∂L(v), v〉v),

(∂L(v)

L(v)‖v‖L+ L

log 1L(v)

‖v‖L+1v

)

\\= L

log 1L (v)

‖v‖L+1v +

〈∂L(v), v〉L(v)‖v‖L

v,

(∂L(v)

L(v)‖v‖L+ L

log 1L(v)

‖v‖L+1v

)

⊥=∂L(v)− 〈∂L(v), v〉v

L‖v‖L,

(∂L(v)

)\\

= 〈∂L(v), v〉v(∂L(v)

)⊥= ∂L(v) − 〈∂L(v), v〉v.

32

Page 33: arxiv.org · 2020. 12. 14. · arXiv:2012.06244v1 [cs.LG] 11 Dec 2020 THE IMPLICIT BIAS FOR ADAPTIVE OPTIMIZATION ALGORITHMS ON HOMOGENEOUS NEURAL NETWORKS A PREPRINT Bohan Wang Microsoft

A PREPRINT

Therefore,

1

2

⟨e−L(v(t))∂L(v(t)), ∂L(v(t))

⟩+

⟨∂L(v(t))

L(v(t))‖v(t)‖L+ L

log 1L(v(t))

‖v‖L+1v(t), ∂L(v(t))

=1

2

⟨(e−L(v(t))∂L(v(t))

)\\

+(e−L(v(t))∂L(v(t))

)⊥,(∂L(v(t))

)\\

+(∂L(v(t))

)⊥

+

⟨(∂L(v(t))

L(v(t))‖v(t)‖L+ L

log 1L(v(t))

‖v(t)‖L+1v(t)

)

\\+

(∂L(v(t))

L(v(t))‖v(t)‖L+ L

log 1L(v(t))

‖v(t)‖L+1v(t)

)

⊥,

(∂L(v(t))

)\\

+(∂L(v(t))

)⊥

(∗)=

1

2

(∥∥∥∥(e−L(v(t))∂L(v(t))

)\\

∥∥∥∥∥∥∥∥(∂L(v(t))

)\\

∥∥∥∥+∥∥∥(e−L(v(t))∂L(v(t))

)⊥

∥∥∥∥∥∥(∂L(v(t))

)⊥

∥∥∥)

+

∥∥∥∥∥∥

(∂L(v(t))

L(v(t))‖v(t)‖L+ L

log 1L(v(t))

‖v(t)‖L+1v(t)

)

\\

∥∥∥∥∥∥

∥∥∥∥(∂L(v(t))

)\\

∥∥∥∥

+

∥∥∥∥∥

(∂L(v(t))

L(v(t))‖v(t)‖L+ L

log 1L(v(t))

‖v(t)‖L+1v(t)

)

∥∥∥∥∥∥∥∥(∂L(v(t))

)⊥

∥∥∥

≥1

2

∥∥∥∥∥∥

(e−L(v(t))∂L(v(t)) + ∂L(v(t))

L(v(t))‖v(t)‖L+ L

log 1L(v(t))

‖v(t)‖L+1v(t)

)

\\

∥∥∥∥∥∥

∥∥∥∥(∂L(v(t))

)\\

∥∥∥∥

+1

2

∥∥∥∥∥

(e−L(v(t))∂L(v(t)) + ∂L(v(t))

L(v(t))‖v(t)‖L+ L

log 1L(v(t))

‖v(t)‖L+1v(t)

)

∥∥∥∥∥∥∥∥(∂L(v(t))

)⊥

∥∥∥ ,

where eq. (∗) is because

〈e−L(v(t))∂L(v(t)), v(t)〉 = −e−L(v(t))〈−∂L(v(t)), v(t)〉 < 0,

e−L(v(t)) > 0,

Llog 1

L(v(t))

‖v(t)‖L+1+

〈∂L(v(t)), v(t)〉L(v(t))‖v(t)‖L

= − 1

L(v(t))‖v(t)‖L+1

(Lν(t)− L(v(t)) log 1

L(v(t))

)< 0,

1

L(v(t))‖v(t)‖L> 0,

〈∂L(v(t)), v(t)〉 = −〈−∂L(v(t)), v(t)〉 < 0.

The proof is completed since ∂γ(v) = −e−L(v)∂L(v)− ∂L(v)

L(v)‖v‖L− L

log 1L(v)

‖v‖L+1 v.

The following lemma gives an equivalent proposition of that the curve length ζ is finite.

Lemma 28. ζ is finite if and only if

∫ ∞

0

∥∥∥∂⊥L (v(t))∥∥∥

‖v(t)‖ dt <∞.

Proof. By definition,

(β(t)⊙ ∂L (v(t)))⊥

=β(t)⊙ ∂L (v(t)) − 〈v(t),β(t)⊙ ∂L (v(t))〉v(t)=∂⊥L (v(t)) + (1 − β(t))⊙ ∂L (v(t))− 〈v(t), (1 − β(t))⊙ ∂L (v(t))〉v(t).

33

Page 34: arxiv.org · 2020. 12. 14. · arXiv:2012.06244v1 [cs.LG] 11 Dec 2020 THE IMPLICIT BIAS FOR ADAPTIVE OPTIMIZATION ALGORITHMS ON HOMOGENEOUS NEURAL NETWORKS A PREPRINT Bohan Wang Microsoft

A PREPRINT

Furthermore, by eq. (13), ‖1− β(t)‖∞ = O( 1

t(log t)2−2L),

‖(1− β(t))⊙ ∂L (v(t)) − 〈v(t), (1 − β(t))⊙ ∂L (v(t))〉v(t)‖≤2‖(1− β(t))⊙ ∂L (v(t)) ‖

=O(

1

t2

).

Therefore,

∫ ∞

0

dζ(t)

dt≤∫ ∞

0

∥∥∥∂⊥L (v(t))∥∥∥

‖v(t)‖ dt+

∫ ∞

0

O(

1

t2

)dt.

The proof is completed.

We then prove

∫ ∞

0

∥∥∥∂⊥L (v(t))∥∥∥

‖v(t)‖ dt <∞.

Theorem 9. There exists a, γ0 > 0 and a definable desingularizing function Φ on [0, a), such that, for large enough t,∥∥∥∂⊥L (v(t))

∥∥∥‖v(t)‖ ≤ −dΨ(γ0 − γ(t))

dt.

Proof. Sincedγ(t)dt ≥ 0, and both

log 1˜L(v(t))

‖v(t)‖L and e−L(v(t)) are upper bounded, γ(t) converges to a limit non-

decreasingly. Define γ0 = limt→∞ γ(t). If γ(t) = γ0 for a finite time t0, thendγ(t)dt = 0 for any t ≥ t0, which

further leads to ‖(∂L(v)(t))⊥‖ = 0, and the proof is then completed by letting Ψ(x) = x. Therefore, we only con-sider the case where γ(t) < γ0 for any finite time t. For any large enough t, we further divide the proof into twocases.

Case I.∥∥∂⊥γ(v(t))

∥∥ ≥ ‖v(t)‖L4

∥∥∂\\γ(v(t))∥∥.

Applying Lemma 11 to γ0− γ(v)|‖v‖>1, there exists an a1 > 0 and a definable desingularizing function Ψ1, such that

if ‖v‖ > 1, γ(v) > γ0 − a1, and ∥∥∂⊥γ(v)∥∥ ≥ ‖v‖L

4

∥∥∂\\γ(v)∥∥ ,

thenΨ′

1(γ0 − γ(v))‖v‖‖∂γ(v(t))‖ ≥ 1.

Since limt→∞ γ(t) = γ0, and limt→∞ ‖v(t)‖ = ∞, there exists a large enough time T1, such that, for every t ≥ T1,‖v(t)‖ > 1, and γ(t) > γ0 − a1.

Therefore, for any t ≥ T1 which satisfies∥∥∂⊥γ(v(t))

∥∥ ≥ ‖v(t)‖L4

∥∥∂\\γ(v(t))∥∥, we have

∥∥∂⊥γ(v(t))∥∥ ≥

∥∥∂\\γ(v(t))∥∥ ,

which further indicates ∥∥∂⊥γ(v(t))∥∥ ≥ 1

2

∥∥∂γ(v(t))∥∥ .

Furthermore, by Lemma 27,

dγ(t)

dt≥1

2

∥∥∂⊥γ(v(t))∥∥∥∥∥∂⊥L(v(t))

∥∥∥

≥1

4‖v(t)‖

∥∥∂γ(v(t))∥∥

∥∥∥∂⊥L(v(t))∥∥∥

‖v(t)‖

≥ 1

4Ψ′1(γ0 − γ(t))

∥∥∥∂⊥L(v(t))∥∥∥

‖v(t)‖ .

34

Page 35: arxiv.org · 2020. 12. 14. · arXiv:2012.06244v1 [cs.LG] 11 Dec 2020 THE IMPLICIT BIAS FOR ADAPTIVE OPTIMIZATION ALGORITHMS ON HOMOGENEOUS NEURAL NETWORKS A PREPRINT Bohan Wang Microsoft

A PREPRINT

Case II.∥∥∂⊥γ(v(t))

∥∥ < ‖v(t)‖L4

∥∥∂\\γ(v(t))∥∥.

Applying Lemma 12 to γ0− γ(v)|‖v‖>1, we have that there exists an a2 > 0 and a desingularizing function on [0, a2),such that if v > 1, and γ(v) > γ0 − a2, then

max

1,

4

L

Ψ′

2(γ0 − γ(v))‖v‖L2 +1‖∂γ(v)‖ ≥ 1.

Similar to Case I., there exists a large enough time T2 and constants C1, such that, for every t ≥ T1, ‖v(t)‖ > 1,

γ(t) > γ0 − a2, C1‖v(t)‖L ≤ log 1L(v(t))

, and L(v(t)) ≤ ‖v(t)‖−2L.

Therefore, ∥∥∥∥∂\\ log1

L(v(t))

∥∥∥∥ =1

L(v(t))Lν(t)

‖v(t)‖ ≥L log 1

L(v(t))

‖v(t)‖ ≥ LC1‖v(t)‖L−1,

and

‖∂\\γ(v(t))‖ =

∥∥∥∥∥Llog 1

L(v(t))

‖v(t)‖L+1v(t) +

〈∂L(v(t)), v(t)〉L‖v(t)‖L

v(t) + e−L(v(t))〈∂L(v(t)), v(t)〉v(t)∥∥∥∥∥

≤− Llog 1

L(v(t))

‖v(t)‖L+1− 〈∂L(v(t)), v(t)〉

L(v(t))‖v(t)‖L− e−L(v(t))〈∂L(v(t)), v(t)〉

≤L(2 lnN + 1)

‖v(t)‖L+1+ ‖∂L(v(t))‖

≤L(2 lnN + 1)

‖v(t)‖L+1+B1L(v(t))‖v(t)‖L−1

≤B1 + L(2 lnN + 1)

‖v(t)‖ ,

which further leads to ∥∥∥∥∂\\ log1

L(v(t))

∥∥∥∥ ≥ LC1

B1 + L(2 lnN + 1)‖v(t)‖2L‖∂\\γ(v(t))‖. (14)

On the other hand, ∥∥∥∥∂⊥ log1

L(v(t))

∥∥∥∥ =1

L(v(t))‖∂L(v(t)) − 〈∂L(v(t)), v(t)〉v(t)‖,

while

‖∂⊥γ(v(t))‖

=

∥∥∥∥∥e−L(v(t))(∂L(v(t))− 〈∂L(v(t)), v(t)〉v(t)) + ∂L(v(t)) − 〈∂L(v(t)), v(t)〉v(t)

L(v(t))‖v(t)‖L

∥∥∥∥∥

=

(e−L(v(t)) +

1

L(v(t))‖v(t)‖L

)∥∥∥∂L(v(t))− 〈∂L(v(t)), v(t)〉v(t)∥∥∥

≥ 1

‖v(t)‖L∥∥∥∥∂⊥ log

1

L(v(t))

∥∥∥∥ . (15)

Since ∥∥∂⊥γ(v(t))∥∥ < ‖v‖L

4

∥∥∂\\γ(v(t))∥∥ ,

combining eqs. (14) and (15), we have that∥∥∥∥∂\\ log

1

L(v(t))

∥∥∥∥ ≥ LC1

B1 + L(2 lnN + 1)‖v(t)‖2L‖∂\\γ(v(t))‖

≥ LC1

B1 + L(2 lnN + 1)‖v(t)‖ 7

4L‖∂⊥γ(v(t))‖

≥ LC1

B1 + L(2 lnN + 1)‖v(t)‖ 3

4L

∥∥∥∥∂⊥ log1

L(v(t))

∥∥∥∥ .

35

Page 36: arxiv.org · 2020. 12. 14. · arXiv:2012.06244v1 [cs.LG] 11 Dec 2020 THE IMPLICIT BIAS FOR ADAPTIVE OPTIMIZATION ALGORITHMS ON HOMOGENEOUS NEURAL NETWORKS A PREPRINT Bohan Wang Microsoft

A PREPRINT

Since ∂L(v(t)) is parallel to ∂ log 1L(v(t))

, we also have

∥∥∥∂\\L∥∥∥ (v(t)) ≥ LC1

B1 + L(2 lnN + 1)‖v(t)‖ 3

4L∥∥∥∂⊥L(v(t))

∥∥∥ .

Furthermore, ∥∥∂γ(v(t))∥∥ =

∥∥∂⊥γ(v(t))∥∥ +

∥∥∂\\γ(v(t))∥∥ ≤ ‖v‖L

4

∥∥∂\\γ(v(t))∥∥ .

Thus, by Lemma 27,

dγ(t)

dt≥1

2

∥∥∂\\γ (v(t))∥∥∥∥∥∂\\L (v(t))

∥∥∥

≥ LC1

2(B1 + L(2 lnN + 1))‖v(t)‖ 1

2L∥∥∂γ (v(t))

∥∥∥∥∥∂⊥L(v(t))

∥∥∥

=LC1

2(B1 + L(2 lnN + 1))‖v(t)‖ 1

2L+1∥∥∂γ (v(t))

∥∥

∥∥∥∂⊥L (v(t))∥∥∥

‖v(t)‖

≥ LC1

2(B1 + L(2 lnN + 1))max1, 4

L

Ψ′

2(γ0 − γ(v(t)))

∥∥∥∂⊥L (v(t))∥∥∥

‖v(t)‖ .

Concluding Case I. and Case II., for any t ≥ maxT1, T2, and Ψ(x) =

max4Ψ1(x),2(B1+L(2 lnN+1))max1, 4

LLC1

Ψ2(x), we have that

Ψ′(γ0 − γ(v(t)))dγ(t)

dt≥

∥∥∥∂⊥L (v(t))∥∥∥

‖v(t)‖ .

The proof is completed.

D Proof for the Discrete Case

We prove the result for AdaGrad and experiential loss, with the result for RMSProp and logistic loss follows exactlyas the continuous case. We slightly change the order of four stages in the flow: First, in Section D.1, we prove thatthe conditioner has a limit with no zero entry; secondly, in Section D.2, we prove that the empirical loss convergesto zero; then, in Section D.3, we construct a further smoothed approximate margin, and prove it has a lower bound;finally, in Section D.4, we prove that every limit point of AdaGrad is along some KKT point of optimization problem(PA) defined in Theorem 5.

D.1 Convergence of conditioners

Before the proof, we give a formal definition of the learning rate bound C(t) in Assumption 2: let M be the smooth

constant in Assumption 2. I. Then, C(t) = maxminihAi (t)

−1/M, 1,C2

1

2LNe−1 , where C0 will be clear below. By

the monotony of hAi (t), apparently C(t) is non-decreasing. Now we can prove

∑∞t=1 ∂L(w(t))2 <∞.

Lemma 29. Suppose L is M smooth with respect to w. Then, for ∂L(w(t))∞t=1 updated by AdaGrad (eq. (3)),∑∞t=1 ∂L(w(t))2 <∞.

Proof. For any t > t0,

L(w(t)) − L(w(t + 1)) = ηt〈∂L(w(t)),hA(t)⊙ ∂L(w(t))〉 − M

2η2t ‖hA(t)⊙ ∂L(w(t))‖2

≥ ηt1

2〈∂L(w(t)),hA(t)⊙ ∂L(w(t))〉

≥ η1

2〈∂L(w(t)),hA(t)⊙ ∂L(w(t))〉.

Thus, since∑∞

t=1 at share the same convergent behavior with∑∞

t=1at

tτ=1 aτ

(at ≥ 0), by similar routine of Lemma

14, the proof is completed.

36

Page 37: arxiv.org · 2020. 12. 14. · arXiv:2012.06244v1 [cs.LG] 11 Dec 2020 THE IMPLICIT BIAS FOR ADAPTIVE OPTIMIZATION ALGORITHMS ON HOMOGENEOUS NEURAL NETWORKS A PREPRINT Bohan Wang Microsoft

A PREPRINT

Therefore, h∞= limt→∞ hA(t) has no zero entry. We can then define a discrete version of adaptive gradient flow as

v(t) = h−1/2∞ ⊙w(t),

β(t) = h−1∞ ⊙ h(t),

L(v) = L(h12∞ ⊙ v),

and

qi(v) = qi(h12∞ ⊙ v),

which further leads to

v(t+ 1)− v(t) = −β(t)⊙ ∂Lind(v(t)), (16)

and β(t) decreases component-wisely to 1p.

By Lemma 29, for any t ≥ t0, L(w(t)) ≤ L(w(t0)) < N . Therefore, there exists a positive real constant C0 only de-

pending on L(w(t0)), such that, ‖w(t)‖ ≥ C0. Furthermore, since h− 1

2∞ ≥ h(t0)− 1

2 , ‖v(t)‖ ≥ C0 maxihi(t0)− 1

2 .

Define C1 = C0 maxihi(t0)− 1

2 , which only depends on L(w(t0)) and ∂L(w(t)) (t ≤ t0).

Moreover, similar to approximate flow, there exists a time t1, such that, for any time t ≥ t1,

p∑

i=1

log1

βi(t)−12

≤ 1

2,

‖β−1(t)‖ ≥√

1

2.

D.2 Convergence of Empirical Loss

Define function γ′ as the rate of log 1L to ‖v‖L

4 :

γ′(t) =log 1

L(v(t))

‖v(t)‖L4

.

Then, we have the following lemma.

Lemma 30. For any t ≥ t1, γ′(t) is non-decreasing.

Proof. Since by Lemma 29, L is non-increasing, if ‖v(t + 1)‖ ≤ ‖v(t)‖, the proposition trivially holds. Therefore,we consider the case that ‖v(t+ 1)‖ > ‖v(t)‖ in the following proof.

The change of ‖v(t)‖ can be calculated as

‖v(t+ 1)‖2 − ‖v(t)‖2

=− ηt〈β(t)⊙ ∂L(v(t)),v(t)〉 + η2t ‖β(t)⊙ ∂L(v(t))‖2.

Let A = −ηt〈β(t)⊙ ∂L(v(t)),v(t)〉, and B = η2t ‖β(t)⊙ ∂L(v(t))‖2. We estimate them separately.

As for A:

A = −ηt〈β(t)⊙ ∂L(v(t)),v(t)〉≤ ηt‖β(t)⊙ ∂L(v(t))‖‖v(t)‖≤ 2ηt‖β

12 (t)⊙ ∂L(v(t))‖‖v(t)‖

≤ 2ηt‖β12 (t)⊙ ∂L(v(t))‖‖v(t)‖‖∂L(v(t))‖‖v(t)‖

〈∂L(v(t)),v(t)〉

≤ 2ηt‖β 1

2 (t)⊙ ∂L(v(t))‖2‖v(t)‖2〈∂L(v(t)),v(t)〉

.

37

Page 38: arxiv.org · 2020. 12. 14. · arXiv:2012.06244v1 [cs.LG] 11 Dec 2020 THE IMPLICIT BIAS FOR ADAPTIVE OPTIMIZATION ALGORITHMS ON HOMOGENEOUS NEURAL NETWORKS A PREPRINT Bohan Wang Microsoft

A PREPRINT

As for B:

B = η2t ‖β(t)⊙ ∂L(v(t))‖2

≤ 2η2t ‖β− 12 (t)⊙ ∂L(v(t))‖2

≤ LNe−1

Lν(t)2η2t ‖β− 1

2 (t)⊙ ∂L(v(t))‖2

≤ LNe−1

Lν(t)C21

2η2t ‖β− 12 (t)⊙ ∂L(v(t))‖2‖v(t)‖2

≤ 2ηt‖β 1

2 (t)⊙ ∂L(v(t))‖2‖v(t)‖2〈∂L(v(t)),v(t)〉

.

Therefore,

‖v(t+ 1)‖2 − ‖v(t)‖2 ≤ 4ηt‖β 1

2 (t)⊙ ∂L(v(t))‖2‖v(t)‖2〈∂L(v(t)),v(t)〉

. (17)

On the other hand, by Lemma 29,

L(v(t)) − L(v(t+ 1)) ≥ 1

2〈∂L(w(t)), ηth

A(t)⊙ ∂L(w(t))〉

= ηt1

2‖β 1

2 (t)⊙ ∂L(v(t))‖2. (18)

Combining eqs. (17) and (18), we have

41

ν(t)(L(v(t)) − L(v(t + 1))) ≥ L

2

‖v(t+ 1)‖2 − ‖v(t)‖2‖v(t)‖2 .

Since ν(t) ≥ L log 1L(t)

, we further have

1

L(v(t)) log 1L(v(t))

(L(v(t)) − L(v(t+ 1))) ≥ L

8

‖v(t+ 1)‖2 − ‖v(t)‖2‖v(t)‖2 .

By the convexity of log log 1x (when x is small) and − log x,

log log1

L(v(t+ 1))− log log

1

L(v(t)) ≥ L

4(log ‖v(t+ 1)‖ − log ‖v(t)‖).

The proof is completed.

Remark 1. Actually, the convexity does not hold for x ∈ [e−1, L(v(t1))] (if L(v(t1)) > e−1). However, we caninstead define

Φ0(x) = log log1

x+

∫ x

0

inf− 1

w log 1w

: w ∈ [x, log1

L(v(t1))]+ 1

x log 1x

dw.

Which satisfies for x ∈ [0, L(v(t1))], 1x log 1

x

≤ −Φ′0(x), Φ0(x) ≤ log log x, and Φ0(x) is convex. We can then replace

log x by e−Φ0(x) and prove the above theorem in exactly the same way.0

With the relationship between ‖v‖ and L, we can now prove that the empirical loss goes to zero.

Theorem 10. limt→∞ L(t) = 0. Furthermore, limt→∞ ρ(t) = ∞.

Proof. By Lemma 33, for any integer time t ≥ t0

L(t+ 1)− L(t) ≤ −1

2ηtL2ν(t)2

‖v(t)‖2 ≤ −1

2ηtL2L(t)2 log 1

L(t)2

‖v(t)‖2 .

38

Page 39: arxiv.org · 2020. 12. 14. · arXiv:2012.06244v1 [cs.LG] 11 Dec 2020 THE IMPLICIT BIAS FOR ADAPTIVE OPTIMIZATION ALGORITHMS ON HOMOGENEOUS NEURAL NETWORKS A PREPRINT Bohan Wang Microsoft

A PREPRINT

Furthermore, since γ′(t) ≥ γ′(t1), we have that

(log( 1

L(t))

ML1 γ(t0)

) 1L

≥ ρ.

Therefore,

L(t+ 1)− L(t) ≤ −1

2ηtL2L(t)2(log 1

L(t))2

‖v(t)‖2

≤ −1

2ηtL

2L(t)2(log

1

L(t)

)2(

γ′(t1)

log( 1L(t)

)

) 8L

= −1

2ηtL

2L(t)2(log

1

L(t)

)2− 8L

(γ′(t1))8L .

Let E0 = L (t1)2(log 1

L(t1)

)2−8/L

, then ψ(x) = minx2(log 1x )

2− 8L , E0. Apparently, ψ(x) is non-decreasing in

(0, L(t1)]. Therefore,E(x) =∫ L(t1)

xψ(s)ds is convex with respect to x, and

E(L(t+ 1))− E(L(t)) ≥ E′(L(t))(L(t+ 1)− L(t))

≥ 1

2ηtL

2 (γ′(t1))8L ,

which further implies

E(L(t))− E(L(t1)) ≥t−1∑

τ=t0

1

2ητL

2 (γ′(t1))2L .

Since limt→∞∑t−1

τ=t012ητL

2 (γ′(t1))8L = ∞, we then have

limt→∞

E(L(t)) = ∞,

and as a result,

limt→∞

L(t) = 0.

D.3 Convergence of surrogate margin

As a preparation, define B1 = max‖∂qi(v)‖ : i ∈ [N ],v ∈ B(1), and B2 = max ‖Hqi(v)‖ : i ∈ [N ],v ∈ B(1).

Furthermore, we define λ(x) = (log 1x )

−1, and µ(x) =log 1

L(v(t1))

log 1x

. Since limt→∞ L(t) = 0, we have the following

lemma.

Lemma 31. There exists a large enough time t2 ≥ t1, such that, for any t ≥ t2,

B21L(v) log7−

8L

1

L(v)/γ′(t1)

8L ≤ λ(L(v(t)))µ(L(v(t))),

and,1

γ′(t1)8−8/LL(v(t)) log8−8/L 1

L(v(t))(B2

1 + C−L1 B2

)≤ µ(t).

Proof. The proposition is obvious since logi( 1x) = o(x), ∀i as x→ 0.

39

Page 40: arxiv.org · 2020. 12. 14. · arXiv:2012.06244v1 [cs.LG] 11 Dec 2020 THE IMPLICIT BIAS FOR ADAPTIVE OPTIMIZATION ALGORITHMS ON HOMOGENEOUS NEURAL NETWORKS A PREPRINT Bohan Wang Microsoft

A PREPRINT

Then, we define a further surrogate margin γ of the discrete case as following:

γ(t) :=eΦ(L)

ρL,

where ρ(t) is defined as ‖β− 12 (t)⊙ v(t)‖, and Φ(x) is defined as

Φ(x) = log log1

x+

∫ x

0

(− sup

1 + 2(1 + λ(w)/L)µ(w)

w log 1w

: w ∈[w, L (t2)

]+

1

w log 1w

)dw.

The following properties hold for γ.

Lemma 32. • Let a series of vi satisfies limi→∞ ‖vi‖ = ∞. Then, limi→∞γ(vi)γ(vi)

= 1;

• If L(v(t)) ≤ L(v(t2)), then γ(t) < γ(t) ≤ γ(t).

Proof. As beginning, we verify the existence of Φ. Actually, when x is small enough,1+2(1+λ(x)/L)µ(x)

x log 1x

decreases,

and limx→01+2(1+λ(x)/L)µ(x)

x log 1x

= ∞. Therefore, there exists a small enough ε, such that, for any w < ε,

sup

1 + 2(1 + λ(w)/L)µ(w)

w log 1w

: w ∈[w, L (t2)

]=

1 + 2(1 + λ(w)/L)µ(w)

w log 1w

,

which further leads to

− sup

1 + 2(1 + λ(w)/L)µ(w)

w log 1w

: w ∈[w, L (t2)

]+

1

w log 1w

= −2(1 + λ(w)/L)µ(w)

w log 1w

,

which is integrable as w → 0. Concretely, for any x < ε,∫ x

0

(− sup

1 + 2(1 + λ(w)/L)µ(w)

w log 1w

: w ∈[w, L (t2)

]+

1

w log 1w

)dw

=

∫ x

0

−2(1 + λ(w)/L)µ(w)

w log 1w

dw

= − log1

L(t1)

(1

log 1x

+1

2L log2 1x

).

Therefore, for a series vi∞i=1 satisfying limi→∞ ‖vi‖ = ∞,

limi→∞

γ(vi)

γ(vi)= lim

i→∞e−O

(

1

log2 1L(vi)

)

= 1.

Furthermore, if x ≤ L(v(t2)),

Φ′(x) ≤ 1

w log 1w

− 1 + 2(1 + λ(w)/L)µ(w)

w log 1w

< 0,

which proves that γ(t) < γ(t).

The proof is completed.

To bound the norm of first and second derivatives of L, we further need the following lemma.

The next lemma characterizes the behavior of surrogate margin γ(t).

Lemma 33. For positive integer time t ≥ t2, γ(t) ≥ e−12 γ(t2).

40

Page 41: arxiv.org · 2020. 12. 14. · arXiv:2012.06244v1 [cs.LG] 11 Dec 2020 THE IMPLICIT BIAS FOR ADAPTIVE OPTIMIZATION ALGORITHMS ON HOMOGENEOUS NEURAL NETWORKS A PREPRINT Bohan Wang Microsoft

A PREPRINT

Proof. For any time t ≥ t2,

ρ(t+ 1)2 − ρ(t)2 = (‖β− 12 (t)⊙ v(t+ 1)‖2 − ρ(t)2) + (ρ(t+ 1)2 − ‖β− 1

2 (t)⊙ v(t+ 1)‖2).We calculate two parts separately

‖β− 12 (t)⊙ v(t+ 1)‖2 − ‖β−1

2 (t)⊙ v(t)‖2

=η2t ‖β12 (t)⊙ ∂L(v(t))‖2 + 2Lηtν(t)

≥0.

On the other hand, since β− 12 is non-decreasing,

ρ(t+ 1) = ‖β− 12 (t+ 1)⊙ v(t+ 1)‖ ≥ ‖β− 1

2 (t)⊙ v(t+ 1)‖.

‖β− 12 (t)⊙ v(t+ 1)‖2 − ‖β− 1

2 (t)⊙ v(t)‖2 can also be upper bounded as follows:

‖β− 12 (t)⊙ v(t+ 1)‖2 − ‖β− 1

2 (t)⊙ v(t)‖2

=η2t ‖β12 (t)⊙ ∂L(v(t))‖2 + 2Lηtν(t)

=2Lηtν(t)

(ηt‖β

12 (t)⊙ ∂L(v(t))‖2

2Lν(t)+ 1

)

(∗)≤2Lηtν(t)

(λ(L(v(t)))µ(L(v(t)))

L+ 1

),

where inequality (*) comes from the estimation of ‖β 12 (t) ⊙ ∂L(v(t))‖2 as follows: by the homogeneity of qi, we

have

‖∂L(v)‖ =

∥∥∥∥∥−N∑

i=1

e−qi(v)∂qi(v)

∥∥∥∥∥(∗∗)≤ B1L(v)‖v‖L−1

(∗∗∗)≤ B1

1

γ′(t1)4L

L(v) log4− 4L

1

L(v),

where inequality (∗∗) is due to ∂qi is (L− 1) homogeneous, and inequality (∗ ∗ ∗) holds by Lemma 30. On the other

hand, ν(t) ≥ L(v(t)) log 1L(v(t))

. Combining the estimation of ν and ‖β(t)− 12 ⊙ ∂L(β(t))‖, we have

ηt‖β12 (t)⊙ ∂L(v(t))‖2

2Lν(t)≤B2

1 L(v) log7−8L 1

L(v)

Lγ′(t1)8L

≤ λ(L(v(t)))µ(L(v(t)))L

.

Similar to Lemma 29, the decrease of L(v) can be calculated by second order Taylor Expansion:

L(t+ 1)− L(t) =− ηt

⟨β(t)⊙ ∂L(v(t)), ∂L(v(t))

+1

2η2t (β(t)⊙ ∂L)TH(L(v(ξ)))(β(t) ⊙ ∂L), (19)

where ξ ∈ (0, 1).

By homogeneity of qi, the norm of Hessian matrix ‖H(L(v(ξ)))‖ can be bounded as

‖H(L(v(ξ)))‖ =

∥∥∥∥∥N∑

i=1

e−qi(∂qi∂q

⊤i −Hqi

)∥∥∥∥∥2

(∗)≤

N∑

i=1

e−qi(B2

1‖v(ξ)‖2L−2 +B2‖v(ξ)‖L−2)

≤ L(v(ξ))‖v(ξ)‖2L−2(B2

1 + C−L1 B2

)

≤ 1

γ′(t1)8−8/LL(v(ξ)) log8−8/L 1

L(v(ξ))(B2

1 + C−L1 B2

).

41

Page 42: arxiv.org · 2020. 12. 14. · arXiv:2012.06244v1 [cs.LG] 11 Dec 2020 THE IMPLICIT BIAS FOR ADAPTIVE OPTIMIZATION ALGORITHMS ON HOMOGENEOUS NEURAL NETWORKS A PREPRINT Bohan Wang Microsoft

A PREPRINT

Therefore,

1

2η2t (β(t)⊙ ∂L)TH(L(v(ξ)))(β(t) ⊙ ∂L)

≤ηt‖β(t)12 ⊙ ∂L)‖2 1

γ′(t1)8−8/LL(v(ξ)) log8−8/L 1

L(v(ξ))(B2

1 + C−L1 B2

)

≤ηt‖β(t)12 ⊙ ∂L)‖2µ(t). (20)

Taking the estimation eq. (20) back to eq. (19), we have

L(t)− L(t+ 1) ≥ (1− µ(t))ηt‖β(t)12 ⊙ ∂L)‖2 (21)

By multiplying1+λ(L(t))µ(L(t))/L

(1−µ(L(t)))ν(t)to both sides of eq. (21), we then have

1 + λ(L(t))µ(L(t))/L(1− µ(L(t)))ν(t)

(L(t+ 1)− L(t))

≤− ηt

(1 +

λ(L(t))µ(L(t))L

)L2ν(t)

ρ(t)2

≤− L

2

‖β− 12 (t)⊙ v(t+ 1)‖2 − ‖β− 1

2 (t)⊙ v(t)‖2‖β− 1

2 (t)⊙ v(t)‖2.

Furthermore, since −Φ′(L(t)) ≥ 1+λ(L(t))µ(L(t))/L

(1−µ(L(t)))L(t)/λ(L(t)), by the convexity of Φ and − log x, we have that

L

(log

1

‖β− 12 (t)⊙ v(t+ 1)‖

− log1

‖β− 12 (t)⊙ v(t)‖

)

+(Φ(L(t+ 1))− Φ(L(t))) ≥ 0.

Therefore,

logΦ(L(t+ 1))

‖β− 12 (t)⊙ v(t+ 1)‖L

− logΦ(L(t)

‖β− 12 (t)⊙ v(t)‖L

≥ 0.

Furthermore, since

‖β− 12 (t+ 1)⊙ v(t+ 1)‖L

‖β− 12 (t+ 1)⊙ v(t)‖L

≥ Πpi=1

β−L/2i (t)

β−L/2i (t+ 1)

,

we have

γ(t+ 1) =e−Φ(t+1)

‖β− 12 (t+ 1)⊙ v(t+ 1)‖L

≥ e−Φ(t+1)

‖β− 12 (t)⊙ v(t+ 1)‖L

Πpi=1

β−L/2i (t)

β−L/2i (t+ 1)

≥ γ(t)Πpi=1

β−L/2i (t)

β−L/2i (t+ 1)

.

Thus, by induction,

γ(t+ 1) ≥ γ(t)Πpi=1

β−L/2i (t0)

β−L/2i (t+ 1)

≥ e−12 γ(t1).

The proof is completed.

Similar to the flow case, we can then prove the convergence of γ.

Lemma 34. There exists a positive real γ∞, such that

limt→∞

γ(t) = γ∞.

42

Page 43: arxiv.org · 2020. 12. 14. · arXiv:2012.06244v1 [cs.LG] 11 Dec 2020 THE IMPLICIT BIAS FOR ADAPTIVE OPTIMIZATION ALGORITHMS ON HOMOGENEOUS NEURAL NETWORKS A PREPRINT Bohan Wang Microsoft

A PREPRINT

Proof. Since for any t ≥ t2

logγ(t+ 1)

γ(t)≥ logΠp

i=1

β− 1

2

i (t)

β− 1

2i (t+ 1)

,

we have that γ(t)Πpi=1

1

β12i (t)

monotonously increases. Furthermore, since γ(t) < γ(t) is bounded, so does

γ(t)Πpi=1

1

β12i (t)

. Therefore, γ(t)Πpi=1

1

β12i (t)

converges to a positive real. Since limt→∞ Πpi=1

1

β12i (t)

= 1, the proof

is completed.

D.4 Verification of KKT point

Similar to the flow case, we have the following construction of (ε, δ) KKT point. The proof is exactly the same asLemma 5, and we omit it here.

Lemma 35. Let λi = q1−2/Lmin ‖v‖ · e−f(qi)f ′ (qi) /‖∂L‖2. Then v(t) is a (ε, δ) KKT point of (P ), where ε, δ are

defined as follows:

ε = C1(1 + cos(θ))

δ = C21

log 1L,

where cos(θ) is defined as 〈v(t), ∂L(t)〉, and C1, C2 are positive real constants.

By Lemma 24, we only need to prove that limt→∞ ˜cos(θ) = 1. Furthermore, the estimation of cos(θ) can be givenby the following lemma.

Lemma 36. For any t3 > t4 ≥ t2,

t4−1∑

τ=t3

(˜cos(θ)(τ)−2 − 1

)(log

1

ρ(t)− log

1

‖β− 12 (τ) ⊙ v(τ + 1)‖

)≤ 1

Llog

γ(t4)

γ(t3)+ log

(Πp

i=1

β− 1

2

i (t3)

β− 1

2

i (t4)

).

Proof. By Lemma 33, for any t ≥ t0

1

Llog

γ(t+ 1)

γ(t)=

1

L(Φ(t+ 1)− Φ(t)) +

(log

1

ρ(t+ 1)− log

1

ρ(t)

)

≥(log

1

ρ(t)− log

1

‖β− 12 (t)⊙ v(t+ 1)‖

)ρ(t)2

L2ν(t)2‖β 1

2 (t)⊙ ∂L(t)‖2

+

(log

1

‖β− 12 (t)⊙ v(t+ 1)‖

− log1

ρ(t)

)− log

(Πp

i=1

β− 1

2

i (t+ 1)

β− 1

2

i (t)

).

Furthermore, since Lν(t) =

β12 (t)⊙ ∂L(t),

β− 12 (t)⊙ v

⟩, we have

1

Llog

γ(t+ 1)

γ(t)≥(log

1

ρ(t)− log

1

‖β− 12 (t)⊙ v(t+ 1)‖

)cos(θ)−2

+

(log

1

‖β− 12 (t)⊙ v(t+ 1)‖

− log1

ρ(t)

)− log

(Πp

i=1

β− 1

2

i (t+ 1)

β− 1

2i (t)

).

The proof is completed.

We still need a lemma to bound the change of the direction of β− 12 ⊙ v.

43

Page 44: arxiv.org · 2020. 12. 14. · arXiv:2012.06244v1 [cs.LG] 11 Dec 2020 THE IMPLICIT BIAS FOR ADAPTIVE OPTIMIZATION ALGORITHMS ON HOMOGENEOUS NEURAL NETWORKS A PREPRINT Bohan Wang Microsoft

A PREPRINT

Lemma 37. For any t ≥ t2,

‖ β− 1

2 (t+ 1)⊙ v(t+ 1)− β− 1

2 (t)⊙ v(t)‖

≤O(1)

p∑

i=1

(β−1i (t+ 1)− β−1

i (t))+

(O(1)

‖β− 12 ⊙ v(t+ 1)‖ρ(t)

+ 1

)log

‖β− 12 (t)⊙ v(t+ 1)‖

ρ(t)

+

(O(1)

‖β− 12 ⊙ v(t+ 1)‖ρ(t)

+ 1

)logΠp

i=1

β− 1

2

i (t+ 1)

β− 1

2

i (t).

Proof. Since triangular inequality,

‖ β− 1

2 (t+ 1)⊙ v(t+ 1)− β− 1

2 (t)⊙ v(t)‖

≤∥∥∥∥

1

ρ(t+ 1)β− 1

2 (t+ 1)⊙ v(t+ 1)− 1

ρ(t+ 1)β− 1

2 (t+ 1)⊙ v(t)

∥∥∥∥

+

∥∥∥∥1

ρ(t+ 1)β− 1

2 (t+ 1)⊙ v(t)− 1

ρ(t+ 1)β− 1

2 (t)⊙ v(t)

∥∥∥∥

+

∥∥∥∥1

ρ(t+ 1)β− 1

2 (t)⊙ v(t)− 1

ρ(t)β− 1

2 (t)⊙ v(t)

∥∥∥∥ .

Let

A =

∥∥∥∥1

ρ(t+ 1)β− 1

2 (t+ 1)⊙ v(t+ 1)− 1

ρ(t+ 1)β− 1

2 (t+ 1)⊙ v(t)

∥∥∥∥ ;

B =

∥∥∥∥1

ρ(t+ 1)β− 1

2 (t+ 1)⊙ v(t)− 1

ρ(t+ 1)β− 1

2 (t)⊙ v(t)

∥∥∥∥ ;

C =

∥∥∥∥1

ρ(t+ 1)β− 1

2 (t)⊙ v(t)− 1

ρ(t)β− 1

2 (t)⊙ v(t)

∥∥∥∥ .

Then,

A =O(1)1

ρ(t+ 1)‖v(t+ 1)− v(t)‖

=O(1)1

ρ(t+ 1)‖ηtβ(t)⊙ ∂L‖ = O(1)

ηtρ(t+ 1)

‖∂L‖

(∗)≤O(1)

ηtν(t)

γ(t2)ρ(t+ 1)ρ(t)≤ O(1)

ρ(t+ 1)2 − ρ(t)2

ρ(t+ 1)ρ(t)

≤O(1)ρ(t+ 1)2 − ρ(t)2

ρ(t+ 1)2ρ(t+ 1)

ρ(t)≤ O(1)

ρ(t+ 1)

ρ(t)log

ρ(t+ 1)

ρ(t),

where eq. (∗) can be derived in the same way as Lemma 7;

B ≤p∑

i=1

(β− 1

2

i (t+ 1)− β− 1

2

i (t))2‖v(t)‖ρ(t+ 1)

≤ O(1)

p∑

i=1

β−1i (t+ 1)− β−1

i (t);

and

C = 1− ρ(t)

ρ(t+ 1).

44

Page 45: arxiv.org · 2020. 12. 14. · arXiv:2012.06244v1 [cs.LG] 11 Dec 2020 THE IMPLICIT BIAS FOR ADAPTIVE OPTIMIZATION ALGORITHMS ON HOMOGENEOUS NEURAL NETWORKS A PREPRINT Bohan Wang Microsoft

A PREPRINT

Therefore,

‖ β− 1

2 (t+ 1)⊙ v(t+ 1)− β− 1

2 (t)⊙ v(t)‖

≤O(1)

p∑

i=1

(β−1i (t+ 1)− β−1

i (t))+

(O(1)

ρ(t+ 1)

ρ(t)+ 1

)log

ρ(t+ 1)

ρ(t)

≤O(1)

p∑

i=1

(β−1i (t+ 1)− β−1

i (t))+

(O(1)

‖β− 12 ⊙ v(t+ 1)‖ρ(t)

+ 1

)log

‖β(t)− 12 (t)⊙ v(t+ 1)‖ρ(t)

+

(O(1)

‖β(t)− 12 ⊙ v(t+ 1)‖ρ(t)

+ 1

)logΠp

i=1

β− 1

2i (t+ 1)

β− 1

2

i (t)

We then prove that∑∞

τ=t2log ‖β−

12 (τ)⊙v(τ+1)‖

ρ(τ) = ∞.

Lemma 38. The sum of log ‖β−12 (τ)⊙v(τ+1)‖

ρ(τ) diverges, that is,∑∞

τ=t2log ‖β−

12 (τ)⊙v(τ+1)‖

ρ(τ) = ∞.

Proof.

∞∑

τ=t

log‖β− 1

2 (τ) ⊙ v(τ + 1)‖ρ(τ)

≥∞∑

τ=t

logρ(τ + 1)

ρ(τ)− logΠp

i=1

1

β− 1

2i (t)

.

The proof is completed since limt→∞ ρ(t) = ∞ and logΠpi=1

1

β−

12

i (t)is bounded.

Now we can prove the following lemma.

Lemma 39. Let v be any limit point of v(t)∞t=1. Then v is a KKT point of optimization problem (P ).

Proof. Let t1 be any integer time larger than t2. We construct a sequence ti∞i=1 by iteration. Suppose t1, · · · , tk−1

have been constructed. Let sk > tk−1 be a large enough time which satisfies

logΠpi=1

1

β− 1

2

i (sk)≤ 1

k3,

‖v(sk)− v‖ ≤ 1

k,

log

(γ∞γ(sk)

)≤ 1

k3

‖β−1(sk)‖ ≤ 1 +1

k − 1.

Then let (sk)′ (guaranteed by Lemma 38) be the first time greater than sk that∑(sk)′−1

τ=sklog ‖β−

12 (τ)⊙v(τ+1)‖

ρ(τ) ≥ 1k .

By Lemma 36, there exists a time tk ∈ [sk, (sk)′ − 1], such that ˜cos(θ)−2 − 1 ≤ 1

k2 .

45

Page 46: arxiv.org · 2020. 12. 14. · arXiv:2012.06244v1 [cs.LG] 11 Dec 2020 THE IMPLICIT BIAS FOR ADAPTIVE OPTIMIZATION ALGORITHMS ON HOMOGENEOUS NEURAL NETWORKS A PREPRINT Bohan Wang Microsoft

A PREPRINT

Moreover,

‖v(tk)− v‖2 ≤ 2

(‖ β− 1

2 (tk)⊙ v(tk)− v‖2 + ‖ β− 1

2 (tk)⊙ v(tk)− v(tk)‖2)

= 2‖ β−1

2 (tk)⊙ v(tk)− v‖2 + 2

(2− 2

β− 12 (tk)⊙ v(tk), v(tk)

⟩)

≤ 2‖ β(tk)⊙ v(tk)− v‖2 +O

(1

k

)

≤ 2‖ β(tk)⊙ v(tk)− β(sk)⊙ v(sk)‖2 +O

(1

k

)

≤ O(1)1

k+O(e

1k )

1

k→ 0.

The proof is completed.

Therefore, similar to the gradient flow case, we then have the following theorem.

Theorem 11. Let w be any limit point of w(t). Then w is along the direction of a KKT point of the followingoptimization problem.

Minimize1

2‖h− 1

2∞ ⊙w‖Subject to: qi(w) ≥ 1.

46