This paper was converted on www.awesomepapers.org from LaTeX by an anonymous user.
Want to know more? Visit the Converter page.

Symmetry-Preserving Paths in Integrated Gradients

Miguel Lerma Northwestern University, Evanston, USA [email protected]  and  Mirtha Lucas DePaul University, Chicago, USA [email protected]
Abstract.

We provide rigorous proofs that the Integrated Gradients (IG) attribution method for deep networks satisfies completeness and symmetry-preserving properties. We also study the uniqueness of IG as a path method preserving symmetry.

1. Introduction

1.1. Integrated Gradients

The Integrated Gradients (IG) attribution method for deep networks is introduced in [5]. A neural network can be interpreted as a function F:nF:\mathbb{R}^{n}\to\mathbb{R} that maps its inputs 𝐱=(x1,,xn)\mathbf{x}=(x_{1},\dots,x_{n}) to an output F(𝐱)F(\mathbf{x}). As an example, assume that the input represents an image, and the xix_{i} are the intensities of its pixels. Assume we want to determine whether the image contains some element, say a dog, or a cat. There are deep networks that have been training with millions of images and can provide an answer by assigning a score (given by the value of F(x)F(x)) to the presence or absence of the chosen element (see e.g. [4]). If the element is present in the image then the score is high, if not it is low. Each possible element has an associated score function, say Fdog(𝐱)F_{\text{dog}}(\mathbf{x}), Fcat(𝐱)F_{\text{cat}}(\mathbf{x}), etc. A problem that IG is designed to solve is to determine the contribution of each of its inputs to the output. In the example where the input is an image, and the network provides scores for the presence of an element (dog, cat,…), the inputs (pixels) that contribute most to the output are expected to be the ones in the area of the image where that element appears. This provides a way to locate that element within the image, e.g. where exactly the dog or the cat appears in the image.

IG was designed requiring it to satisfy a number of desirable properties, particularly sensitivity and implementation invariance. To define sensitivity we will need to establish a baseline input where the element or feature is absent (in image recognition a black image 𝐱base=(0,,0)\mathbf{x}^{\text{base}}=(0,\dots,0) may serve the purpose). Then, the definitions of sensitivity and implementation invariance are as follows.

  • An attribution method satisfies Sensitivity if for every input and baseline that differ in one feature but have different predictions then the differing feature should be given a non-zero attribution.

  • An attribution method satisfies Implementation Invariance if the attributions are always identical for two functionally equivalent networks (two networks are functionally equivalent if their outputs are equal for all inputs).

The solution proposed in [5] is as follows. Given a baseline 𝐱base\mathbf{x}^{\text{base}} and an input 𝐱input\mathbf{x}^{\text{input}}, then the attribution assigned to the ii-th coordinate of the input is:

(1) IGi=(xiinputxibase)01F(γ(t))xi𝑑t,\text{IG}_{i}=(x_{i}^{\text{input}}-x_{i}^{\text{base}})\int_{0}^{1}\frac{\partial F(\gamma(t))}{\partial x_{i}}\,dt\,,

where γ(t)=𝐱base+t(𝐱input𝐱base)\gamma(t)=\mathbf{x}^{\text{base}}+t\,(\mathbf{x}^{\text{input}}-\mathbf{x}^{\text{base}}).

In the next section we will justify this formula, and discuss two additional properties of IG:

  • Completeness. The attributions add up to the difference between the output of FF at the input 𝐱input\mathbf{x}^{\text{input}} and the baseline 𝐱base\mathbf{x}^{\text{base}}.

  • Symmetry Preserving. Two input variables are symmetric w.r.t. a function FF if swapping them does not change the function. An attribution method is symmetry preserving, if for all inputs that have identical values for symmetric variables and baselines that have identical values for symmetric variables, the symmetric variables receive identical attributions.

1.2. IG Completeness and Symmetry Preserving

1.2.1. Completeness

The solution proposed by the authors can be understood as an application of the Gradient Theorem for line integrals. Under appropriate assumptions on function FF we have:

(2) F(γ(1))F(γ(0))=γF(𝐱)𝑑𝐱=γi=1nF(𝐱)xidxi=i=1nγF(𝐱)xi𝑑xi,F(\gamma(1))-F(\gamma(0))=\int_{\gamma}\nabla F(\mathbf{x})\cdot\,d\mathbf{x}=\int_{\gamma}\sum_{i=1}^{n}\frac{\partial F(\mathbf{x})}{\partial x_{i}}\,dx_{i}=\sum_{i=1}^{n}\int_{\gamma}\frac{\partial F(\mathbf{x})}{\partial x_{i}}\,dx_{i}\,,

where γ:[0,1]n\gamma:[0,1]\to\mathbb{R}^{n} is a smooth (continuously differentiable) path, and \nabla is the nabla operator, i.e., F=(Fx1,,Fxn)\nabla F=\left(\frac{\partial F}{\partial x_{1}},\dots,\frac{\partial F}{\partial x_{n}}\right). The attribution IGi\text{IG}_{i} to the ii-th variable xix_{i} of the score increase with respect to baseline F(𝐱input)F(𝐱base)F(\mathbf{x}^{\text{input}})-F(\mathbf{x}^{\text{base}}) is the ii-th term of the final sum, i.e., IGi=γF(𝐱)xi𝑑xi\text{IG}_{i}=\int_{\gamma}\frac{\partial F(\mathbf{x})}{\partial x_{i}}\,dx_{i}.

Note that the result is highly dependent on the path γ\gamma chosen. The authors of IG claim that the best (in fact canonical) choice of path is a straight line from baseline to input, i.e., γ(t)=𝐱base+t(𝐱input𝐱base)\gamma(t)=\mathbf{x}^{\text{base}}+t\,(\mathbf{x}^{\text{input}}-\mathbf{x}^{\text{base}}). With this choice we get dxi=(xiinputxibase)dtdx_{i}=(x_{i}^{\text{input}}-x_{i}^{\text{base}})\,dt, and equation (1) follows.

If (2) holds, then it is easy to see that IG also satisfies the completeness property:

F(𝐱input)F(𝐱base)=i=1nIGi.F(\mathbf{x}^{\text{input}})-F(\mathbf{x}^{\text{base}})=\sum_{i=1}^{n}\text{IG}_{i}\,.

This is Proposition 1 in their paper. However this result depends on the Gradient Theorem for line integrals, which requires the function FF to be continuously differentiable everywhere. This cannot be the case for deep networks, which involve functions such as ReLU and max pooling that are not everywhere differentiable. As a fix the authors restate the Gradient Theorem assuming only that FF is continuous everywhere and differentiable almost everywhere. However this cannot work, the particular case of the Gradient Theorem in 1-dimension is the (second) Fundamental Theorem of Calculus, which does not hold in general under those premises—Cantor’s staircase function provides a well known counterexample. There are also issues concerning allowable paths, e.g. for the function f(x,y)=max(x,y)f(x,y)=\text{max}(x,y) the partial derivatives f/x{\partial f}/{\partial x} and f/y{\partial f}/{\partial y} don’t even exist at the points of the line x=yx=y.

Our Proposition 1 in the next section states a version of the Gradient Theorem that can be applied to deep networks.

1.2.2. Symmetry Preserving

Theorem 1 of [5] states that IG, with paths that are a straight line from baseline to input, is the unique path method that is symmetry-preserving. However this theorem is not stated in the paper with full mathematical rigor, and the proof provided contains some inconsistencies.111Look for instance at the value of function ff in the region where xiax_{i}\leq a and xjbx_{j}\geq b. Here we will present a completely rigorous formulation of a theorem that we believe captures the original authors’ intention, and provide its full proof.

2. Main Results

Here we state an appropriate generalization of the Gradient Theorem that can be applied to deep networks, and study the symmetry-preserving property of IG with straight-line paths.

First, we need to extend the class of functions to which we want to apply the theorem so that it includes functions implemented by common deep networks. We will do so by introducing Lipschitz continuous functions.

Definition 1.

A function F:SnmF:S\subseteq\mathbb{R}^{n}\to\mathbb{R}^{m} is Lipschitz continuous if there is a constant K0K\geq 0 such that f(𝐱)f(𝐲)K𝐱𝐲||f(\mathbf{x})-f(\mathbf{y})||\leq K||\mathbf{x}-\mathbf{y}|| for every 𝐱,𝐲S\mathbf{x},\mathbf{y}\in S, where ||||||\cdot|| represents Euclidean distance.

For univariate functions, continuity and almost everywhere differentiability is a necessary condition to make sense of the integral of a derivative. The following result ensures that such condition is satisfied for multivariate Lipschitz continuous functions.

Rademacher’s theorem.

If UU is an open subset of n\mathbb{R}^{n} and F:UmF:U\to\mathbb{R}^{m} is Lipschitz continuous, then f is differentiable almost everywhere in UU.

Proof.

See e.g. [1], Theorem 3.1.6., or [2] Theorem 3.1. ∎

Rademacher’s theorem ensures that the function in our proposition 1, that we state next, is differentiable almost everywhere. However that does not mean that such function is differentiable almost everywhere on a given path—e.g. the function max(x,y)\text{max}(x,y) is everywhere continuous, and almost everywhere differentiable on 2\mathbb{R}^{2}, but is not differentiable at any point of the line x=yx=y. So differentiability on the path needs to be included as an additional premise.

Proposition 1 (Gradient Theorem for Lipschitz Continuous Functions).

Let UU be an open subset of n\mathbb{R}^{n}. If F:UF:U\to\mathbb{R} is Lipschitz continuous, and γ:[0,1]U\gamma:[0,1]\to U is a smooth path such that FF is differentiable at γ(t)\gamma(t) for almost every t[0,1]t\in[0,1], then

γF(𝐱)𝑑𝐱=F(γ(1))F(γ(0)).\int_{\gamma}\nabla F(\mathbf{x})\cdot d\mathbf{x}=F(\gamma(1))-F(\gamma(0))\,.
Proof.

The path γ\gamma is continuously differentiable on a compact set (the interval [0,1][0,1]), hence it is Lipschitz continuous (because its derivative is continuous and so bounded on [0,1][0,1]). The composition of two Lipschitz continuous functions is Lipschitz continuous, hence tF(γ(t))t\mapsto F(\gamma(t)) is Lipschitz continuous, which implies absolutely continuous. By the Fundamental Theorem of Calculus for absolutely continuous functions222See [3] sec. 33.2, theorem 6. we have

F(γ(1))F(γ(0))=01ddtF(γ(t))𝑑t.F(\gamma(1))-F(\gamma(0))=\int_{0}^{1}\frac{d}{dt}F(\gamma(t))\,dt\,.

By the multivariate chain rule we have

ddtF(γ(t))=F(γ(t))γ(t)\frac{d}{dt}F(\gamma(t))=\nabla F(\gamma(t))\cdot\gamma^{\prime}(t)

wherever FF is differentiable (for almost every t[0,1]t\in[0,1] by hypothesis). Hence

01ddtF(γ(t))𝑑t=01F(γ(t))γ(t)𝑑t=γF(𝐱)𝑑𝐱,\int_{0}^{1}\frac{d}{dt}F(\gamma(t))\,dt=\int_{0}^{1}\nabla F(\gamma(t))\cdot\gamma^{\prime}(t)\,dt=\int_{\gamma}\nabla F(\mathbf{x})\cdot d\mathbf{x}\,,

and the result follows.

Next, we will look at results intended to capture the symmetry-preserving properties of IG.

Definition 2.

A multivariate function FF with nn variables is symmetric in variables xix_{i} and xjx_{j}, iji\neq j, if F(x1,,xi,,xj,,xn)=F(x1,,xj,,xi,,xn)F(x_{1},\dots,x_{i},\dots,x_{j},\dots,x_{n})=F(x_{1},\dots,x_{j},\dots,x_{i},\dots,x_{n}).

Definition 3.

The smooth path γ:[0,1]𝐑n\gamma:[0,1]\to\mathbf{R}^{n} is said to be IG-symmetry preserving for variables xix_{i} and xjx_{j} if γi(0)=γj(0)\gamma_{i}(0)=\gamma_{j}(0) and γi(1)=γj(1)\gamma_{i}(1)=\gamma_{j}(1) implies γF(𝐱)xi𝑑xi=γF(𝐱)xj𝑑xj\int_{\gamma}\frac{\partial F(\mathbf{x})}{\partial x_{i}}\,dx_{i}=\int_{\gamma}\frac{\partial F(\mathbf{x})}{\partial x_{j}}\,dx_{j} for every function FF verifying the hypotheses of proposition 1 that is symmetric in its variables xix_{i} and xjx_{j}.

Definition 4.

A path γ:[0,1]n\gamma:[0,1]\to\mathbb{R}^{n} is monotonic if for each i=1,,ni=1,\dots,n we have t1t2γi(t1)γi(t2)t_{1}\leq t_{2}\Rightarrow\gamma_{i}(t_{1})\leq\gamma_{i}(t_{2}) or t1t2γi(t1)γi(t2)t_{1}\leq t_{2}\Rightarrow\gamma_{i}(t_{1})\geq\gamma_{i}(t_{2}), where γi(t)\gamma_{i}(t) represents the ii-th coordinate of γ(t)\gamma(t).333In other words, each γi\gamma_{i} is either increasing, or decreasing. Note that γ\gamma could be increasing in some coordinates and decreasing in other The path γ\gamma is strictly monotonic if the inequalities hold replacing them with strict inequalities, i.e., t1<t2γi(t1)<γi(t2)t_{1}<t_{2}\Rightarrow\gamma_{i}(t_{1})<\gamma_{i}(t_{2}) or t1<t2γi(t1)>γi(t2)t_{1}<t_{2}\Rightarrow\gamma_{i}(t_{1})>\gamma_{i}(t_{2}).

Next theorem is intended to capture the symmetry-preserving properties of IG. The proof follows closely the one given in [5].

Theorem 1.

Given i,j{1,,n}i,j\in\{1,\dots,n\}, iji\neq j, real numbers a<ba<b, and a strictly monotonic smooth path γ:[0,1](a,b)n\gamma:[0,1]\to(a,b)^{n} such that γi(0)=γj(0)\gamma_{i}(0)=\gamma_{j}(0) and γi(1)=γj(1)\gamma_{i}(1)=\gamma_{j}(1), then the following statements are equivalent:

  • (1)

    For every t[0,1]t\in[0,1], γi(t)=γj(t)\gamma_{i}(t)=\gamma_{j}(t).

  • (2)

    For every function F:[a,b]nF:[a,b]^{n}\to\mathbb{R} symmetric in xix_{i} and xjx_{j} and verifying the premises of proposition 1 with U=(a,b)nU=(a,b)^{n} we have γF(𝐱)xi𝑑xi=γF(𝐱)xj𝑑xj\int_{\gamma}\frac{\partial F(\mathbf{x})}{\partial x_{i}}\,dx_{i}=\int_{\gamma}\frac{\partial F(\mathbf{x})}{\partial x_{j}}\,dx_{j} (i.e., γ\gamma is IG-symmetry preserving for variables xix_{i} and xjx_{j}).

Proof.

Proof of (1) \Rightarrow (2). Since γi(t)=γj(t)\gamma_{i}(t)=\gamma_{j}(t) for every t[0,1]t\in[0,1], and FF is symmetric with respect to variables xix_{i} and xjx_{j}, we have F(γ(t))xi=F(γ(t))xj\frac{\partial F(\gamma(t))}{\partial x_{i}}=\frac{\partial F(\gamma(t))}{\partial x_{j}} for almost every t[0,1]t\in[0,1]. Hence they have the same integral.

Proof of (2) \Rightarrow (1). Without loss of generality we will assume that γi\gamma_{i} and γj\gamma_{j} are increasing, so that γi(0)=γj(0)<γi(1)=γj(1)\gamma_{i}(0)=\gamma_{j}(0)<\gamma_{i}(1)=\gamma_{j}(1). Next, assume that (1) is not true. Then, for some t0(0,1)t_{0}\in(0,1) we have γi(t0)γj(t0)\gamma_{i}(t_{0})\neq\gamma_{j}(t_{0}). Assume wlog γi(t0)<γj(t0)\gamma_{i}(t_{0})<\gamma_{j}(t_{0}). Let (u,v)(u,v) be the maximum interval containing t0t_{0} such that γi(t)<γj(t)\gamma_{i}(t)<\gamma_{j}(t) for every t(u,v)t\in(u,v). Since (u,v)(u,v) is maximum, and γi\gamma_{i}, γj\gamma_{j} are increasing, then γi(t),γj(t)<γi(u)=γj(u)\gamma_{i}(t),\gamma_{j}(t)<\gamma_{i}(u)=\gamma_{j}(u) for t<ut<u, and γi(t),γj(t)>γi(v)=γj(v)\gamma_{i}(t),\gamma_{j}(t)>\gamma_{i}(v)=\gamma_{j}(v) for t>vt>v.

Define g:[a,b]g:[a,b]\to\mathbb{R} as follows:

g(x)={0 if ax<uxu if uxvv if v<xbg(x)=\begin{cases}0&\text{ if }a\leq x<u\\ x-u&\text{ if }u\leq x\leq v\\ v&\text{ if }v<x\leq b\end{cases}

and F(𝐱)=g(xi)g(xj)F(\mathbf{x})=g(x_{i})g(x_{j}). Then FF is symmetric in xix_{i} and xjx_{j}, and verifies the premises of proposition 1. For t[a,b]t\notin[a,b] we have that F(γ(t))F(\gamma(t)) is constant, hence F(γ(t))xi=F(γ(t))xj=0\frac{\partial F(\gamma(t))}{\partial x_{i}}=\frac{\partial F(\gamma(t))}{\partial x_{j}}=0. For t[u,v]t\in[u,v] we have

F(γ(t))xi\displaystyle\frac{\partial F(\gamma(t))}{\partial x_{i}} =xi((xiu)(xju))=(xju)=γj(t)u,\displaystyle=\frac{\partial}{\partial x_{i}}((x_{i}-u)(x_{j}-u))=(x_{j}-u)=\gamma_{j}(t)-u\,,
F(γ(t))xj\displaystyle\frac{\partial F(\gamma(t))}{\partial x_{j}} =xj((xiu)(xju))=(xiu)=γi(t)u.\displaystyle=\frac{\partial}{\partial x_{j}}((x_{i}-u)(x_{j}-u))=(x_{i}-u)=\gamma_{i}(t)-u\,.

By hypothesis γi(t)<γj(t)\gamma_{i}(t)<\gamma_{j}(t), hence γF(𝐱)xi𝑑xi>γF(𝐱)xj𝑑xj\int_{\gamma}\frac{\partial F(\mathbf{x})}{\partial x_{i}}\,dx_{i}>\int_{\gamma}\frac{\partial F(\mathbf{x})}{\partial x_{j}}\,dx_{j}, which is a contradiction.

This completes the proof. ∎

Corollary 1.

Let 𝐩=(p1,,pn)\mathbf{p}=(p_{1},\dots,p_{n}) and 𝐪=(q1,,qn)\mathbf{q}=(q_{1},\dots,q_{n}) be two points in the open set UnU\subseteq\mathbb{R}^{n}, such that pi=pjp_{i}=p_{j} and qi=qjq_{i}=q_{j}. Then, the (straight line) path γ(t)=𝐩+t(𝐪𝐩)\gamma(t)=\mathbf{p}+t\,(\mathbf{q}-\mathbf{p}) is IG-symmetry preserving for variables xix_{i}, xjx_{j} for every function FF that is symmetric in xix_{i} and xjx_{j} and verifies the hypotheses of proposition 1. Furthermore, if pj1==pjrp_{j_{1}}=\dots=p_{j_{r}} and qj1==qjrq_{j_{1}}=\dots=q_{j_{r}}, and γ\gamma is IG-symmetric preserving for (each pair of) variables xi1,,xirx_{i_{1}},\dots,x_{i_{r}} in the sense of definition 3, then t(γi1(t),,γir(t))t\mapsto(\gamma_{i_{1}}(t),\dots,\gamma_{i_{r}}(t)) is a straight line in r\mathbb{R}^{r}.

Next, we will look at a simple example illustrating the result stated in theorem 1.

Example 1.

Figure (1) illustrates the failure to preserve symmetry when the path is not a straight line between baseline and input. The function used is F(x1,x2)=x1x2F(x_{1},x_{2})=x_{1}x_{2}, which is symmetric, and the path is a curve joining (0,0)(0,0) and (1,1)(1,1). The attributions from IG are

IG1\displaystyle\text{IG}_{1} =γF(x1,x2)x1𝑑x1=o1x2𝑑x1,\displaystyle=\int_{\gamma}\frac{\partial F(x_{1},x_{2})}{\partial x_{1}}\,dx_{1}=\int_{o}^{1}x_{2}\,dx_{1}\,,
IG2\displaystyle\text{IG}_{2} =γF(x1,x2)x2𝑑x2=01x1𝑑x2.\displaystyle=\int_{\gamma}\frac{\partial F(x_{1},x_{2})}{\partial x_{2}}\,dx_{2}=\int_{0}^{1}x_{1}\,dx_{2}\,.
Refer to caption
Figure 1. Example.

We have IG1=\text{IG}_{1}= area under the path γ\gamma, and IG2=\text{IG}_{2}= area above the path. Their sum is IG1+IG2=1\text{IG}_{1}+\text{IG}_{2}=1, the whole area of the square with vertices (0,0)(0,0) and (1,1)(1,1), which equals F(1,1)=1F(1,1)=1, and the Gradient Theorem holds. However IG1<IG2\text{IG}_{1}<\text{IG}_{2}. If we used instead the straight line path joining (0,0)(0,0) and (1,1)(1,1), then the IG attributions would be equal.

Note that the uniqueness of IG (using straight line paths) is not fully captured by our results, and in general does not hold (under the definition of “symmetry-preserving” as worded in the IG paper, which we tried to formally capture in definition 3). All we can tell is that if pi=pjp_{i}=p_{j} and qi=qjq_{i}=q_{j} then γi(t)=γj(t)\gamma_{i}(t)=\gamma_{j}(t) for every t[0,1]t\in[0,1], but if the premises fail (i.e., pipjp_{i}\neq p_{j} or qiqjq_{i}\neq q_{j}) then nothing forces the path to be a straight line. To illustrate this point we give next a counterexample showing a way to select paths that are not straight lines in general, and still verify the definition of IG-symmetry preserving.

Counterexample 1.

Consider the following path between points 𝐩=(p1,p2)\mathbf{p}=(p_{1},p_{2}) and 𝐪=(q1,q2)\mathbf{q}=(q_{1},q_{2}) in 2\mathbb{R}^{2}:

γ1(t)\displaystyle\gamma_{1}(t) =p1+t(q1p1)+t(t1)((p1p2)2+(q1q2)2)sgn(q1p1),\displaystyle=p_{1}+t\,(q_{1}-p_{1})+t(t-1)((p_{1}-p_{2})^{2}+(q_{1}-q_{2})^{2})\,\,\text{sgn}(q_{1}-p_{1})\,,
γ2(t)\displaystyle\gamma_{2}(t) =p2+t(q2p2)+t(t1)((p1p2)2+(q1q2)2)sgn(q2p2).\displaystyle=p_{2}+t\,(q_{2}-p_{2})+t(t-1)((p_{1}-p_{2})^{2}+(q_{1}-q_{2})^{2})\,\,\text{sgn}(q_{2}-p_{2})\,.

We multiply the last term of each expression by the sign function444The sign function is defined sgn(x)=x|x|\text{sgn}(x)=\frac{x}{|x|} if x0x\neq 0, and sgn(0)=0\text{sgn}(0)=0. to make sure that the paths are monotonic (this shows that requiring the paths to be monotonic does not affect the result.) Also note that the assignment of path (𝐩,𝐪)γ(\mathbf{p},\mathbf{q})\mapsto\gamma, where γ\gamma is defined as above, is symmetric in the sense that swapping the indexes 11 and 22 in the expression produces another expression that is equivalent to the original, so the assignment of path is symmetric with respect to the coordinates x1x_{1} and x2x_{2}. Also, we have that if p1=p2p_{1}=p_{2} and q1=q2q_{1}=q_{2} then γ(t)=𝐩+t(𝐪𝐩)\gamma(t)=\mathbf{p}+t\,(\mathbf{q}-\mathbf{p}), which, according to theorem 1, is IG-symmetry preserving for variables x1x_{1}, x2x_{2} for every function FF that is symmetric in x1x_{1} and x2x_{2} and verifies the hypotheses of proposition 1. However, if p1p2p_{1}\neq p_{2} or q1q2q_{1}\neq q_{2}, the quadratic terms in tt make γ\gamma a curve that is not a straight line in general, and hence it differs from the path used in the IG attribution method. The symmetry preserving property is not violated because in the cases where the path is not a straight line the premises of the definition don’t apply, so theorem 1 still holds.

Admittedly this counterexample is an artificial modification of a straight-line. IG (with straight line paths) is a simple path-based symmetry-preserving attribution method, and we see no reason to replace it with a different method using non straight-line paths without justification.

3. Conclusions

We have rigorously stated and proved that Integrated Gradients has completeness and symmetry preserving properties. The premises used to prove the result makes it suitable for functions implemented by common deep networks.

On the other hand we have shown that IG with straight line paths is not the unique path method that is symmetry-preserving, in fact there are path methods that verify the definition of symmetry-preserving but don’t necessarily use straight line paths for all combinations of baseline and input. Note that this should not be taken as an argument against using straight line paths in IG, in fact straight lines are still the simplest paths that provide the desired results.

References

  • [1] Federer, Herbert (1969). Geometric measure theory, Die Grundlehren der mathematischen Wissenschaften, 153, Berlin–Heidelberg–New York: Springer-Verlag, pp. xiv+676, ISBN 978-3-540-60656-7, MR 0257325, Zbl 0176.00801.
  • [2] Heinonen, Juha (2004). Lectures on Lipschitz Analysis. URL: http://www.math.jyu.fi/research/reports/rep100.pdf
  • [3] A. N. Kolmogorov, S. V. Fomin (1970). Introductory Real Analysis. Dover Books on Mathematics.
  • [4] Karen Simonyan, Andrew Zisserman (2015). Very Deep Convolutional Networks for Large-Scale Image Recognition. arXiv preprint arXiv:1409.1556 [cs.CV]
  • [5] Mukund Sundararajan, Ankur Taly, and Qiqi Yan (2017). Axiomatic attribution for deep networks. arXiv preprint arXiv:1703.01365 [cs.LG]