arxiv_dump / txt /2111.13236.txt
billxbf's picture
Upload 101 files
8f1929a verified
raw
history blame
69.2 kB
Joint inference and input optimization
in equilibrium networks
Swaminathan Gurumurthy
Carnegie Mellon UniversityShaojie Bai
Carnegie Mellon UniversityZachary Manchester
Carnegie Mellon University
J. Zico Kolter
Carnegie Mellon University
Bosch Center for AI
Abstract
Many tasks in deep learning involve optimizing over the inputs to a network to
minimize or maximize some objective; examples include optimization over latent
spaces in a generative model to match a target image, or adversarially perturbing
an input to worsen classifier performance. Performing such optimization, however,
is traditionally quite costly, as it involves a complete forward and backward pass
through the network for each gradient step. In a separate line of work, a recent
thread of research has developed the deep equilibrium (DEQ) model, a class of
models that foregoes traditional network depth and instead computes the output of a
network by finding the fixed point of a single nonlinear layer. In this paper, we show
that there is a natural synergy between these two settings. Although, naively using
DEQs for these optimization problems is expensive (owing to the time needed to
compute a fixed point for each gradient step), we can leverage the fact that gradient-
based optimization can itself be cast as a fixed point iteration to substantially
improve the overall speed. That is, we simultaneously both solve for the DEQ
fixed point andoptimize over network inputs, all within a single “augmented”
DEQ model that jointly encodes both the original network and the optimization
process. Indeed, the procedure is fast enough that it allows us to efficiently train
DEQ models for tasks traditionally relying on an “inner” optimization loop. We
demonstrate this strategy on various tasks such as training generative models while
optimizing over latent codes, training models for inverse problems like denoising
and inpainting, adversarial training and gradient based meta-learning.
1 Introduction
Many settings in deep learning involve optimization over the inputs to a network to minimize some
desired loss. For example, for a “generator” network G:Z!X that maps from latent space Zto
an observed space X, it may be desirable to find a latent vector z2Z that most closely produces
some target output x2X by solving the optimization problem (e.g. [10, 13])
minimize
z2ZkxG(z)k2
2: (1)
As another example, constructing adversarial examples for classifiers [ 28,53] typically involves
optimizating over a perturbation to a given input; i.e., given a classifier network g:X!Y , task loss
Correspondence to: Swaminathan Gurumurthy <[email protected]>
Code available at https://github.com/locuslab/JIIO-DEQ
Preprint. Under review.arXiv:2111.13236v1 [cs.LG] 25 Nov 2021`:Y!R+, and a sample x2X, we want to solve
maximize
kk`(g(x+)): (2)
More generally, a wide range of inverse problems [ 10] and other auxiliary tasks [ 22,3] in deep
learning can also be formulated in such a manner.
Orthogonal to this line of work, a recent trend has focused on the use of an implicit layer within deep
networks to avoid traditional depth. For instance, Bai et al. [5]introduced deep equilibrium models
(DEQs) which instead treat the network as repeated applications of a single layer and compute the
output of the network as a solution to an equilibrium-finding problem instead of simply specifying
a sequence of non-linear layer operations. Bai et al. [5]and subsequent work [ 6] have shown that
DEQs can achieve results competitive with traditional deep networks for many realistic tasks.
In this work, we highlight the benefit of using these implicit models in the context of input optimization
routines. Specifically, because optimization over inputs itself is typically done via an iterative method
(e.g., gradient descent), we can combine this optimization fixed-point iteration with the forward
DEQ fixed point iteration all within a single “augmented” DEQ model that simultaneously performs
forward model inference as well as optimization over the inputs. This enables the models to more
quickly perform both the inference and optimization procedures, and the resulting speedups further
allow us to train networks that use such “bi-level” fixed point passes. In addition, we also show
a close connection between our proposed approach and the primal-dual methods for constrained
optimization.
We illustrate our methods on 4 tasks that span across different domains and problems: 1) training
DEQ-based generative models while optimizing over latent codes; 2) training models for inverse
problems such as denoising and inpainting; 3) adversarial training of implicit models; and 4) gradient-
based meta-learning. We show that in all cases, performing this simultaneous optimization and
forward inference accelerates the process over a more naive inner/outer optimization approach. For
instance, using the combined approach leads to a 3.5-9x speedup for generative DEQ networks,
a 3x speedup in adverarial training of DEQ networks and a 2.5-3x speedup for gradient based
meta-learning. In total, we believe this work points to a variety of new potential applications for
optimization with implicit models.
2 Related Work
Implicit layers. Layers with implicitly defined depth have gained tremendous popularity in recent
years[ 46,19,29]. Rather than a static computation graph, these layers define a condition on the output
that the model must satisfy, which can represent “infinite” depth, be directly differentiated through
via the implicit function theorem [ 47], and are memory-efficient to train. Some recent examples of
implicit layers include optimization layers [ 16,1], deep equilibrium models[ 5,6,68,40,52], neural
ordinary differential equations (ODEs) [ 14,18,61], logical structure learning [ 67], and continuous
generative models [30].
In particular, deep equilibrium models (DEQs) [ 5] define the output of the model as the fixed point
of repeated applications of a layer. They compute this using black-box root-finding methods[ 5] or
accelerated fixed-point iterations [ 36] (e.g., Broyden’s method [ 11]). In this work, we propose an
efficient approach to perform input optimization with the DEQ by simultaneously optimizing over
the inputs and solving the forward fixed point of an equilibrium model as a joint, augmented system.
As related work, Jeon et al. [36] introduce fixed point iteration networks that generalize DEQs to
repeated applications of gradient descent over variables. However, they don’t address the specific
formulation presented in this paper, which has a number of practical use cases (e.g., adversarial
training). Lu et al. [52] proposes an implicit version of normalizing flows by formulating a joint
root-finding problem that defines an invertible function between the input xand outputz?. Perhaps
the most relevant approach to our work is Gilton et al. [26], which specifically formulates inverse
imaging problems as a DEQ model. In contrast, our approach focuses on solving input optimization
problems where the network of interest is already a DEQ, and thus the combined optimization and
forward inference task leads to a substantially different set of update equations and tradeoffs.
2Input optimization in deep learning. Many problems in deep learning can be framed as optimizing
over the inputs to minimize some objective . Some canonical examples of this include finding
adversarial examples [ 53,45], solving inverse problems [ 10,13,56], learning generative models [ 9,
72], meta-learning [ 58,22,74,32] etc. For most of these examples, input optimization is typically
done using gradient descent on the input, i.e., we feed the input through the network and compute
some loss, which we minimize by optimizing over the input with gradient descent. While some of
these problems might not require differentiating through the entire optimization process, many do
(introduced below), and can further slow down training and impose massive memory requirements.
Input optimization has recently been applied to train generative models. Zadeh et al. [72], Bojanowski
et al. [9]proposed to train generator networks by jointly optimizing the parameters and the latent
variables corresponding to each example. Similarly, optimizing a latent variable to make the
corresponding output match a target image is common in decoder-only models like GANs to get
correspondences [ 10,39], and has been found useful to stabilize GAN training [ 71]. However, in all
of these cases, the input is optimized for just a few (mostly 1) iterations. In this work, we present a
generative model, where we optimize and find the optimal latent code for each image at each training
step. Additionally, Bora et al. [10], Chang et al. [13] showed that we can take a pretrained generative
model and use it as a prior to solve for the likely solutions to inverse problems by optimizing on
the input space of the generative model (i.e., unsupervised inverse problem solving). Furthermore,
Diamond et al. [15], Gilton et al. [25], Gregor and LeCun [31] have shown that networks can also be
trained to solve specific inverse problems by effectively unrolling the optimization procedure and
iteratively updating the input. We demonstrate our approach in the unsupervised setting as in Bora
et al. [10], Chang et al. [13], but also show flexible extension of our framework to train implicit
models for supervised inverse problem solving.
Another crucial application of input optimization is to find adversarial examples [ 64,28]. This
manifests as optimizing an objective that incentivices an incorrect prediction by the classifier, while
constraining the input to be within a bounded region of the original input. Many attempts have been
made on the defense side [ 57,37,65,69]. The most successful strategy thus far has been adversarial
training with a projected gradient descent (PGD) adversary [ 53] which involves training the network
on the adversarial examples computed using PGD online during training . We show that our joint
optimization approach can be easily applied to this setting, allowing us to train implicit models to
perform competitively with PGD in guaranteeing adversarial robustness, but at much faster speeds.
While the examples above were illustrated with non-convex networks, attempts have also been made
to design networks whose output is a convex function of the input [ 2]. This allows one to use more
sophisticated optimization algorithms, but usually at a heavy cost of model capacity. They have been
demonstrated to work in a variety of problems including multi-label prediction, image completion
[2], learning stable dynamical systems [44] and optimal transport mappings [54], MPC [12], etc.
3 Joint inference and input optimization in DEQs
Here we present our main methodological contribution, which sets up an augmented DEQ that jointly
performs inference and input optimization over an existing DEQ model. We first define the base
DEQ model, and then illustrate a joint approach that simultaneously finds it’s forward fixed point and
optimizes over its inputs. We discuss several methodological details and extensions.
3.1 Preliminaries: DEQ-based models
To begin with, we recall the deep equilibrium model setting from Bai et al. [5], but with the notation
slightly adapted to better align with its usage in this paper. Specifically, we consider an input-injected
layerf:ZX!Z whereZdenotes the hidden state of the network, Xdenotes the input space,
anddenotes the parameters of the layer. Given an input x2X, computing the forward pass in a
DEQ model involves finding a fixed point z?(x)2Z, such that
z?
(x) =f(z?
(x);x); (3)
which (under proper stability conditions) corresponds to the “infinite depth” limit of repeatedly
applying the ffunction. We emphasize that under this setting, we can effectively think of z?
itself
3as the implicitly defined network (which thus is also parameterized by ), and one can differentiate
through this “network” via the implicit function theorem [8, 47].
The fixed point of a DEQ could be computed via the simple forward iteration
z+:=f(z;x) (4)
starting at some artibrary initial value of z(typically 0). However, in practice DEQ models
will typically compute this fixed point not simply by iterating the function f, but by using a
more accelerated root-finding or fixed-point approach such as Broyden’s method [ 11] or Anderson
acceleration [ 4,66]. Further, although little can be said about e.g., the existence or uniqueness of these
fixed points in general (though there do exist restrictive settings where this is possible [ 68,59,23]), in
practice a wide suite of techniques have been used to ensure that such fixed points exist, can be found
using relatively few function evaluations, and are able to competitively model large-scale tasks [ 5,6].
3.2 Joint inference and input optimization
Now we consider the setting of performing input optimization for such a DEQ model. Specifically,
consider the task of attempting to optimize the input x2X to minimize some loss `:ZY! R+.
minimize
x2X`(z?
(x);y) (5)
wherey2Y represents the data point. To solve this, we typically perform such an optimization via
e.g., gradient descent, which repeats the update
x+:=x @`(z?
(x);y)
@x>
(6)
until convergence, where we use term z?alone to denote the fixed output of the network z?
(i.e., just
as a fixed output rather than a function). Using the chain rule and the implicit function theorem, we
can further expand update (6) using the following analytical expression of the gradient:
@`(z?
(x);y)
@x=@`(z?;y)
@z?@z?
(x)
@x=@`(z?;y)
@z?
I@f(z?;x)
z?>@f(z?;x)
@x(7)
Thinking about z?
as an implicit function of xpermits us to combine the fixed-point equation in Eq. 4
(onz) with this input optimization update (on x), thus performing a joint forward update:

z+
x+
:=f(z;x)
x @f(z;x)
@x>
I@f(z;x)
@z>@`(z;y)
@z>
(8)
It should be apparent that, if both iterates converge, then they have converged to a simultaneous fixed
pointz?and an optimal x?value for the optimization problem (5). However, simply performing
this update can still be inefficient, because computing the inverse Jacobian in Eq. (7)is expensive
and typically computed via an iterative update – namely, we would first compute the variable
=
I@f(z;x)
@z>@`(z;y)
@z>via the following iteration (i.e., a Richardson iteration [60]):
+:=@f(z;x)
@z>
+@`(z;y)
@z>
: (9)
Therefore, to efficiently solve the joint inference and input optimization problem, we propose
combining all three iterative procedures into the update
2
4z+
+
x+3
5:=2
4f(z;x)@f(z;x)
@z>+@`(z;y)
@z>
x @f(z;x)
@x>3
5 (10)
Like Eq. (8), if this joint process converges to a fixed point, then it corresponds to a simultaneous
optimum of both the inference and optimization processes. Such a formulation is especially appealing,
as the iteration (10) isitself just an augmented DEQ network v?
(y)(i.e., with input injection y)
whose forward pass optimizes on a joint inference-optimization space v= (x;;z ). Moreover, we
can use standard techniques to differentiate through thisprocess, though there are also optimizations
4.
Figure 1: Left: Performing each gradient update on DEQ inputs requires a fixed point computation in
the forward and backward pass. Right: Solving the 3 fixed points simultaneously as an “augmented”
DEQ where the targets yact as input and the function Frepresents the joint fixed point updates in
Eq. 11
we can apply in several settings that we discuss below. This is in contrast to prior works where f
is an explicit deep neural network, where the model forward-pass and optimization processes are
disentangled and have to be dealt with separately. We illustrated this in Figure 1 where the figure on
the left shows the input optimization naively performed using gradient descent in DEQs v/s the figure
on the right which shows the joint updates performed using the augmented DEQ network.
As a final note, we mention that, in practice, just as the gradient-descent update has a step size , it is
often beneficial to add similar “damping” step sizes to the other updates as well. This leads to the full
iteration over the augmented DEQ
2
4z+
+
x+3
5:=2
6664(1 z)z+ zf(z;x)
(1 )+ 
@f(z;x)
@z>
+
@`(z;y)
@z>
x x
@f(z;x)
@x>
3
7775(11)
Finally, in order to speed up convergence, as is common in DEQ models, we apply a more involved
fixed point solver, such as Anderson acceleration, on top of this naive iteration. We analyze the effect
of these different root-finding approaches in the Appendix.
Notes on Convergence Our treatment of the above system as an augmented DEQ allows us to
borrow results from [ 68,7] to ensure convergence of the fixed point iteration. Specifically, if we
assume the joint Jacobian of the fixed point iterations we describe are strongly monotone with
smoothness parameter mand Lipschitz constant L, then by standard arguments (see e.g., Section
5.1 of [ 62]), the fixed point iteration with step size < m=L2will converge. Note that these are
substantially weaker rates and constants than required for typical gradient descent or the minimization
of locally convex function because the coupling between the three fixed point iterations introduce
cross-terms in the joint Jacobian.
Due to these cross-terms, going from the strong monotonicity assumption on the joint fixed point
iterations to specific assumptions on fand`is less straightforward. However, empirically, we
observed that as long as the step sizes ’s were kept reasonably small and the functions fand`were
designed appropriately (e.g `respecting notions of local convexity and fwith Jacobian eigenvalues
less than 1, etc.) the fixed point iterations converged reliably.
3.3 Iterpretation as a primal-dual optimization
While the problem above was introduced as an input-optimization problem, its formulation as a
joint optimization problem in the augmented DEQ system (10) can also be viewed as a constrained
5optimization problem where the DEQ fixed-point conditions are treated as constraints,
minimize
x;z`(z;y)subject toz=f(z;x) (12)
which yield the Lagrangian
minimize
x;zmaximize
L(x;z; )`(z;y) +>(f(z;x)z) (13)
and the corresponding KKT conditions2
4f(z;x)z
@L(x;z; )
@z@L(x;z; )
@x3
5=2
4f(z;x)z
@`(z;y)
@z+>(@f(z;x)
@zI)
>@f(z;x)
@x3
5="0
0
0#
2R2n+d(14)
whereare the dual variables corresponding to the equality constraints. Rearranging the terms in
the KKT conditions of the above problem, introducing the step size parameters 0sand treating it as
fixed point iteration gives us the updates in Eq. (11). Indeed, performing such iterations is a variation
of the classical primal-dual gradient method for solving equality-constrained optimization problems
[20, 34, 17].
3.4 Outer Optimization (Backward Pass)
A notable advantage of formulating the entire joint inference and input optimization problem as an
augmented DEQ v?
(y)is that it allows us to abstract away the detailed function of v, and simply
train parameters of this joint process as an outer optimization problem:
minimize
`outer(v?
(y);y) (15)
where`outer:XZZY! R+
Given the solutions x?;z?, to the inner problem, computing updates to in the outer optimization
problem is equivalent to the backward pass of the augmented DEQ and correspondingly is optimized
using standard stochastic gradient optimizers like Adam [ 42]. Thus, as with any other DEQ model,
we assume the inner problem was solved to a fixed point, and apply the implicit function theorem to
compute the gradients w.r.t the augmented system (11). This gives us a constant memory backward
pass which is invariant to the underlying optimizer used to solve the inner problem.
@`outer(v?
;y)
@=@`outer(y;v?)
@v?@v?

@=@`outer(y;v?)
@v?@K(v?)
@v?1@K(v?)
@(16)
wherev= [x;z; ]>andK(v) = 0 represents the KKT conditions from (14). As with the original
DEQ, instead of computing the above expression explicitly, we first solve the following linear system
to computeuand then substitute it back in the equation above to obtain the full gradient,
u>=@`outer(y;v?)
@v?@K(v)
@v1
()@`outer(y;v?)
@v?>
+@K(v)
@v>
u= 0 (17)
Although we can train any joint DEQ in this manner, doing so in practice (e.g., via automatic
differentiation), will require double backpropagation, because the definition of K(v)above already
includes vector-Jacobian products, and this expression will require differentiating again. However,
in the case that `outeris the same as`(or in fact where it is the negation of `), then there exists a
substantial simplification of the outer optimization gradient. These cases are indeed quite common, as
we may want e.g., to train the parameters of a generative model to minimize the same reconstruction
error that we attempt to optimize via the latent variable; or in the case of adversarial examples, the
inner adversarial optimization is precisely the negation of the outer objective.
In these cases, we have that
`outer(y;v?
(y)) =`(y;z?
(y)) (18)
so we have that
@`outer(y;v?
(y))
@=@`(y;z?)
@z?
I@f(z?;x?)
@z?1@f(z?;x?)
@= (?)>@f(z?;x?)
@(19)
In other words, we can compute the exact needed derivatives with respect to by simply re-using
the converged solution v?, without the need to double backpropagate through the KKT system. The
same considerations, but just negative, apply to the case where `outer=`.
64 Experiments
As our approach provides a generic framework for joint modeling of an implicit network’s forward
dynamics and the “inner” optimization over the input space, we demonstrate its effectiveness and
generalizability on 4 different types of problems that are popular areas of research in machine learning:
generative modeling [ 43,27], inverse problems [ 10,25,31], adversarial training [ 69,53] and gradient
based meta-learning[ 22,58] (results for the latter are in the appendix). In all cases, we show that our
joint inference and input optimization (JIIO) provides significant speedups over projected gradient
descent applied to DEQ models and that the models trained using JIIO achieve results competitive
with standard baselines. In all of our experiments, the design of our model layer ffollows from the
prior work on multiscale deep equilibrium (MDEQ) models [ 6] that have been applied on large-scale
computer vision tasks, and where we replace all occurrences of batch normalization [ 35] with group
normalization [ 70] in order to ensure the inner optimization can be done independently for each
instance in the batch. We elaborate on the details of the choice of other hyperparameters and design
decisions of our model (such as the damping parameters in the update step (11)), as well as that of
the datasets for each task in the Appendix.
We introduce below each problem instantiation, how they fit into our methodology described in
Sec. 3.2, and the result of applying the JIIO framework compared to the alternative methods trained
in similar settings. Overall, our results provide strong evidence of benefits of performing joint
optimizations on implicit models, thus opening new opportunities for future research in this direction.
4.1 Generative Modeling
We study the application of JIIO to learning decoder-only generative models that compute the latent
representations by directly minimizing the reconstruction loss [ 72,9]; i.e., given a decoder network
D, the latent representation xof a sampley(e.g., an image) is x= minx2XkD(x)yk2
2.2Moreover,
instead of placing explicit regularizations on the latent space X(as in V AEs), we follow [ 24] to
directly train the decoder for reconstruction (and then after training, we fit the resulting latents using
a simple density model, post-hoc, for sampling). Formally, given sample data y1;:::;yn(e.g.,n
images), the generative model we study takes the following form:
minimize
nX
i=1kyih(z?
i)k2
subject tox?
i;z?
i= argmin
x;z:z=f(z;x)kyih(z)k2; i= 1;:::;n(20)
wherehis a final output layer that transform the activations z?to the target dimensionality. We
train the MDEQ-based fwith the JIIO framework on standard 64 64 cropped images from CelebA
dataset, which consists of 202,599 images. We use the standard train-val-test split as used in Liu
et al. [51] and train the model for 50k training steps. We use Fréchet inception distance (FID) [ 33] to
measure the quality of the sampling and test-time reconstruction of the implicit model trained with
JIIO and compare with the other standard baselines such as V AEs [ 43]. The results are shown in
Table 1. JIIO-MDEQ refers to the MDEQ model trained using our setup with 40 JIIO iterations in
the inner loop during training (and tested with 100 iterations). MDEQ-V AE refers to an equivalent
MDEQ model but with an encoder and a decoder trained as a V AE. We observe that our model’s
generation quality is competitive with, or better than, each of these encoder-decoder based approaches.
Moreover, with the joint optimization proposed, JIIO-MDEQ achieves the best reconstruction quality.
We additionally apply JIIO on pre-trained MDEQ-V AEs (i.e., train an MDEQ-based V AE as usual
on optimizing ELBO [ 43], and take the decoder out) for test-time image reconstruction. The result
(shown in Table 1) suggests that the reconstructions obtained as a result are better even than the
original MDEQ-V AE. In other words, JIIO can be used with general implicit-mode-based decoders at
test time even if the decoder wasn’t trained with JIIO.
2This notation differs from the “standard” notation of latent variable models (where the latent variable is
typically denoted by z). However, because x;y;z all have standard meanings in setting above, we change from
the common notation here to be more consistent with the remainder of this paper.
7Figure 2: Samples generated with
JIIO on a small MDEQ network.Model Generation Reconstruction
V AE [43] 48.12 39.12
RAE [24] 40.96 36.01
MDEQ-V AE 57.15 45.81
MDEQ-V AE (w/ JIIO) - 42.36
JIIO-MDEQ 46.82 32.52
Table 1: Comparison of FID scores attained by standard
generative models with our method, which performs joint
optimization. We use 40 solver iterations (for the augmented
DEQ) to train the JIIO model reported in this table.
One of the key advantages presented by JIIO is the relative speed of optimization over simply running
gradient descent (or its adaptive variants like Adam [ 41]). Table 2 shows our timing results for
one optimization run on a single example for various models (averaged over 200 examples). We
observe that performing 40 iterations of projected Adam takes more than 9the time taken by
40 iterations of JIIO, which we used during training and more than 3:5the time taken by 100
iterations of JIIO which we use for reconstructions at test time (e.g., to produce the results in Table 1,
though both of them lead to similar levels of reconstruction loss). Fig 3 shows the reconstruction
loss as it evolves across a single optimization run for an MDEQ model trained with JIIO. This again
clearly shows that JIIO converges vastly faster (in terms of wall-clock time) than if we handle the
inner optimization separately as in prior works, demonstrating the advantage of joint optimization.
However, it’s interesting to note that JIIO optimization seems somewhat unstable (see Fig. 3) and
fluctuates more as well. This seems to be an artifact of the specific acceleration scheme we use (see
more details in Appendix A.2).
4.2 Inverse Problems
We also extend the setup mentioned in section 4.1 directly to inverse problems. These problems,
specifically, can be approached as either an unsupervised or a supervised learning problem, which
we discuss separately in this section. To demonstrate how JIIO can be applied, we will be using
image inpainting and image denoising as example inverse problem tasks, which was extensively
studied in prior works like Chang et al. [13], Gilton et al. [26]. For the inpainting task, we randomly
mask a 20x20 window from the image and train the model to adequately fill the missing pixels based
on the surrounding image context. For the image denoising tasks, we add random gaussian noise
"N(0;2I)with= 0:2and= 0:4, respectively, to all pixels in the image, and train the model
to recover the original image. We use the same datasets and train-test setups as in the generative
modeling experiments in Sec. 4.1.
Figure 3: Cost changing with time for Adam v/s
JIIO optimization. Tested on models trained with
40 and 100 JIIO iterations respectivelyModel time taken (ms)
PGD : 20 iters 4360
JIIO : 80 iters 1401
Table 2: Time taken to compute adversarial
example of a MDEQ model on MNIST
Model time taken (ms)
Adam : 40 iters 7659
JIIO : 40 iters 862
JIIO : 100 iters 2156
Table 3: Time taken to perform JIIO optimization
v/s Adam in the generative modeling/inverse
problem experiments
8Task Model Inpainting Denoising (= 0:2) Denoising (= 0:4)
AE 17.9 18.72 18.32
Supervised MDEQ-AE 17.06 18.58 18.49
JIIO-MDEQ-100 16.90 18.22 17.89
V AE (Adam) [10] 15.34 15.31 15.24
Unsupervised MDEQ-V AE (Adam) 16.62 16.96 16.87
JIIO-MDEQ-40 15.88 17.08 16.03
JIIO-MDEQ-100 15.87 17.86 17.55
Table 4: Comparison of Median PSNR values for supervised and unsupervised inverse problem
solving approaches. The top 3 rows show models that are trained for the specific inverse problem and
the latter 5 show pre-trained generative models re-purposed for solving inverse problems
4.2.1 Unsupervised inverse problem solving
Bora et al. [10], Chang et al. [13] have showed that we can solve most inverse problems by taking a
pre-trained generative model and using that as a prior to solve for the likely solutions to the inverse
problems by optimizing on the input space of the generative model. Specifically, given a “generator”
networkG:X !Y , mapping from the latent space Xto an observed space Y, that models the
data generating distribution, they show that one can solve any inverse problem by optimizing the
following objective:
minimize
x2Xk^yAG(x)k2
2: (21)
where ^y=Ay2Y represents the corrupted data point, y2Y is the uncorrupted data and A:Y!Y
denotes the measurement matrix that defines the specific type of inverse problem that we try to solve
(e.g., for image inpainting, it would be a mask with the missing regions filled in with zeros. For
deblurring, it would be a convolution with a gaussian blur operator etc.). They call it unsupervised
inverse problem solving. Likewise, we can use the pre-trained generator from section 4.1 to solve
most inverse problems by simply solving a slightly modified version of the inner problem in (20):
minimize
x;z:z=f(z;x)k^yAh(z)k2(22)
In table 4, the unsupervised results for V AE and MDEQ-V AE generators are obtained by optimizing
(21) using Adam for 40 iterations, while for the JIIO trained models, we optimize (22) with 100 JIIO
iterations. JIIO-MDEQ-40 and JIIO-MDEQ-100 refer to JIIO-MDEQ models trained with 40 and
100 inner-loop iterations respectively. The results in Table 4 show that on all 3 problems, JIIO trained
generators produce results comparable to the V AE and MDEQ-V AE generators. Moreover, as shown
in section 4.1, JIIO also converges much faster than Adam applied to a MDEQ-V AE generator.
4.2.2 Supervised Inverse problem solving
While the unsupervised inverse problem solving works reasonably well, we can also learn models to
solve specific inverse problems to obtain better performance. Specifically, given uncorrupted data
y1;:::;yn, and the measurement matrix A, we can train a network G:Y!Y mapping from the
corrupted sample ^yi=Ayito the uncorrupted sample yiby minimizing:
minimize
nX
i=1kyiG(Ayi)k2(23)
Now, instead of modeling Gas an explicit network, we could also model it as a solution to the
inverse problem in (22) and the resulting parameters can be trained as follows:
minimize
nX
i=1kyih(z?
i)k2
subject tox?
i;z?
i= argmin
x;z:z=f(z;x)kAyiAh(z)k2; i= 1;:::;n(24)
As shown in Table 4 this yields models competitive with their autoencoder based counterparts, while
being better than all the unsupervised approaches. Each of the baseline models in the supervised
section of Table 4 are trained by simply optimizing (23) with the corresponding model replacing G.
However, note that given the models here are trained on specific inverse problems, one would have to
train a new model for each new problem as opposed to the unsupervised approach.
9Datasets Train (#) Test (!) Clean PGD JIIO
Clean 99.450.03 80.11.87 65.884.72
MNIST PGD 99.180.03 96.530.05 95.740.04
JIIO 99.320.09 95.740.22 96.630.58
Clean 78.470.94 2.380.41 3.714.01
CIFAR PGD 54.911.01 37.40.26 36.170.55
JIIO 55.540.82 37.310.67 37.770.84
Table 5: Comparison of adversarial training approaches on L2 norm perturbations with = 1. The
rows represent the training procedure and the columns represent the testing procedure
4.3 Adversarial Training
Although the previous two tasks were based on image generation, we note that our approach can
be used more generally for input optimization in DEQs and illustrate that by applying it to `2
adversarial training on DEQ-based classification models. Specifically, given inputs xi;yi;i=
1;:::;n , adversarial training seeks to optimize the objective
minimize
nX
i=1max
kk2`(h(xi+i);yi) (25)
We apply our setting to a DEQ-based classifier h(x) =h(z?
(x))wherez?=f(z?;x). In
this setting, we embed the iterative optimization over (with projected gradient descent) into the
augmented DEQ, and write the problem as
minimize
nX
i=1`(h(z?
i);y)
subject toz?
i;?
i= argmin
z;:z=f(z;x+);kk2`(h(z);y); i= 1;:::;n(26)
Specifically we train MDEQ models on CIFAR10 [ 48] and MNIST [ 50] using adversarial training
against L2 attacks with = 1for 20 epochs and 10 epochs respectively using the standard train-val-
test splits. Table 5 shows the robust and clean accuracy of models trained using PGD adversarial
training, JIIO adversarial training and vanilla training. We find that models trained using JIIO have
comparable robust and clean accuracy on both datasets. Furthermore, when tested on models trained
without adversarial training, we observe that JIIO serves as a comparable attack method to PGD.
Table 4.2 shows the time taken to find the adversarial example for a single image of MNIST using 20
iterations of PGD and 80 iterations of JIIO. We again observe more than 3x speedups when using JIIO
over using PGD while obtaining competitive robust accuracy. However, note that, unlike previous
experiments in generative modeling/inverse problems, performing adversarial training with truncated
JIIO iterations would lead to significant reduction in robust performance due to the adversarial setting.
5 Concluding remarks
We present a novel optimization procedure for jointly optimizing over the input and the forward fixed
point in DEQ models and show that, for the same class of models, it is 39faster than performing
vanilla gradient descent or Adam on the inputs. We also apply this approach to 3 different settings to
show it’s effectiveness: training generative models, solving inverse problems, adversarial training of
DEQs and gradient based meta-learning. In the process, we also introduce an entirely new type of
decoder only generative model that performs competitively with it’s autoencoder based counterparts.
Despite these features, we note that there is substantial room for future work in these directions.
Notably, despite the fact that the augmented joint inference and input optimization DEQ can embed
both processes in a “single” DEQ model, in practice these joint models take substantially more
iterations to converge as well (often in the range of 50-100) than traditional DEQs (often in 10-20
iterations), and correspondingly often use larger memory caches within methods like Anderson
acceleration. Thus, while we make the statement that these augmented DEQs are “just” another
DEQ, this relative difficulty in finding a fixed point likely adds challenges to training the underlying
models. Thus, despite ongoing work in improving the inference time of typical DEQ models, there is
substantial room for improvement here in making these joint models truly efficient.
106 Acknowledgements
Swaminathan Gurumurthy and Shaojie Bai are supported by a grant from the Bosch Center for
Artificial Intelligence. We further thank Zhichun Huang and Taylor Howell for their help with some
implementations and for various brainstorming sessions and feedback throughout the course of this
project. We would also like to thank Vladlen Koltun for brainstorming ideas especially in the initial
stages of the project.
References
[1]B. Amos and J. Z. Kolter. Optnet: Differentiable optimization as a layer in neural networks. In
International Conference on Machine Learning (ICML) , pages 136–145, 2017.
[2]B. Amos, L. Xu, and J. Z. Kolter. Input convex neural networks. In D. Precup and Y . W. Teh,
editors, Proceedings of the 34th International Conference on Machine Learning , volume 70 of
Proceedings of Machine Learning Research , pages 146–155. PMLR, 06–11 Aug 2017. URL
http://proceedings.mlr.press/v70/amos17b.html .
[3]B. Amos, I. Jimenez, J. Sacks, B. Boots, and J. Z. Kolter. Differentiable mpc for end-to-end
planning and control. In S. Bengio, H. Wallach, H. Larochelle, K. Grauman, N. Cesa-Bianchi,
and R. Garnett, editors, Advances in Neural Information Processing Systems , volume 31. Curran
Associates, Inc., 2018. URL https://proceedings.neurips.cc/paper/2018/
file/ba6d843eb4251a4526ce65d1807a9309-Paper.pdf .
[4]D. G. Anderson. Iterative procedures for nonlinear integral equations. Journal of the ACM
(JACM) , 12(4):547–560, 1965.
[5]S. Bai, J. Z. Kolter, and V . Koltun. Deep equilibrium models. Advances in Neural Information
Processing Systems , 32:690–701, 2019.
[6]S. Bai, V . Koltun, and J. Z. Kolter. Multiscale deep equilibrium models. In Advances in
Neural Information Processing Systems , volume 33, 2020. URL https://github.com/
locuslab/mdeq .
[7]S. Bai, V . Koltun, and Z. Kolter. Stabilizing equilibrium models by jacobian regularization. In
M. Meila and T. Zhang, editors, Proceedings of the 38th International Conference on Machine
Learning , volume 139 of Proceedings of Machine Learning Research , pages 554–565. PMLR,
18–24 Jul 2021. URL https://proceedings.mlr.press/v139/bai21b.html .
[8]K. Binmore and J. Davies. Calculus: concepts and methods . Cambridge University Press, 2001.
[9]P. Bojanowski, A. Joulin, D. Lopez-Paz, and A. Szlam. Optimizing the latent space of generative
networks. arXiv preprint arXiv:1707.05776 , 2017.
[10] A. Bora, A. Jalal, E. Price, and A. G. Dimakis. Compressed sensing using generative models.
InInternational Conference on Machine Learning (ICML) , pages 537–546, 2017.
[11] C. G. Broyden. A class of methods for solving nonlinear simultaneous equations. Mathematics
of Computation , 1965.
[12] F. Bünning, A. Schalbetter, A. Aboudonia, M. H. de Badyn, P. Heer, and J. Lygeros. Input
convex neural networks for building mpc. ArXiv , abs/2011.13227, 2020.
[13] J. R. Chang, C.-L. Li, B. Póczos, B. Vijaya Kumar, and A. C. Sankaranarayanan. One network
to solve them all — solving linear inverse problems using deep projection models. In 2017
IEEE International Conference on Computer Vision (ICCV) , pages 5889–5898, 2017. doi:
10.1109/ICCV .2017.627.
[14] R. T. Q. Chen, Y . Rubanova, J. Bettencourt, and D. K. Duvenaud. Neural ordinary differential
equations. In S. Bengio, H. Wallach, H. Larochelle, K. Grauman, N. Cesa-Bianchi, and
R. Garnett, editors, Advances in Neural Information Processing Systems , volume 31. Curran
Associates, Inc., 2018. URL https://proceedings.neurips.cc/paper/2018/
file/69386f6bb1dfed68692a24c8686939b9-Paper.pdf .
11[15] S. Diamond, V . Sitzmann, F. Heide, and G. Wetzstein. Unrolled optimization with deep priors.
arXiv preprint arXiv:1705.08041 , 2017.
[16] J. Djolonga and A. Krause. Differentiable learning of submodular models. In
I. Guyon, U. V . Luxburg, S. Bengio, H. Wallach, R. Fergus, S. Vishwanathan, and
R. Garnett, editors, Advances in Neural Information Processing Systems , volume 30. Curran
Associates, Inc., 2017. URL https://proceedings.neurips.cc/paper/2017/
file/192fc044e74dffea144f9ac5dc9f3395-Paper.pdf .
[17] S. S. Du and W. Hu. Linear convergence of the primal-dual gradient method for convex-concave
saddle point problems without strong convexity. In The 22nd International Conference on
Artificial Intelligence and Statistics , pages 196–205. PMLR, 2019.
[18] E. Dupont, A. Doucet, and Y . W. Teh. Augmented neural odes. In H. Wallach,
H. Larochelle, A. Beygelzimer, F. d 'Alché-Buc, E. Fox, and R. Garnett, editors,
Advances in Neural Information Processing Systems , volume 32. Curran Associates,
Inc., 2019. URL https://proceedings.neurips.cc/paper/2019/file/
21be9a4bd4f81549a9d1d241981cec3c-Paper.pdf .
[19] L. El Ghaoui, F. Gu, B. Travacca, A. Askari, and A. Y . Tsai. Implicit deep learning. arXiv
preprint arXiv:1908.06315 , 2, 2019.
[20] H. C. Elman and G. H. Golub. Inexact and preconditioned uzawa algorithms for saddle point
problems. SIAM Journal on Numerical Analysis , 31(6):1645–1661, 1994.
[21] H.-r. Fang and Y . Saad. Two classes of multisecant methods for nonlinear acceleration.
Numerical Linear Algebra with Applications , 16(3):197–221, 2009.
[22] C. Finn, P. Abbeel, and S. Levine. Model-agnostic meta-learning for fast adaptation of
deep networks. In D. Precup and Y . W. Teh, editors, Proceedings of the 34th International
Conference on Machine Learning , volume 70 of Proceedings of Machine Learning Research ,
pages 1126–1135. PMLR, 06–11 Aug 2017. URL http://proceedings.mlr.press/
v70/finn17a.html .
[23] S. W. Fung, H. Heaton, Q. Li, D. McKenzie, S. Osher, and W. Yin. Fixed point networks:
Implicit depth models with Jacobian-free backprop. arXiv:2103.12803 , 2021.
[24] P. Ghosh, M. S. Sajjadi, A. Vergari, M. Black, and B. Schölkopf. From variational to
deterministic autoencoders. In International Conference on Learning Representations , 2020.
[25] D. Gilton, G. Ongie, and R. Willett. Neumann networks for linear inverse problems in imaging.
IEEE Transactions on Computational Imaging , 6:328–343, 2020. doi: 10.1109/TCI.2019.
2948732.
[26] D. Gilton, G. Ongie, and R. Willett. Deep equilibrium architectures for inverse problems in
imaging. arXiv preprint arXiv:2102.07944 , 2021.
[27] I. Goodfellow, J. Pouget-Abadie, M. Mirza, B. Xu, D. Warde-Farley, S. Ozair, A. Courville, and
Y . Bengio. Generative adversarial nets. In Z. Ghahramani, M. Welling, C. Cortes, N. Lawrence,
and K. Q. Weinberger, editors, Advances in Neural Information Processing Systems , volume 27.
Curran Associates, Inc., 2014. URL https://proceedings.neurips.cc/paper/
2014/file/5ca3e9b122f61f8f06494c97b1afccf3-Paper.pdf .
[28] I. J. Goodfellow, J. Shlens, and C. Szegedy. Explaining and harnessing adversarial examples.
arXiv preprint arXiv:1412.6572 , 2014.
[29] S. Gould, R. Hartley, and D. Campbell. Deep declarative networks: A new hope.
arXiv:1909.04866 , 2019.
[30] W. Grathwohl, R. T. Q. Chen, J. Bettencourt, and D. Duvenaud. Scalable reversible generative
models with free-form continuous dynamics. In International Conference on Learning
Representations , 2019. URL https://openreview.net/forum?id=rJxgknCcK7 .
12[31] K. Gregor and Y . LeCun. Learning fast approximations of sparse coding. In Proceedings of the
27th international conference on international conference on machine learning , pages 399–406,
2010.
[32] S. Gurumurthy, S. Kumar, and K. Sycara. Mame: Model-agnostic meta-exploration. In
Conference on Robot Learning , pages 910–922. PMLR, 2020.
[33] M. Heusel, H. Ramsauer, T. Unterthiner, B. Nessler, and S. Hochreiter. Gans trained by a two
time-scale update rule converge to a local nash equilibrium. arXiv preprint arXiv:1706.08500 ,
2017.
[34] M. Hong, M. Razaviyayn, and J. Lee. Gradient primal-dual algorithm converges to second-order
stationary solution for nonconvex distributed optimization over networks. In International
Conference on Machine Learning , pages 2009–2018. PMLR, 2018.
[35] S. Ioffe and C. Szegedy. Batch normalization: Accelerating deep network training by reducing
internal covariate shift. In International Conference on Machine Learning (ICML) , 2015.
[36] Y . Jeon, M. Lee, and J. Y . Choi. Differentiable forward and backward fixed-point iteration
layers. IEEE Access , 9:18383–18392, 2021.
[37] H. Kannan, A. Kurakin, and I. Goodfellow. Adversarial logit pairing. ArXiv , abs/1803.06373,
2018.
[38] T. Karras, S. Laine, and T. Aila. A style-based generator architecture for generative adversarial
networks, 2019 ieee. In CVF Conference on Computer Vision and Pattern Recognition (CVPR) ,
pages 4396–4405, 2018.
[39] T. Karras, S. Laine, M. Aittala, J. Hellsten, J. Lehtinen, and T. Aila. Analyzing and improving
the image quality of stylegan. In Proceedings of the IEEE/CVF Conference on Computer Vision
and Pattern Recognition , pages 8110–8119, 2020.
[40] K. Kawaguchi. On the theory of implicit deep learning: Global convergence with implicit layers.
InInternational Conference on Learning Representations (ICLR) , 2021.
[41] D. Kingma and J. Ba. Adam: A method for stochastic optimization. In International Conference
on Learning Representations (ICLR) , 2015.
[42] D. P. Kingma and J. Ba. Adam: A method for stochastic optimization. In Y . Bengio and
Y . LeCun, editors, 3rd International Conference on Learning Representations, ICLR 2015,
San Diego, CA, USA, May 7-9, 2015, Conference Track Proceedings , 2015. URL http:
//arxiv.org/abs/1412.6980 .
[43] D. P. Kingma and M. Welling. Auto-encoding variational bayes. arXiv preprint arXiv:1312.6114 ,
2013.
[44] J. Z. Kolter and G. Manek. Learning stable deep dynamics models. In NeurIPS , 2019.
[45] J. Z. Kolter and E. Wong. Provable defenses against adversarial examples via the convex outer
adversarial polytope. In ICML , 2018.
[46] J. Z. Kolter, D. Duvenaud, and M. Johnson. Deep implicit layers - neural odes, deep equilibirum
models, and beyond. 2020. URL http://implicit-layers-tutorial.org/ .
[47] S. G. Krantz and H. R. Parks. The implicit function theorem: History, theory, and applications .
Springer, 2012.
[48] A. Krizhevsky, G. Hinton, et al. Learning multiple layers of features from tiny images. 2009.
[49] B. M. Lake, R. Salakhutdinov, and J. B. Tenenbaum. Human-level concept learning through
probabilistic program induction. Science , 350(6266):1332–1338, 2015.
[50] Y . LeCun, L. Bottou, Y . Bengio, and P. Haffner. Gradient-based learning applied to document
recognition. Proceedings of the IEEE , 86(11):2278–2324, 1998.
13[51] Z. Liu, P. Luo, X. Wang, and X. Tang. Deep learning face attributes in the wild. In Proceedings
of International Conference on Computer Vision (ICCV) , December 2015.
[52] C. Lu, J. Chen, C. Li, Q. Wang, and J. Zhu. Implicit normalizing flows. In International
Conference on Learning Representations (ICLR) , 2021.
[53] A. Madry, A. Makelov, L. Schmidt, D. Tsipras, and A. Vladu. Towards deep learning models
resistant to adversarial attacks. ArXiv , abs/1706.06083, 2018.
[54] A. Makkuva, A. Taghvaei, S. Oh, and J. Lee. Optimal transport mapping via input convex
neural networks. In ICML , 2020.
[55] A. Nichol and J. Schulman. Reptile: a scalable metalearning algorithm. arXiv: Learning , 2018.
[56] G. Ongie, A. Jalal, C. A. Metzler, R. G. Baraniuk, A. G. Dimakis, and R. Willett. Deep learning
techniques for inverse problems in imaging. IEEE Journal on Selected Areas in Information
Theory , 1(1):39–56, 2020.
[57] N. Papernot, P. D. McDaniel, X. Wu, S. Jha, and A. Swami. Distillation as a defense to
adversarial perturbations against deep neural networks, 2016. In 2016 IEEE Symposium on
Security and Privacy (SP) , 2015.
[58] A. Rajeswaran, C. Finn, S. M. Kakade, and S. Levine. Meta-learning with implicit
gradients. In H. Wallach, H. Larochelle, A. Beygelzimer, F. d 'Alché-Buc, E. Fox, and
R. Garnett, editors, Advances in Neural Information Processing Systems , volume 32. Curran
Associates, Inc., 2019. URL https://proceedings.neurips.cc/paper/2019/
file/072b030ba126b2f4b2374f342be9ed44-Paper.pdf .
[59] M. Revay, R. Wang, and I. R. Manchester. Lipschitz bounded equilibrium networks.
arXiv:2010.01732 , 2020.
[60] L. F. Richardson. Ix. the approximate arithmetical solution by finite differences of physical
problems involving differential equations, with an application to the stresses in a masonry dam.
Philosophical Transactions of the Royal Society of London. Series A, Containing Papers of a
Mathematical or Physical Character , 210(459-470):307–357, 1911.
[61] Y . Rubanova, R. T. Chen, and D. Duvenaud. Latent ODEs for irregularly-sampled time series.
arXiv:1907.03907 , 2019.
[62] E. K. Ryu and S. Boyd. Primer on monotone operator methods. Appl. Comput. Math , 15(1):
3–43, 2016.
[63] T. Salimans, A. Karpathy, X. Chen, and D. P. Kingma. Pixelcnn++: Improving the pixelcnn with
discretized logistic mixture likelihood and other modifications. arXiv preprint arXiv:1701.05517 ,
2017.
[64] C. Szegedy, W. Zaremba, I. Sutskever, J. Bruna, D. Erhan, I. Goodfellow, and R. Fergus.
Intriguing properties of neural networks. 2014.
[65] G. Tao, S. Ma, Y . Liu, and X. Zhang. Attacks meet interpretability: Attribute-steered detection
of adversarial samples. In S. Bengio, H. Wallach, H. Larochelle, K. Grauman, N. Cesa-Bianchi,
and R. Garnett, editors, Advances in Neural Information Processing Systems , volume 31. Curran
Associates, Inc., 2018. URL https://proceedings.neurips.cc/paper/2018/
file/b994697479c5716eda77e8e9713e5f0f-Paper.pdf .
[66] H. F. Walker and P. Ni. Anderson acceleration for fixed-point iterations. SIAM Journal on
Numerical Analysis , 49(4):1715–1735, 2011.
[67] P.-W. Wang, P. Donti, B. Wilder, and Z. Kolter. SATNet: Bridging deep learning and logical
reasoning using a differentiable satisfiability solver. In K. Chaudhuri and R. Salakhutdinov,
editors, Proceedings of the 36th International Conference on Machine Learning , volume 97 of
Proceedings of Machine Learning Research , pages 6545–6554. PMLR, 09–15 Jun 2019. URL
http://proceedings.mlr.press/v97/wang19e.html .
14[68] E. Winston and J. Z. Kolter. Monotone operator equilibrium networks. In Neural Information
Processing Systems , 2020.
[69] E. Wong, L. Rice, and J. Z. Kolter. Fast is better than free: Revisiting adversarial training. In
International Conference on Learning Representations , 2020. URL https://openreview.
net/forum?id=BJx040EFvH .
[70] Y . Wu and K. He. Group normalization. In Proceedings of the European conference on computer
vision (ECCV) , pages 3–19, 2018.
[71] Y . Wu, J. Donahue, D. Balduzzi, K. Simonyan, and T. Lillicrap. {LOGAN}: Latent optimisation
for generative adversarial networks, 2020. URL https://openreview.net/forum?
id=rJeU_1SFvr .
[72] A. Zadeh, Y .-C. Lim, P. P. Liang, and L.-P. Morency. Variational auto-decoder. arXiv preprint
arXiv:1903.00840 , 2019.
[73] J. Zhang, B. O’Donoghue, and S. Boyd. Globally convergent type-i anderson acceleration for
nonsmooth fixed-point iterations. SIAM Journal on Optimization , 30(4):3170–3197, 2020.
[74] L. Zintgraf, K. Shiarli, V . Kurin, K. Hofmann, and S. Whiteson. Fast context adaptation via
meta-learning. In International Conference on Machine Learning , pages 7693–7702. PMLR,
2019.
15A Additional Discussions on Optimization Stability and Speed
A.1 Regularization
As was observed in previous work [ 7,68], the stability of convergence of a DEQ is directly related to
the conditioning of the Jacobian matrix at the equilibrium point. To that end, we primarily adopt the
regularization proposed by [ 7] which upper bounds the implicit model’s stability by estimating their
trace with the Hutchinson estimator: tr(Jz) =tr
@f(z;x)
@z
=E2N(0;I)[>J>
zJz]. However, our
exact implementation is subtly different from the original proposal in that we regularize the Jacobian
matrix at a randomly chosen iterate along the optimization trajectory (x(k);z(k))instead of just the
last iterate (x;z). This modification is especially important as the optimization trajectories become
long (which is the case for our problems; e.g., which could take >80 iterations), which essentially
encourages the Jacobian to be not only stable at the end but also during the root-solving process.
Specifically, with this modification, the outer objective of JIIO becomes :
minimize
`outer(v?
(y);y) +E2N(0;1)[>J>
zJz] (27)
whereis the regularization coefficient, and in practice we sample 1 or 2 ’s to produce a Monte-Carlo
estimation of the expectation term.
A.2 Choice of Acceleration method
We perform ablation experiments on the choice of the acceleration methods for performing the joint
optimization. Figure 4 shows various approaches that can be used to accelerate JIIO applied to a
reconstruction task on a pre-trained MDEQ-V AE decoder. Broyden’s method treats its solution as
a solution to a root finding problem on the KKT conditions instead of as a minimization problem
and hence, given the non-convex nature of the problem, could end up chasing arbitrary stationary
points. The Anderson based approaches treat the problem as a minimization problem and hence
are able to perform much better. Specifically, Type-I Anderson mixing is usually more unstable (a
phenomenon that had been discussed in prior work like Fang and Saad [21] and Zhang et al. [73]),
and yet manages to attain the lowest loss values in 100 JIIO iterations. Type-II Anderson, on the
other hand, allows for much smoother optimization process although it plateaus at higher loss values.
However, since speed was an important point of consideration for our experiments, we nevertheless
went with Anderson Type-I over Type-II and picked the output of the optimizer using a heuristic that
traded off between a small KKT residual and a small cost. Overall, we observed that the criterion for
this iterate selection can be somewhat flexible but preferably kept consistent between training and
testing runs when we perform JIIO.
Figure 4: Comparing different acceleration techniques on a MDEQ-V AE decoder
16B Supplementary Experiments and Results
B.1 Gradient based Meta-Learning
Gradient (or optimization) based meta-learning defines a bi-level optimization problem where the
outer loop learns meta-parameters for a distribution of tasks while the inner loop learns task-specific
parameters, typically using a small amount of data. This bi-level structure blends itself naturally
into the types of input optimization problems we have been looking at. Specifically, taking few-shot
learning as the use case, we are given a collection of tasks fTigN
i=1, each associated with a dataset
Di, from which we can sample two disjoint sets: Dtr
i=f(str
i;k;ytr
i;k)gK
k=1andDte
if(ste
i;k;yte
i;k)gK
k=1,
where each (s;y)is a data-label pair. Gradient based meta-learning for few shot learning problem
can be framed as a bi-level optimization problem with input optimization as the inner loop:
minimize
`(x?
i;h(x?
i;ste
i;k);yte
i;k); i = 1;:::;N
subject tox?
i= argmin
xi`(xi;h(xi;str
i;k);ytr
i;k); k = 1;:::;K(28)
wherex?
iare the task specific parameters inferred during the inner-loop optimization and are the
meta-parameters. Note that the task specific parameters can be treated as inputs, and thus the inner
problem becomes an input optimization problem. In this case, with a DEQ network, the above
problem can be modified slightly as:
minimize
`(x?
i;h(zi;k?(te));yte
i;k); i = 1;:::;N
subject tox?
i;z?
i;k(tr);z?
i;k(te)= argmin
xi;z(tr=te )
i;k=f(z(tr=te )
i;k;xi)`(xi;h(ztr
i;k);ytr
i;k); k = 1;:::;K
(29)
Clearly, this modified problem can now to solved using JIIO in the inner loop. Table 6 shows the
accuracies obtained by a JIIO-trained DEQ model and various baseline meta-learning approaches
like Implicit-MAML [ 58], MAML [ 22] and Reptile[ 55] on the 5-way, 1-shot task on Omniglot
[49]. We observe that the JIIO-trained DEQ model achieves comparable accuracy to the baselines.
The partially lower accuracy numbers of the DEQ model may be attributed to the fact that we do
not differentiate through the fixed point variable x?
iwhile optimizing the outer loop objective. We
observed very poor conditioning in our experiments when trying to differentiate through x?
i(thus
requiring a large number of fixed-point updates for the backward pass and resulting in poor quality
gradients) and instead hope to explore that further in future work. Table 7 shows the time taken by
JIIO v/s Adam on the DEQ model to perform optimization in the inner loop. Again, we observe that
JIIO takes more than 2.7x lesser time to converge, demonstrating the main advantage of using JIIO
for input optimization problems with DEQs.
Algorithm/Model 5-way, 1-shot
MAML 98.7
Reptile 97.68
iMAML, GD 99.16
JIIO-DEQ 97.33
Table 6: Accuracy obtained on the 5-way, 1-shot
task from the omniglot datasetModel time taken (ms)
PGD : 20 iters 28948
JIIO : 100 iters 11836
Table 7: Time taken by JIIO vs. Adam to
perform inner-loop optimization in DEQ-based
meta-learning tasks
B.2 Adversarial training
We showed the adversarial training results for L2 perturbations with = 1in section 4.3. However,
for CIFAR10, it is also common to use = 0:5with L2 perturbations. Thus, we also show the robust
and clean accuracy of models trained with PGD and JIIO with = 0:5for CIFAR10 in Table 8. We
again observe that the models trained with JIIO and PGD show similar robust and clean accuracies
showing that JIIO is as effective as PGD towards finding adversarial examples while being about 3x
faster.
17Datasets Train (#) Test (!) Clean PGD JIIO
Clean 78.470.94 4.541.33 4.852.93
CIFAR PGD 68.480.81 51.770.75 50.140.74
JIIO 67.792.33 51.390.56 51.250.66
Table 8: Comparison of adversarial training approaches on L2 norm perturbations with = 0:5. The
rows represent the training procedure and the columns represent the testing procedure
B.3 Generative Models
We provide additional samples and reconstructions from the JIIO-MDEQ model trained on the CelebA
64x64 images in Figure 5. We also tried training the JIIO-MDEQ model with minor modifications
(increasing latent dimensions to 384 and switching the downscaling factor to 4) from the original
model and training it with 100 inner-loop JIIO iterations. The reconstructions from the resulting
model can be seen in Figure 6. We observe that the reconstructions are blurry and speculate that it’s
likely due to the squared error loss used in training and the limited capacity of the latent space (which
is constrained by the size of the post-hoc density model we wish to learn). Salimans et al. [63] have
proposed more suitable losses and representation approaches for high dimensional images which we
hope to test in future work to scale up our models to large scale generative modeling problems.
(a) Generated Samples
(b) Reconstructions
Figure 5: Generated Samples and Reconstructions obtained from the JIIO-MDEQ generative model
trained in Section 4 on CelebA 64x64 dataset.
B.4 Inverse Problems
We showed results on 3 inverse problems in section 4. Figures 7, 8, 9, 10, 11, 12 show examples from
unsupervised and supervised trained JIIO-MDEQ models on the tasks. We see that the unsupervised
models approach the performance of the supervised alternatives on most tasks, however, the supervised
approaches do tend to perform significantly better on the inpainting task.
C modeling and Training Details
In this section, we discuss various modeling decisions and training/testing details for all the tasks.
C.1 Datasets
Generative modeling and inverse problems. For all the tasks in the generative modeling and
inverse problems sections, we work with the CelebA 6464dataset and use the standard train-
18Figure 6: Reconstructions obtained from JIIO-MDEQ trained on 256x256 FFHQ dataset.
Figure 7: Supervised Image Denoising with additive noise sampled from N(0;0:2): (top) Noisy
image; (bottom) Recovered Image
val-test split prescribed in the original paper [ 51] with 162770 images in the training set, 19867 in
the validation set and 19962 in the test set. We follow the procedure in [ 51] to crop and resize the
images to obtain the 6464images from the dataset. In section B.3, we also show some preliminary
reconstruction results on the FFHQ dataset with 70000 images aligned and cropped as done in [ 38]
and then resized to 256256. We additionally perform data augmentation on both datasets by
performing random horizontal flips on each example.
Adversarial training. We use the well known CIFAR10 [ 48] and MNIST [ 50] datasets for our
experiments on adversarial training. The CIFAR10 dataset consists of 60k 3232color images
equally distributed over 10 classes, 50k of which are used for training and 10k for testing. The
19Figure 8: Unsupervised Image Denoising with additive noise sampled from N(0;0:2): (top) Noisy
image; (bottom) Recovered Image
Figure 9: Supervised Image Denoising with additive noise sampled from N(0;0:4): (top) Noisy
image; (bottom) Recovered Image
Figure 10: Unsupervised Image Denoising with additive noise sampled from N(0;0:4): (top) Noisy
image; (bottom) Recovered Image
Figure 11: Supervised Image Inpaiting: (top) Incomplete image; (middle) Inpainted image; (bottom)
Reconstructed Image
Figure 12: Unsupervised Image Inpaiting: (top) Incomplete image; (middle) Inpainted image;
(bottom) Reconstructed Image
20MNIST dataset consists of 70k grayscale 2828images equally distributed over 10 classes, with
60k used for training and 10k for testing.
Gradient based meta-learning. We use the Omniglot dataset [ 49] for our experiments with
gradient based meta-learning. The Omniglot dataset contains 1623 different handwritten characters
from 50 different alphabets, where each image is 2828dimensions. We create the tasks and the
corresponding meta-training and meta-testing sets using the procedure described in [58].
C.2 Architecture
For all experiments in the paper, we use the multiscale architectures proposed for DEQs (MDEQ) in
Bai et al. [6]. In this section, we provide the specific instantiation and hyperparameters of the models
used in each of our settings.
Generative modeling and inverse problems. We use a 4 branch MDEQ-LARGE model in all our
experiments with a latent vector of size 128 dimensions mapped to the input injection for each branch
using a single fully connected layer, while the output layer h(z)merges the outputs of all the branches
to obtain the output image as done in the Segmentation Networks in [ 6]. The hyperparameters of
the network are provided in Table 9. Importantly, as mentioned in the experiments section of the
paper, we replace all instances of Batch Normalization [ 35] with Group Normalization [ 70], in order
to make the inner optimization independent for each example. Furthermore, for the MDEQ-V AE
and MDEQ-AE baselines, we use a similar architecture for the encoder as well, except that the
input injection and output layers are parameterized as in the input injection layers in the MDEQ
classification networks in [ 6]. Additionally, for the baseline V AE, RAE and AE models, we borrow
the architectures and code from [24].
For the generative modeling experiments, we additionally fit a density model on the inferred latent
vectors inferred on the training set for sampling purposes. In our experiments, we fit a V AE with 4
fully connected layers each in the encoder and the decoder with the V AE latent dimension of size
384. The hidden dimensions for all non-latent layers of the V AE are set to 2096. We additionally also
fit a 20 component GMM on the latents of the V AE given the larger latent dimension.
Adversarial training. We adopt a 2 branch MDEQ-SMALL model in all our experiments and use
the structure of a standard classification network in [ 6]. The specific hyperparameters are provided
in Table 9. As with experiments above, we replace all Batch Normalization layers with Group
Normalization.
Gradient based meta-learning. We adopt a 2 branch MDEQ-SMALL model for our experiments
and use the classification network proposed in [ 5]. As with experiments above, we replace all
BatchNorm layers with GroupNorm. We additionally feed the task vector xithrough a fully connected
layer and use it as a film layer on top of the first GroupNorm in the residual block.
C.3 Joint Inference and Input Optimization (JIIO)
As opposed to vanilla DEQ models, the joint optimization problem in the augmented DEQ requires
a larger number of iterations and the corresponding acceleration techniques require larger memory
sizes to converge than the forward inference in DEQ models. Moreover, the additional instability also
benefits from additional damping. We accelerate the fixed point iterations using (Type-I) Anderson
acceleration [ 4] in all our experiments. We run the optimization for the maximum number iterations
as detailed in the experiments and pick the iterate with the least cost for each example in the batch.
Generative modeling and inverse problems. We use a memory size of 40 and observe that
further increasing memory sizes can lead to faster convergence (at the cost of additional GPU
memory requirements). Additionally, we damp the fixed point iterations with = [ z; ; x] =
[0:8;0:6;0:01]. For experiments involving 100 inner loop iterations, we further reduce x= 0:003
after 65 iterations in order to obtain finer solutions.
21Adversarial training. We use a memory size of 20 for all tasks. For the experiments on CIFAR10
and MNIST, we use = [ z; ; ] = [0:8;0:6;0:1]and = [ z; ; ] = [0:8;0:6;0:6]
respectively. Additionally, for MNIST, we reduce = 0:2after 65 iterations. After each update, we
additionally project the iterates onto an L2 ball with radius = 1in order to ensure the perturbations
stay inside the L2 ball around the example.
Gradient based meta-learning. We use a memory size of 10 for this task and set =
[ z; ; x] = [0:8;0:6;0:04]. As with the previous experiments we reduce x= 0:01after
65 iterations in order to obtain finer solutions. Additionally, we use a task/input vector of size 400 for
the meta-learning task.
C.4 Regularization
We use the regularization coefficient = 0:1and 2 for the generative modeling and inverse problems
experiments, and with 40 and 100 JIIO iterations, respectively. We use = 0:01for the adversarial
training experiments and = 0:5for our meta-learning experiments. For all experiments, we
compute the Hutchinson estimator E2N(0;1)[>J>
zJz]using 2 samples of .
C.5 Compute and Runtime
The generative model and inverse problems experiments were trained on 4 RTX-2080 Ti GPUs.
The generative modeling experiments were run for 50k training steps. For the inverse problems
experiments, we trained models with 100 JIIO iterations for 25k training steps, taking 4.5-5 days
for each training run. Adversarial training experiments with JIIO trained models take 9-10 hours
for CIFAR10 on 3 RTX-2080 Ti GPUs while taking 5-6 hours on 2 RTX-2080 Ti GPUs for MNIST
experiments. The models trained with projected gradient descent take 20-21 hours to train on 3
RTX-2080 Ti GPUs for the CIFAR10 experiments while taking 13-14 hours to train on 2 RTX-2080
gpus for the MNIST experiments. For our meta-learning experiments, the models trained using JIIO
used 4 RTX-2080 Ti GPUs for roughly 2 days.
C.6 Space complexity of the method
As pointed out earlier, JIIO has higher memory requirements than the corresponding ADAM version
due to the higher memory sizes used in computing the fixed point. For example, in the generative
modeling/inverse problems, optimization using JIIO requires 17.45 GB of GPU memory as opposed
to vanilla Adam based optimization which simply costs 7.47 GB GPU memory for a batch of 48
images from celebA. However, for the adversarial training problems, the memory requirements were
comparable - JIIO requires 2.19 GB memory as opposed to 2.15 for projected gradient descent for a
batch with 96 images from CIFAR10.
22HyperparameterAdversarial Training Generative Model/Inverse Problems Meta-Learning
MNIST CIFAR10 CelebA64 Omniglot
Input Image Size 2828 3232 6464 2828
Batch Size 96 96 48 80
Optimizer Adam Adam Adam Adam
(Start) Learning Rate 0.001 0.001 0.001 0.001
Nesterov Momentum 0.9 0.9 0.9 0.9
Weight Decay 0 0 0 0
Number of Scales 2 2 4 2
# of Channels for Each Scale [24, 24] [24, 24] [32,64,128,256] [64, 128]
Width Expansion (in the residual block) 5 5 5 5
Normalization (# of groups) GroupNorm(8) GroupNorm(8) GroupNorm(8) GroupNorm(8)
Weight Normalization Yes Yes Yes Yes
# of Downsamplings Before Equilirbium Solver 0 0 0 0
Forward Quasi-Newton Threshold Tf 18 18 18 18
Backward Quasi-Newton Threshold Tb 20 20 20 20
Broyden’s Method Storage Size m 20 20 20 20
Anderson JIIO Memeory Storage Size M 20 20 40 10
Anderson JIIO Damping [0.8, 0.6, 0.6] [0.8, 0.6, 0.1] [0.8, 0.6, 0.01] [0.8, 0.6, 0.04]
Variational Dropout Rate 0.0 0.0 0.0 0.0
Table 9: MDEQ hyperparameters for each task
23