Kolmogorov–Arnold Network

Neural network architecture
Part of a series on
Machine learning
and data mining
Paradigms
  • Supervised learning
  • Unsupervised learning
  • Online learning
  • Batch learning
  • Meta-learning
  • Semi-supervised learning
  • Self-supervised learning
  • Reinforcement learning
  • Curriculum learning
  • Rule-based learning
  • Quantum machine learning
Problems
Learning with humans
Machine-learning venues
  • v
  • t
  • e

A Kolmogorov–Arnold Network (or KAN) is a neural network that is trained by learning the activation function at each node, rather than the weight on each edge as in a traditional multilayer perceptron. This is a practical application of the Kolmogorov–Arnold representation theorem.

Description

By the Kolmogorov–Arnold representation theorem, any multivariate continuous function f {\displaystyle f} can be written as

f ( x ) = f ( x 1 , , x n ) = q = 0 2 n Φ q ( p = 1 n ϕ q , p ( x p ) ) {\displaystyle f(\mathbf {x} )=f(x_{1},\ldots ,x_{n})=\sum _{q=0}^{2n}\Phi _{q}\!\left(\sum _{p=1}^{n}\phi _{q,p}(x_{p})\right)} .

with suitable values of the functions ϕ q , p : [ 0 , 1 ] R {\displaystyle \phi _{q,p}\colon [0,1]\to \mathbb {R} } and Φ q : R R {\displaystyle \Phi _{q}\colon \mathbb {R} \to \mathbb {R} } . To create a deep network, the KAN extends this to a graph of such representations:

f ( x ) = i L 1 = 1 n L 1 ϕ L 1 , i L , i L 1 ( i L 2 = 1 n L 2 ( i 0 = 1 n 0 ϕ 0 , i 1 , i 0 ( x i 0 ) ) ) {\displaystyle f(\mathbf {x} )=\sum _{i_{L-1}=1}^{n_{L-1}}\phi _{L-1,i_{L},i_{L-1}}\left(\sum _{i_{L-2}=1}^{n_{L-2}}\cdots \left(\sum _{i_{0}=1}^{n_{0}}\phi _{0,i_{1},i_{0}}(x_{i_{0}})\right)\right)}

where each ϕ l , i , j {\displaystyle \phi _{l,i,j}} is a learnable B-spline. Since this representation is differentiable, the values can be learned by any standard backpropagation technique.

For regularization, the L1 norm of the activation function is not sufficient by itself.[1] To keep the network sparse and prevent overfitting, an additional entropy term S {\displaystyle S} is introduced for each layer Φ {\displaystyle \Phi } :

S ( Φ ) = i = 1 n i n j = 1 n o u t | ϕ i , j | 1 | Φ | 1 log ( | ϕ i , j | 1 | Φ | 1 ) {\displaystyle S(\Phi )=-\sum _{i=1}^{n_{in}}\sum _{j=1}^{n_{out}}{\frac {|\phi _{i,j}|_{1}}{|\Phi |_{1}}}\log \left({\frac {|\phi _{i,j}|_{1}}{|\Phi |_{1}}}\right)}

A linear combination of these terms and the L1 norm over all layers produces an effective regularization penalty. This sparse representation helps the deep network to overcome the curse of dimensionality.[2]

Properties

The number of parameters in such a representation is O ( N 2 L G ) {\displaystyle O(N^{2}LG)} , where N {\displaystyle N} is the output dimension, L {\displaystyle L} is the depth of the network, and G {\displaystyle G} is the number of intervals over which each spline is defined. This might appear to be greater than the O ( N 2 L ) {\displaystyle O(N^{2}L)} parameters needed to train a multilayer perceptron of depth L {\displaystyle L} and output dimension N {\displaystyle N} ; however, Liu et al. argue that in scientific domains the KAN can achieve equivalent performance with fewer parameters since many natural functions can be decomposed efficiently into splines.[1]

KANs have been shown to perform well on problems from knot theory and physics (such as Anderson localization), although they have not yet been scaled to language models.[1]

History

Computing the optimal Kolmogorov–Arnold representation for a given function has been a research challenge since at least 1993.[3] Early work on applying this representation to a network used a fixed depth of two and did not appear to be a promising alternative to perceptrons for image processing.[4][5]

More recently, a deep learning algorithm was proposed for constructing these representations using Urysohn operators.[6] In 2021, the representation was successfully applied to a deeper network.[7] In 2022, ExSpliNet brought together the Kolmogorov–Arnold representation with B-splines and probabilistic trees to show good performance on the Iris flower, MNIST, and FMNIST datasets, and argued for the better interpretability of this representation compared to perceptrons.[8]

The term "Kolmogorov–Arnold Network" was introduced by Liu et al. in 2024, which generalizes the networks to arbitrary width and depth, and demonstrated strong performance on a realistic class of multivariate functions although training is inefficient.[1]

References

  1. ^ a b c d Liu, Ziming; Wang, Yixuan; Vaidya, Sachin; Ruehle, Fabian; Halverson, James; Soljačić, Marin; Hou, Thomas Y.; Tegmark, Max (2024). "KAN: Kolmogorov-Arnold Networks". arXiv:2404.19756 [cs.LG].
  2. ^ Poggio, Tomaso (2022). "How deep sparse networks avoid the curse of dimensionality: Efficiently computable functions are compositionally sparse". Center for Brains, Minds and Machines Memo. 10.
  3. ^ Lin, Ji-Nan; Unbehauen, Rolf (January 1993). "On the Realization of a Kolmogorov Network". Neural Computation. 5 (1): 18–20. doi:10.1162/neco.1993.5.1.18.
  4. ^ Köppen, Mario (2022). "On the Training of a Kolmogorov Network". Artificial Neural Networks — ICANN 2002. Lecture Notes in Computer Science. 2415: 474–479. doi:10.1007/3-540-46084-5_77. ISBN 978-3-540-44074-1.
  5. ^ Leni, Pierre-Emmanuel; Fougerolle, Yohan D.; Truchetet, Frédéric (2013). "The Kolmogorov Spline Network for Image Processing". Image Processing: Concepts, Methodologies, Tools, and Applications. IGI Global. ISBN 9781466639942.
  6. ^ Polar, Andrew; Poluektov, M. (30 December 2020). "A deep machine learning algorithm for construction of the Kolmogorov–Arnold representation". Engineering Applications of Artificial Intelligence. 2020: 104–137. arXiv:2001.04652. doi:10.1016/j.engappai.2020.104137.
  7. ^ Schmidt-Hieber, Johannes (May 2021). "The Kolmogorov–Arnold representation theorem revisited". Neural Networks. 137: 119–126. doi:10.1016/j.neunet.2021.01.020.
  8. ^ Fakhoury, Daniele; Fakhoury, Emanuele; Speleers, Hendrik (August 2022). "ExSpliNet: An interpretable and expressive spline-based neural network". Neural Networks. 152: 332–346. doi:10.1016/j.neunet.2022.04.029.