Skip to main content

Parametrisation Independence of the Natural Gradient in Overparametrised Systems

  • Conference paper
  • First Online:
Geometric Science of Information (GSI 2021)

Abstract

In this paper we study the natural gradient method for overparametrised systems. This method is based on the natural gradient field which is invariant with respect to coordinate transformations. One calculates the natural gradient of a function on the manifold by multiplying the ordinary gradient of the function by the inverse of the Fisher Information Matrix (FIM). In overparametrised models, the FIM is degenerate and therefore one needs to use a generalised inverse. We show explicitly that using a generalised inverse, and in particular the Moore-Penrose inverse, does not affect the parametrisation independence of the natural gradient. Furthermore, we show that for singular points on the manifold the parametrisation independence is not even guaranteed for non-overparametrised models.

This is a preview of subscription content, log in via an institution to check access.

Access this chapter

Chapter
USD 29.95
Price excludes VAT (USA)
  • Available as PDF
  • Read on any device
  • Instant download
  • Own it forever
eBook
USD 84.99
Price excludes VAT (USA)
  • Available as EPUB and PDF
  • Read on any device
  • Instant download
  • Own it forever
Softcover Book
USD 109.99
Price excludes VAT (USA)
  • Compact, lightweight edition
  • Dispatched in 3 to 5 business days
  • Free shipping worldwide - see info

Tax calculation will be finalised at checkout

Purchases are for personal use only

Institutional subscriptions

Notes

  1. 1.

    Note that this becomes the FIM when g is the Fisher metric.

References

  1. Amari, S.I.: Natural gradient works efficiently in learning. Neural Comput. 10(2), 251–276 (1998)

    Article  Google Scholar 

  2. Ay, N.: On the locality of the natural gradient for learning in deep Bayesian networks. Inf. Geom. 1–49 (2020). https://doi.org/10.1007/s41884-020-00038-y

  3. Ay, N., Montúfar, G., Rauh, J.: Selection criteria for neuromanifolds of stochastic dynamics. In: Yamaguchi, Y. (ed.) Advances in Cognitive Neurodynamics (III), pp. 147–154. Springer, Dordrecht (2013). https://doi.org/10.1007/978-94-007-4792-0_20

    Chapter  Google Scholar 

  4. Bernacchia, A., Lengyel, M., Hennequin, G.: Exact natural gradient in deep linear networks and application to the nonlinear case. In: NIPS (2019)

    Google Scholar 

  5. Grosse, R., Martens, J.: A Kronecker-factored approximate fisher matrix for convolution layers. In: International Conference on Machine Learning, pp. 573–582. PMLR (2016)

    Google Scholar 

  6. Martens, J.: New insights and perspectives on the natural gradient method. arXiv preprint arXiv:1412.1193 (2014)

  7. Ollivier, Y.: Riemannian metrics for neural networks I: feedforward networks. Inf. Infer. J. IMA 4(2), 108–153 (2015)

    MathSciNet  MATH  Google Scholar 

  8. Van Hasselt, H.: Reinforcement learning in continuous state and action spaces. In: Wiering, M., van Otterlo, M. (eds.) Reinforcement Learning. ALO, vol. 12, pp. 207–251. Springer, Heidelberg (2012). https://doi.org/10.1007/978-3-642-27645-3_7

    Chapter  Google Scholar 

  9. Várady, C., Volpi, R., Malagò, L., Ay, N.: Natural wake-sleep algorithm. arXiv preprint arXiv:2008.06687 (2020)

  10. Watanabe, S.: Algebraic geometry and statistical learning theory. Cambridge University Press, Cambridge (2009)

    Book  Google Scholar 

  11. Zhang, G., Martens, J., Grosse, R.: Fast convergence of natural gradient descent for overparameterized neural networks. arXiv preprint arXiv:1905.10961 (2019)

Download references

Acknowledgements

The authors acknowledge the support of the Deutsche Forschungsgemeinschaft Priority Programme “The Active Self” (SPP 2134).

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to Jesse van Oostrum .

Editor information

Editors and Affiliations

4Appendix

4Appendix

Example 1

Let us consider the specific example where:

$$\begin{aligned} \mathcal {Z}&= \varXi = H = \mathbb {R}^2, \end{aligned}$$
(32)
$$\begin{aligned} \phi (\xi ^1, \xi ^2)&= (\xi ^1 + \xi ^2, 0)\end{aligned}$$
(33)
$$\begin{aligned} f(\eta ^1, \eta ^2)&= (2\eta ^1, \eta ^2)\end{aligned}$$
(34)
$$\begin{aligned} \mathcal {L}(x,y)&= x^2 \end{aligned}$$
(35)

Plugging this into the expressions derived above gives:

$$\begin{aligned} \tilde{\phi }(\eta ^1, \eta ^2)&= (2\eta ^1 + \eta ^2, 0)\end{aligned}$$
(36)
$$\begin{aligned} \partial _1(\xi )&= \partial _2(\xi ) = \frac{\partial {}}{\partial {x}}|_{\phi (\xi )}&F(\eta ) = \begin{bmatrix} 2 &{} 0 \\ 0 &{} 1 \end{bmatrix}\end{aligned}$$
(37)
$$\begin{aligned} \tilde{\partial }_1(\eta )&= 2\frac{\partial {}}{\partial {x}}|_{\tilde{\phi }(\eta )}&G(\xi ) = \begin{bmatrix} 1 &{} 1 \\ 1 &{} 1 \end{bmatrix}\end{aligned}$$
(38)
$$\begin{aligned} \tilde{\partial }_2(\eta )&= \frac{\partial {}}{\partial {x}}|_{\tilde{\phi }(\eta )}&\widetilde{G}(\eta ) = \begin{bmatrix} 4 &{} 2 \\ 2 &{} 1 \end{bmatrix}\end{aligned}$$
(39)
$$\begin{aligned} \nabla _{\partial (\xi )} \mathcal {L}&= (2(\xi ^1 + \xi ^2), 2(\xi ^1 + \xi ^2)) \end{aligned}$$
(40)
$$\begin{aligned} \nabla _{\tilde{\partial }(\eta )} \mathcal {L}&= (4(2\eta ^1 + \eta ^2), 2(2\eta ^1 + \eta ^2)) \end{aligned}$$
(41)

Now we fix \(\eta = (1,1)\) and \(\xi = f(\eta ) = (2,1)\). We start by computing \(y_\varXi \). From the above we know that:

$$\begin{aligned} y_\varXi = \mathop {\mathrm {arg\,min}}\limits _y\{ ||y|| : G(\xi ) y = \nabla _{\partial (\xi )} \mathcal {L}\}. \end{aligned}$$
(42)

It can be easily verified that this gives \(y_\varXi = (3,3)\). For \(y_H\) we get:

$$\begin{aligned} y_H = \mathop {\mathrm {arg\,min}}\limits _y\{ ||\left( F^T\right) ^{-1} y|| : G(\xi ) y = \nabla _{\partial (\xi )} \mathcal {L}\} \end{aligned}$$
(43)

which gives: \(y_H = (4 \frac{4}{5}, 1\frac{1}{5})\). Evidently we have \(y_\varXi \ne y_H\). Note however that when we map the difference of the two gradient vectors from \(T_{(2,1)} \varXi \) to \(T_{(3,0)}\mathcal {M}\) through \(d\phi _{(2,1)}\) we get:

$$\begin{aligned} d\phi _\xi \left( df_\eta \mathrm {grad}^H_\eta \mathcal {L}- \mathrm {grad}_\xi ^\varXi \mathcal {L}\right)&= \left( y_H - y_\varXi \right) ^i \partial _i(\xi )\end{aligned}$$
(44)
$$\begin{aligned}&= (4\frac{4}{5} - 3) \frac{\partial {}}{\partial {x}}|_{(3,0)} + (1\frac{1}{5} - 3) \frac{\partial {}}{\partial {x}}|_{(3,0)} \end{aligned}$$
(45)
$$\begin{aligned}&= 0 \end{aligned}$$
(46)

which shows that although the gradient vectors can be dependent on the parametrisation or inner product on the parameter space, when mapped to the manifold \(\mathcal {M}\) they are invariant.

Example 2

We illustrate the discussion in Sect. 2.2 with a specific example. Let us consider the following parametrisation:

$$\begin{aligned} \phi : (-1/8\pi , 5/8\pi )&\rightarrow (\mathbb {R}^2, \bar{g}) \end{aligned}$$
(47)
$$\begin{aligned} t&\mapsto (x,y) = (\sin (2t), \sin (t)) \end{aligned}$$
(48)

This gives \(\xi _1 = 0, \xi _2 = \frac{1}{2} \pi \) in the above discussion. We get the following calculation for \(\overline{\mathrm {grad}}_p^{\partial (0)} \mathcal {L}\):

$$\begin{aligned} \partial ^{(0)}&= d\phi _t \left( \frac{\partial {}}{\partial {t}}|_t\right) \end{aligned}$$
(49)
$$\begin{aligned}&= 2\cos (2t) \frac{\partial {}}{\partial {x}}|_{\phi (t)} + \cos (t)) \frac{\partial {}}{\partial {y}}|_{\phi (t)}\end{aligned}$$
(50)
$$\begin{aligned} \partial ^{(0)}&= 2 \frac{\partial {}}{\partial {x}}|_{(0,0)} + \frac{\partial {}}{\partial {y}}|_{(0,0)}\end{aligned}$$
(51)
$$\begin{aligned} G(0)&= 2^2 + 1^1 = 5 \end{aligned}$$
(52)
$$\begin{aligned} \overline{\mathrm {grad}}_p^{\partial (0)} \mathcal {L}&= \frac{1}{5} \frac{\partial {}}{\partial {t}}|_{t=0} \bigg (\mathcal {L}(\sin (2t), \sin (t))\bigg ) \left( 2 \frac{\partial {}}{\partial {x}}|_{(0,0)} + \frac{\partial {}}{\partial {y}}|_{(0,0)} \right) \end{aligned}$$
(53)

Now let:

$$\begin{aligned} f: (-1/8\pi , 5/8\pi )&\rightarrow (-1/8\pi , 5/8\pi )\end{aligned}$$
(54)
$$\begin{aligned} t&\mapsto -(t-\frac{1}{4}\pi ). \end{aligned}$$
(55)

We define the alternative parametrisation \(\tilde{\phi } = \phi \circ f\). Note that we have \(\tilde{\phi }(t) = (-\sin (2t), \sin (t))\) and thus \(\tilde{\phi }(0) = \phi (0) = (0,0)\). A similar calculation as before gives:

$$\begin{aligned} \tilde{\partial }^{(0)}&= -2 \frac{\partial {}}{\partial {x}}|_{(0,0)} + \frac{\partial {}}{\partial {y}}|_{(0,0)}\end{aligned}$$
(56)
$$\begin{aligned} \overline{\mathrm {grad}}_p^{\tilde{\partial }(0)} \mathcal {L}&= \frac{1}{5} \frac{\partial {}}{\partial {t}}|_{t=0} \bigg (\mathcal {L}(-\sin (2t), \sin (t))\bigg ) \left( -2 \frac{\partial {}}{\partial {x}}|_{(0,0)} + \frac{\partial {}}{\partial {y}}|_{(0,0)} \right) . \end{aligned}$$
(57)

Note that because \(\partial ^{(0)} \ne \tilde{\partial }^{(0)}\) (53) and (57) are not equal to each other. We can therefore conclude that in this case the expression for the gradient is parametrisation-dependent.

Rights and permissions

Reprints and permissions

Copyright information

© 2021 Springer Nature Switzerland AG

About this paper

Check for updates. Verify currency and authenticity via CrossMark

Cite this paper

van Oostrum, J., Ay, N. (2021). Parametrisation Independence of the Natural Gradient in Overparametrised Systems. In: Nielsen, F., Barbaresco, F. (eds) Geometric Science of Information. GSI 2021. Lecture Notes in Computer Science(), vol 12829. Springer, Cham. https://doi.org/10.1007/978-3-030-80209-7_78

Download citation

  • DOI: https://doi.org/10.1007/978-3-030-80209-7_78

  • Published:

  • Publisher Name: Springer, Cham

  • Print ISBN: 978-3-030-80208-0

  • Online ISBN: 978-3-030-80209-7

  • eBook Packages: Computer ScienceComputer Science (R0)

Publish with us

Policies and ethics