Skip to main content

Advertisement

Log in

Adaptation of the optimal fingerprint method for climate change detection using a well-conditioned covariance matrix estimate

  • Published:
Climate Dynamics Aims and scope Submit manuscript

Abstract

The “optimal fingerprint” method, usually used for detection and attribution studies, requires to know, or, in practice, to estimate the covariance matrix of the internal climate variability. In this work, a new adaptation of the “optimal fingerprints” method is presented. The main goal is to allow the use of a covariance matrix estimate based on an observation dataset in which the number of years used for covariance estimation is close to the number of observed time series. Our adaptation is based on the use of a regularized estimate of the covariance matrix, that is well-conditioned, and asymptotically more precise, in the sense of the mean square error. This method is shown to be more powerful than the basic “guess pattern fingerprint”, and than the classical use of a pseudo-inverted truncation of the empirical covariance matrix. The construction of the detection test is achieved by using a bootstrap technique particularly well-suited to estimate the internal climate variability in real world observations. In order to validate the efficiency of the detection algorithm with climate data, the methodology presented here is first applied with pseudo-observations derived from transient regional climate change scenarios covering the 1960–2099 period. It is then used to perform a formal detection study of climate change over France, analyzing homogenized observed temperature series from 1900 to 2006. In this case, the estimation of the covariance matrix is only based on a part of the observation dataset. This new approach allows the confirmation and extension of previous results regarding the detection of an anthropogenic climate change signal over the country.

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

Access this article

Price excludes VAT (USA)
Tax calculation will be finalised during checkout.

Instant access to the full article PDF.

Fig. 1
Fig. 2
Fig. 3
Fig. 4
Fig. 5

Similar content being viewed by others

References

  • Allen M, Stott P (2003) Estimating signal amplitudes in optimal fingerprinting, part I: theory. Clim Dyn 21:477–491. doi:10.1007/s00382-003-0313-9

    Google Scholar 

  • Allen M, Tett S (1999) Cheking for model consistency in optimal fingerprinting. Clim Dyn 15(6):419–434

    Article  Google Scholar 

  • Barnett T, Pierce D, Schnur R (2001) Detection of anthropogenic climate change in the world’s oceans. Science 292:270–274

    Article  Google Scholar 

  • Caussinus H, Mestre O (2004) Detection and correction of artificial shifts in climate series. J R Stat Soci: Series C (Appl Stat) 53(3):405–425

    Article  Google Scholar 

  • Gibelin AL, Déqué M (2003) Anthropogenic climate change over the mediterranean region simulated by a global variable resolution model. Clim Dyn 20(4):327–339

    Google Scholar 

  • Hasselmann K (1979) On the signal-to-noise problem in atmospheric response studies: meteorology of tropical oceans, pp 251–259

  • Hasselmann K (1993) Optimal fingerprints for the detection of time-dependent climate change. J Clim 6(10):1957–1971

    Article  Google Scholar 

  • Hasselmann K (1997) Multi-pattern fingerprint method for detection and attribution of climate change. Clim Dyn 13(9):601–611

    Article  Google Scholar 

  • Hegerl G, Von~Storch H, Santer B, Cubash UPD Jones P (1996) Detecting greenhouse-gas-induced climate change with an optimal fingerprint method. J Clim 9(10):2281–2306

    Google Scholar 

  • Hegerl G, Hasselmann K, Cubash U, Mitchell J, Roeckner E, Voss R, Waszkewitz J (1997) Multi-fingerprint detection and attribution analysis of greenhouse gas, greenhouse gas-plus-aerosol and solar forced climate change. Clim Dyn 13(9):613–634

    Article  Google Scholar 

  • IPCC (2001) Climate change 2001: the scientific basis. Contribution of Working Group I to the third assessment report of the Intergovernmental Panel on Climate Change. Cambridge University Press, Cambridge, UK and New York, NY, USA, 881 pp, [Houghton JT, Ding Y, Griggs DJ, Noguer M, van der Linden PJ, Dai X, Maskell K, Johnson CA (eds)]

  • Ledoit O, Wolf M (2004) A well-conditioned estimator for large-dimensional covariance matrices. J Multivar Anal 88(2):365–411

    Article  Google Scholar 

  • Mardia K, Kent J, Bibby J (1979) Multivariate analysis. Academic Press, London

    Google Scholar 

  • Planton S, Déqué M, Douville H, Mestre O (2005) Impact du réchauffement climatique sur le cycle hydrologique. CR Geosci 337:193–202

    Article  Google Scholar 

  • Spagnoli B, Planton S, Déqué M, Mestre O, Moisselin JM (2002) Detecting climate change at the regional scale: the case of france. Geophys Res Lett 29(10):90–1, 90–4

    Google Scholar 

  • Stott P (2003) Attribution of regional-scale temperature changes to anthropogenic and natural causes. Goephys Res Lett 30(14):2–1, 2–4. doi:10.1029/2003GL017324

    Google Scholar 

  • Stott P, Tett S (1998) Scale-dependent detection of climate change. J Clim 11(12):3282–3294

    Article  Google Scholar 

  • Tett S, Jones G, Stott P, Hill D, Mitchell J, Allen M, Ingram W, Johns T, Johnson C, Jones A, Roberts D, Sexton D, Woodage M (2002) Estimation of natural and anthropogenic contributions to twentieth century temperature change. J Geophys Res 107(D16):10–1 10–24. doi:10.1029/2000JD000028

    Google Scholar 

  • Zwiers F, Zhang X (2003) Toward regional scale climate detection. J Clim 16(5):793–797

    Article  Google Scholar 

Download references

Acknowledgments

This work was supported by the European Commission Programme Energy, Environment and Sustainable Development under contract GOCE 036961 (CIRCE), and by the European Marie Curie network SEAMOCS. The authors are very grateful for the careful work done by two anonymous reviewers which has resulted in real improvements of the manuscript.

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to Aurélien Ribes.

Appendices

Appendix 1

Ledoit estimate \(\hat{C}_I\)

The main concepts of Ledoit and Wolf (2004) to lead to the Ledoit regularized estimate \({\hat{C}}_I\) are here reviewed.

The first question addressed is the following: given the empirical estimate of a p × p covariance matrix, \({\hat{C}},\) deduced from a n × p sample X, can a linear combination of the form \(\tilde{C} = \gamma \hat{C} + \rho I_p\) be a more precise estimate of C than \(\hat{C}\) ? The answer is yes, in the sense of the mean square error:

$$ \hbox{E} \left(|\tilde{C}-C|_T^2 \right), $$
(21)

for the classical norm:

$$ |A|_T ^2=\frac{\hbox{Tr}(A^{\prime}A)}{{p}}. $$
(22)

When C is known, the optimal values γ o and ρ o of γ and ρ are given by Ledoit and Wolf (2004):

$$ \gamma_o=\frac{{\alpha^2}}{{\delta^2}}, $$
(23)
$$ \rho_o=\frac{{\beta^2 \nu}}{{\delta^2}}, $$
(24)

where

$$ \nu=\langle C,I_p \rangle_T=\frac{\hbox{Tr}(C)}{{p}}, $$
(25)

〈.,.〉 T being the scalar product associated with the norm |.| T , and:

$$ \alpha^2=| C - \nu I_p |_T^2, $$
(26)
$$ \beta^2=\hbox{E}\left(| \hat{C}-C |_T^2 \right), $$
(27)
$$ \delta^2=\hbox{E}\left(| \hat{C} - \nu I_p |_T^2 \right). $$
(28)

In the case when C is unknown, consistent estimators \(\hat{\gamma}\) and \(\hat{\rho}\) of those two optimal coefficients can be constructed, and they are shown to be convergent under general asymptotics hypothesis. Note that the “general asymptotics” framework, also called “large dimensional”, is larger than the classical asymptotics framework for statistics, and it especially covers the case n = p, and n,p →∞. More precisely, estimates of the coefficients ν,α,β and δ are first defined:

$$ \hat{\nu}=(\hat{C},I_p)=\frac{{\hbox{Tr}}(\hat{C})}{{p}}, $$
(29)
$$ \hat{\delta}^2=|\hat{C} - \hat{\nu} I_p |_T^2, $$
(30)
$$ \hat{\beta}^2=\hbox{min} \left(\hat{\delta}^2, \frac{{1}}{{n^2}} \sum_{i=1}^{n} | \psi_i \psi_i^{\prime} - \hat{C} |_T^2 \right), $$
(31)
$$ \hat{\alpha}^2=\hat{\delta}^2 - \hat{\beta}^2. $$
(32)

Then, the estimates of γ o and ρ o are deduced:

$$ \hat{\gamma}=\frac{{\hat{\alpha}^2}}{{\hat{\delta}^2}}, $$
(33)
$$ \hat{\rho}=\frac{{\hat{\beta}^2 \hat{\nu}}}{{\hat{\delta}^2}}. $$
(34)

These estimates lead to a new estimate of the covariance matrix, namely the Ledoit regularized estimate:

$$ \hat{C}_I=\hat{\gamma} \hat{C}+\hat{\rho} I_p. $$
(35)

Appendix 2

Bootstrap and validation implementation

4.1 Simulating \({{\mathcal{D}}}_{\hat{C}_I,g}\)

The procedure implemented to simulate the distribution \({{\mathcal{D}}}_{\hat{C}_I,g}\) is presented here.

Step 0 From the observed n × p sample X, containing the observation vectors \((\psi _{i} )_{{1 \le i \le n}},\) the estimate \(\hat{C}_I\) of C is computed.

Step 1 \(\hat{C}_I\) is used to simulate a set of N samples \((X_{j}^{*} )_{{1 \le j \le N}},\) each one of size n × p, with a Monte Carlo technique: similarly to the matrix X, the lines of a X * j matrix (equivalent to one of the ψ i ) are independent random vectors, the distribution of which is a \(N(0,\hat{C}_I).\) The X * j are independent between them.

Step 2 From X * j , an estimate \(\hat{C}_j^*\) of \(\hat{C}_I\) is computed, using the estimation method presented in Appendix 1.

Furthermore, an estimate \(\hat{f}_j^* = {(\hat{C}_j^*)}^{-1} g\) of \(\hat{f}_o\) is computed.

Step 3 For each j, the conditional distribution \({{\mathcal{D}}}_{\hat{C}_I,g}\) given X * j is known:

$$ {\mathcal{D}}_{\hat{C}_I,g} |X_j^*=N \left(0,\frac{{\hat{f}_j^{*'} \hat{C}_I \hat{f}_j^*}}{{\hat{f}_j^{*'} \hat{C}_j^* \hat{f}_j^*}} \right), $$
(36)

so that the conditional probability density function associated can be deduced.

Now, the \({{\mathcal{D}}}_{\hat{C}_I,g}\) distribution being the mixture of the conditional distributions, its density is approximated by computing the mean of the N conditional densities simulated.

4.2 Validation

The question addressed here is: in what measure the \({{\mathcal{D}}}_{\hat{C}_I,g}\) -distribution can be used instead of the \({{\mathcal{D}}}_{C,g}\) one for computing the p-value of the test?

Using the distribution \({{\mathcal{D}}}_{\hat{C}_I,g},\) we can proposed a ROF test totally defined by the sample \((\psi _{i} )_{{1 \le i \le n}}.\) This test, denoted \({{\mathbb{T}}}_{\hat{C}_I^{-1} g}\) is characterized by its rejection region \({{\mathbb{W}}}_{\hat{C}_I^{-1}g}:\)

$$ {{\mathbb{W}}}_{\hat{C}_I^{-1}g}=\left\{ (\psi_{i})_{1 \leq i \leq n+1}, \frac{{\langle \psi_{n+1}, \hat{C}_I^{-1} g \rangle}}{ {\sqrt{(\hat{C}_I^{-1} g)' \hat{C}_I (\hat{C}_I^{-1} g)}}} \geq {\mathcal{D}}_{\hat{C}_I,g}^{(\alpha)} \right\}, $$
(37)

where \({\mathcal{D}} _{\hat{C}_I,g}^{(\alpha)}\) denotes the (1−α) quantile of the distribution \({{\mathcal{D}}}_{\hat{C}_I,g},\) estimated from the \((\psi _{i} )_{{1 \le i \le n}}\) using the procedure presented in the previous paragraph. Note that we use here a specific notation \({{\mathbb{T}}}_{\hat{C}_I^{-1} g}\) , because this test couldn’t be defined by a rejection region of the form (6). So that \({{\mathbb{T}}}_{\hat{C}_I^{-1} g}\) isn’t part of the T f family.

The validation proposed here focuses on the test level. To verify that the \({{\mathbb{T}}}_{\hat{C}_I^{-1} g}\) level is nominal, it is equivalent to verify that:

$$ \hbox{P}_{H_0} ({{\mathbb{W}}}_{\hat{C}_I^{-1}g})=\alpha. $$
(38)

The validation of this equality can be made similarly with the previous procedure.

Step 0 A “true” covariance matrix, C is fixed arbitrarily.

Step 1 C is used to simulate a set of N 0 samples \( (X_{k} )_{{1 \le k \le N_{0} }} ,\) similarly to the (X * j ) of the previous paragraph: each (X k ) is of dimension n × p, and its lines are independent random vectors, the distribution of which is a N(0,C). Note that X k is equivalent to the unique set of available observations X. The regularized covariance estimate \(\hat{C}_I^{(k)}\) is computed, for each k.

Step 2 Following the previously presented simulation procedure, the distribution \({{\mathcal{D}}}_{\hat{C}_I^{(k)},g}\) is generated. The candidate threshold \({{\mathcal{D}}}_{\hat{C}_I^{(k)},g}^{(\alpha)}\) is deduced.

Step 3 The conditional distribution \({{\mathcal{D}}}_{C,g}\) given X k being known, the level and the power conditionally to X k can be deduced. Finally, the unconditioned level and power are approximated, using a relatively large number of samples N 0.

As mentioned in Sect. 3.3, this validation procedure has been successfully applied to different starting matrices C.

Appendix 3

Discussing the sensitivity to the choice of q

We first study the tests \(T_{C_q^+g}\) . We denote respectively by λ 2 i , g i and ψ i , the eigenvalues of C, and the coordinates of g and ψ in the C eigenvectors basis. The detection variable associated with \(T_{C_q^+g}\) can be written:

$$ d_{C_q^+ g} = \langle \psi,C_q^+ g \rangle=\sum_{i=1}^{q} \frac{{\psi_i g_i}}{{\lambda_i^2}}. $$
(39)

Note that \(\hbox{Var} (\frac{{\psi_1}}{{\lambda_1}},\ldots,\frac{{\psi_p}}{{\lambda_p}})=I_p,\) that is to say that the \(z_i=\frac{{\psi_i}}{{\lambda_i}}\) are random variables, independents and identically distributed, and, under the hypothesis H 0, with N(0,1) distribution. So,

$$ d_{C_q^+ g}=\sum_{i=1}^{q} z_i \frac{{g_i}}{{\lambda_i}}, $$
(40)

is a random walk, with weights \(\frac{{g_i}}{{\lambda_i}}\) .

This formula highlights the role of the family \((\frac{{g_i}}{ {\lambda_i}})\) , and shows that the results are generally “unstable” with respect to the choice of q, even when q is large. Indeed, some stability occurs only when the coefficients \((\frac{{g_i}}{{\lambda_i}})\) become “small”, as q goes to p. In such a case, the tests \(T_{C_q^+g}\) tend to \(T_{C^{-1} g}\) (and they are close to each other), so the choice of q doesn’t matter, as soon as q is large.

However, in the case of climate study, there isn’t any evidence that this condition on the \((\frac{{g_i}}{{\lambda_i}})\) family is satisfied. Consequently, under the hypothesis H 0, the probability to reject H 0 for at least one of the considered tests can become important (clearly greater than the individual level α of each test). This illustrate the potential danger of the method and of the choice of q.

A similar study can be made for the tests \(T_{\hat{C}^+_q g}.\) Denoting respectively by \(\hat{\lambda}_i^2,\) \(\hat{g}_i\) and \(\hat{\psi}_i\) the eigenvalues of \(\hat{C},\) and the coordinates of g and ψ in the \(\hat{C}\) eigenvectors basis,

$$ \hat{d}_{\hat{C}_q^+ g}=\langle \psi,\hat{C}_q^+ g \rangle=\sum_{i=1}^{q} \frac{{\hat{\psi}_i}}{{\hat{\lambda}_i}} \frac{{\hat{g}_i}}{{\hat{\lambda}_i}}. $$
(41)

Here, the main differences with the previous case is that, first, the random variables \(\frac{{\hat{\psi}_i}}{{\hat{\lambda}_i}}\) are not exactly independent and identically distributed, and, second, the \(\hat{\lambda}_i\) family decreases more rapidly than the λ i one, due to estimation errors. Consequently to this last point, the sensitivity to the choice of q is increased.

Rights and permissions

Reprints and permissions

About this article

Cite this article

Ribes, A., Azaïs, JM. & Planton, S. Adaptation of the optimal fingerprint method for climate change detection using a well-conditioned covariance matrix estimate. Clim Dyn 33, 707–722 (2009). https://doi.org/10.1007/s00382-009-0561-4

Download citation

  • Received:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1007/s00382-009-0561-4

Keywords

Navigation