09.03.2013 Views

Contents - Max-Planck-Institut für Physik komplexer Systeme

Contents - Max-Planck-Institut für Physik komplexer Systeme

Contents - Max-Planck-Institut für Physik komplexer Systeme

SHOW MORE
SHOW LESS

Create successful ePaper yourself

Turn your PDF publications into a flip-book with our unique Google optimized e-Paper software.

where π and ¯π are probability distributions with πk =<br />

P(y = k|p), and ¯πk = P(y = k), respectively. The first<br />

term is the inherent uncertainty of y—this quantity is<br />

not affected by the forecasts. The second term quantifies<br />

reliability; note that this term is always positive unless<br />

p = π, which is the mathematical definition of reliability.<br />

The third term quantifies the information content<br />

of π. It contributes negatively to (i.e. improves) the<br />

score, unless π = ¯π, in which case π is constant. Hence,<br />

this term penalises lack of variability of π; the larger the<br />

variability of π, the better the score. As a whole, the decomposition<br />

If constructed yields credence from a to finite thenumber practiceof ofsamples, assessingthe<br />

forecast Talagrand quality diagram through is subject properto scoring randomrules. fluctuations. In<br />

-diagram, deviations from flatness due to finite<br />

Ensemble samples forecasts are taken into account. Under the assumption of<br />

Modern consistency,thenumberofcasesinwhichtheverification<br />

weather forecasts are generated using large<br />

dynamical atmospheric should models, followrunning a binomial on supercom- distribution<br />

puters. In order to initialise these simulations properly, is<br />

the current state of the atmosphere has to be known<br />

at least approximately, and subsequently projected into<br />

the state space of the model; this process is known as<br />

data assimilation. The fact that the initial condition is<br />

not known with certainty (and also that the model is incorrect)<br />

is accounted for by generating not one but several<br />

simulations with minutely perturbed initial conditions,<br />

resulting in an ensemble of forecasts.<br />

Although ensemble forecasts already provide vital information<br />

as to the inherent uncertainty, they need to<br />

be post-processed before they can be interpreted as<br />

probabilities. The interpretation of ensembles and how<br />

to generate useful forecast probabilities using ensembles<br />

is a very active area of research.<br />

Several different interpretations exist, a very common<br />

one being the following Monte–Carlo interpretation:<br />

An ensemble is a collection x1,...,xK of random variables,<br />

drawn independently from a common distribution<br />

function p, the forecast distribution. The forecast distribution<br />

p can be considered as the distribution of the<br />

ensemble members conditioned on the internal state<br />

of the forecasting scheme. The forecast distribution<br />

p however is but a mental construct and not operationally<br />

available.<br />

In this interpretation, the forecasting scheme is called<br />

reliable if the observation y along with the ensemble<br />

members x1 ...xK are independent draws from the<br />

forecast distribution p. Less formally stated, the observation<br />

behaves like just another ensemble member.<br />

A necessary consequence of reliability is that the rank<br />

of y among all ensemble members assumes the values<br />

1,...,K +1 with equal probability (namely 1/(K +1)).<br />

This means that the histogram of rank(y) should be flat,<br />

which can be statistically tested, see for example [3].<br />

In operational ensembles though, histograms are often<br />

found to be u–shaped, with the outermost ranks<br />

being too heavily populated (see Fig.1 for an example);<br />

in other words, outliers happen more often than<br />

they should in a reliable ensemble. This can have several<br />

reasons, such as insufficient spread or conditional<br />

bias. On the other hand, this means that we should be<br />

able to predict such outliers by looking at characteristic<br />

patterns in the ensemble. Indeed, as we could show,<br />

even for reliable ensemble forecasts, the spread of the<br />

actual specific ensemble is indicative of the probability<br />

that the future observation will be an outlier [10].<br />

v l<br />

0.999<br />

0.99<br />

0.9<br />

0.5<br />

0.1<br />

0.01<br />

0.001<br />

1 52<br />

Figure 1: Rank diagram for temperature forecasts in Hannover: The<br />

verification falls much too often into the first and last bins, indicating<br />

that outliers are too frequent for reliability. This ensemble features<br />

51 members. The y–axis shows Binomial probabilities, rather than<br />

actual counts.<br />

Unfortunately, rank based reliability tests are restricted<br />

to scalar predictions. In [9], a rank analysis for vector<br />

valued predictions was suggested by measuring the<br />

length of a minimum spanning tree. Thereby, a new<br />

scalar ensemble is created which however ceases to be<br />

independent, that is, the Monte–Carlo interpretation no<br />

longer applies. This puts the assumptions behind the<br />

entire rank histogram analysis into question. However,<br />

as it could be shown in [2] the rank based reliability<br />

analysis can still be applied to such ensembles, due to<br />

some inherent symmetries called exchangeability. In<br />

particular, these investigations demonstrated that the<br />

minimum spanning tree approach is mathematically<br />

sound.<br />

[1] Glenn W. Brier. Monthly Weather Review, 78(1):1–3, 1950.<br />

[2] J. Bröcker and H. Kantz. Nonlinear Processes in Geophysics,<br />

18(1):1–5, 2011.<br />

[3] Jochen Bröcker. Nonlinear Processes in Geophysics, 15(4):661–673,<br />

2008.<br />

[4] Jochen Bröcker. Quarterly Journal of the Royal Meteorological Society,<br />

135(643):1512 – 1519, 2009.<br />

[5] Jochen Bröcker, David Engster, and Ulrich Parlitz. Chaos, 19,<br />

2009.<br />

[6] Jochen Bröcker and Leonard A. Smith. Weather and Forecasting,<br />

22(2):382–388, 2007.<br />

[7] Thomas A. Brown. Technical Report RM–6299–ARPA, RAND<br />

Corporation, Santa Monica, CA, June 1970.<br />

[8] I. J. Good. Journal of the Royal Statistical Society, XIV(1):107–114,<br />

1952.<br />

[9] J.A. Hansen and L.A. Smith. Monthly Weather Review,<br />

132(6):1522–1528, 2004.<br />

[10] S. Siegert, J. Bröcker, and H. Kantz. Quarterly Journal of the Royal<br />

Meteorological Society, 2011 (submitted).<br />

2.15. Probabilistic Forecasting 71

Hooray! Your file is uploaded and ready to be published.

Saved successfully!

Ooh no, something went wrong!