Welcome to the Atmospheric Group

 

Changes in atmospheric composition are important in the context of stratospheric ozone depletion, global change and related environmental problems. The Michelson Interferometer for Passive Atmospheric Sounding (MIPAS), which is a core instrument of the ENVISAT polar platform successfully launched on 1st March 2002 by the European Space Agency (ESA), is a powerful tool to measure vertical profiles of trace species on a global scale. While operational data processing by ESA covers only analysis of pressure, temperature, and the mixing ratios of the species H2O, O3, HNO3, CH4, N2O and NO2, MIPAS infrared spectral limb emission measurements contain information on a bulk of further species relevant to environmental problems mentioned above. The goal of AMIL2DA was to generate data analysis tools for these supplemental species along with thorough validation of these algorithms.

Instead of merging the contributions of all participants to one data analysis algorithm which fits all purposes, the AMIL2DA strategy was to maintain the diversity of different computer codes of each group which are custom-tailored to their specific scientific needs.

Generally, data analysis consists of forward modeling of radiance spectra and inversion of measurement data. As a first step, forward radiative transfer algorithms and retrieval processors were adapted to the physical and computational needs of the MIPAS experiment. This includes adaptation to high resolution limb emission measurements, acceleration of numerical methods, and automated provision of input data as well as generation of spectroscopic line data not included in the current databases but presumably important to MIPAS applications.

In a second step, these codes were cross-validated by a blind-test intercomparison to reveal potential weaknesses of models. In particular the relevance of breakdown of thermodynamic equilibrium in the atmosphere was emphasized. After successful cross-validation of forward radiative transfer models and subsequent upgrading, these were operated in the context of an inversion computer code, which inferred atmospheric constituent abundances from measured spectra.

For purpose of cross-validation, different inversion algorithms were applied to a common set of synthetic measurement data in a blind-test mode. After upgrading the inversion models and fine-tuning of processing parameters, a common agreed set of real MIPAS measurements was used for further testing. Residuals between measured and best-fitting modeled spectra were analyzed for systematic patterns. Emphasis was put on candidate explanations such as inappropriate predictions on instrument characteristics; different use of initial guess and a priori data; over or under-regularization of the retrieval, and possible deficiencies in spectroscopic data.

These activities are the basis to better exploit MIPAS data by inferring vertical profiles of species relevant to ozone destruction and global change. Deficiencies in forward radiative transfer as well as inversion algorithms were detected and removed, and confidence in retrieval strategies and data products was strengthened. Completeness and appropriateness of physical effects included in the involved radiative transfer models were proved. Standardization of data products was gained while the diversity of data analysis strategies used by different European groups was maintained.

 

 

 

 

 

The measurement space solution (MSS) is a new type of representation of the information retrieved about the vertical profile of an atmospheric constituent. In this new representation the profile is not given, as in classical retrievals, by a sequence of values as a function of altitude, but as the combination of a set of functions each weighted with a measured amplitude. The set of functions are those that belong to the functional space in which the measurement is performed (the so called "measurement space"). The profile obtained in this new way does not directly provide, as classical retrievals do, a useful graphical representation (as seen in Fig. 1.1 and discussed in section 5, some "maquillage" is needed for a graphical representation), but has other important advantages that make it a tool for the full extraction of the information that the observations contain about the profile that is being retrieved.

 

Fig._1.1

Fig. 1.1 Comparison between the graphical representations of the MSS and of a classical retrieval of an ozone volume mixing ratio vertical profile. The MSS does not directly provide a useful graphical representation.

 

 

 

 

 

 

 

 

 

The MSS solution provides the optimal exploitation of the information retrieved from level 2 analyses since it fully satisfies both requirements of providing the profile in a retrieval grid as fine as needed and of using no a priori information. As discussed below in a classical retrieval a compromise must be made between these two requirements.

In the retrieval of the vertical distribution of an atmospheric constituent the observations do not provide enough information to retrieve the continuous function of the vertical profile, therefore, the problem arises of how to represent, in terms of which and how many parameters, the finite number of independent pieces of information provided by the observations. The first choice is that of the vertical grid on which to represent the profile. This vertical grid should be as fine as possible in order both to make sure that all the information included in the observations is adequately represented by the profile and to limit the need for interpolation in the subsequent applications of the data. Indeed, operations of interpolation do introduce a loss of information [1]. However, whenever the information obtained from the observations is not sufficient to determine as many parameters as the grid points, a too fine grid makes the inversion problem ill-posed or ill-conditioned. This problem is generally overcome using some external information. In the optimal estimation method [2], according to the Bayesian approach, an a priori probability distribution function of the atmospheric state is used to fill the gaps present in the observations. In this case the solution is represented by the state corresponding to the maximum of the a posteriori probability distribution function, which combines the a priori information with that of the observations. An alternative method consists in applying a smoothness constraint to the profile. This can be done adding to the chi-square function minimized by the retrieval a term that increases when the profile oscillates [3-8]. In this case the information not coming from the observations is in the smoothness constraint. A further method that allows to make the problem well conditioned consists in writing the profile as a linear combination of a finite number of continuous functions (generally triangular functions, boxcars, polynomials, sine and cosine functions etc...) and to determine the coefficients of these functions fitting the observations [2]. In this case the information not coming from the observations consists in the number and in the kind of the chosen functions. In the following discussion the external information not coming from the observations is referred to as "a priori information" in all cases, even when it is applied in the form of a constraint.

A priori information makes it possible to obtain a vertical profile on a fine grid, but generates other types of problems in utilizing the data. In validation activities, when the compared profiles use different a priori information, it is necessary to assess the error due to this difference [9]. The estimation of the error component due to the use of different a priori information is a difficult task because it requires a reliable knowledge of the climatological variability of the quantity to be compared, which often is not available. Also in data assimilation, which is the combining of diverse data sampled at different times and different locations into atmospheric models, the measurements containing a priori information pose some problems [2]. Profiles in different geolocations obtained using the same a priori information are correlated among themselves as well as profiles obtained using nearby retrieved profiles as a priori. If not taken into account, these correlations can produce a bias in the products of the assimilation. Also in the case of data fusion, which consists in the synergetic use of more than one measurement of the same atmospheric state obtained with different instruments, the presence of a priori information in the original measurements can produce a bias in the final product.

The above considerations suggest that the optimal exploitation of the information retrieved from level 2 analyses requires data represented on a retrieval grid as fine as needed without any a priori information. These two requirements are not simultaneously satisfied by the products currently provided by level 2 analyses, and are met in the case of the MSS solution.

 

 

 

 

 


 

 

We represent the observations (radiances) with a vector y of m elements and the vertical profile of the unknown atmospheric parameter with a vector x of n elements corresponding to a predefined altitude grid. The relationship between the vectors x and y is:

y = F (x) + ɛ , (3.1)

 

where the function F(x) is the forward model and ɛ is the vector containing the experimental errors of the observations, characterized by a variance-covariance matrix (VCM) Sy.

We expand F(x) up to the first order around a specific value of x, identified by x0 and referred to as the linearization point, and, after some rearrangement, Eq. (3.1) becomes equal to:

 

y - F(x0) + K X0 = KX + ɛ , (3.2)

where K is the Jacobian matrix (that is the partial derivatives of F(x) with respect to the elements of x) calculated at x0. Eq. (3.2) implies that the elements of  y-F(X0) + K X0 are the scalar products between x and the rows of K plus the errors.  It follows that the knowledge of  y-F(X0) + K X0 determines the knowledge of the component of x that lies in the space generated by the rows of K. This space is referred to as measurement space and its orthogonal complement in Rn is referred to as null space.
In order to weigh the observations with their errors and avoid the complication of correlated errors it is useful to consider the quantity Sy -1/2 (which is characterized by a VCM that is the unity matrix) instead of the observations y. Multiplying both terms of Eq. (3.2) on the left by Sy -1/2 we obtain:

 

Sy -1/2 [ y - F(x0) + K X0] = Sy -1/2 Kx + Sy -1/2 ɛ , (3.3)

 

 

Since the space Rn can be split into the direct sum of the measurement space and of the null space, we can write:

X = Xa + Xb ,   (3.4)

where xa and xb, respectively, belong to the measurement space and to the null space (see Fig. 3.1).


Fig._3.1

Fig. 3.1 Decomposition of the vertical profile in the measurement-space and null-space components.

 

 

xa and xb can be expressed as:

Xa = Va , (3.5)

Xb = Wb , (3.6)

 

where V is a matrix whose columns are an orthonormal basis of the measurement space, W is a matrix whose columns are an orthonormal basis of the null space and a and b are the projections of x on these orthornormal bases:

a = VT x ,  (3.7)

b = WT x , (3.8)

 

 

where the superscript T denotes transposed matrices. The component xa is the only quantity that can be derived from the observations. In order to find xa from Eq. (3.5) we need to identify V and a. To this purpose we perform the singular value decomposition (SVD) of:

 

Sy-1/2 K = U Λ VT ,  (3.9)

 

The columns of V are an orthonormal basis of the space generated by the rows of Sy-1/2K. Since Sy-1/2 is a nonsingular matrix the space generated by the rows of Sy-1/2K coincides with the space generated by the rows of K, therefore, the columns of V are an orthonormal basis of the measurement space and, among all the possible orthonormal bases of the measurement space, it can be chosen for representing xa with Eq. (3.5).  We can now determine the components of  xa relative to this orthonormal basis. Substituting Eq. (3.9) in Eq. (3.3), multiplying both terms on the left by Λ-1UT and using Eq. (3.7), after some rearrangements, we obtain that â, i.e. the estimation of a deduced from the observations, is given by:

â = a + ɛa =  VT x + ɛa = VT x0 + Λ-1 UT Sy -1/2 ( y - F(x0) ) , (3.10)


where:

ɛa = Λ-1 UT Sy -1/2 ɛ , (3.11)

 

is the error that we make taking â as the estimation of a and is characterized by the diagonal VCM:

 

Sa = Λ-2 ,   (3.12)

 

From Eq. (3.12) we see that the components of the vector â are independent of each other and are characterized by variances given by the inverse of the squared singular values of Sy-1/2K. Therefore, components corresponding to large singular values are well determined while components corresponding to small singular values are poorly determined. The definition of xa in terms of V and â, determined respectively by Eq. (3.9) and by Eq. (3.10), is referred to as the measurement-space solution (MSS).

 

 

 


 



The possibility to split the profile into a component retrieved from the observations and a component obtained from a priori information is particularly suitable to perform data fusion among independent measurements. Indeed the usual approach to data fusion implies the merging of products retrieved from the individual measurements, thus leading to a direct transfer into the fused data of any a priori information contained in a product. However, when a component of the profile is determined by a measurement there is no need for constraining this component in the other measurement by means of a priori information. This consideration suggests that the MSS is the optimal quantity to be used as input for data fusion.

 

In this section we analyze how the MSS can be used for the fusion of two independent indirect measurements of the same profile x. The results of this section can be easily extended to an arbitrary number of independent measurements.

 

Since the predetermined grid can be chosen freely and be common to the two measurements, we can assume that the two measurement spaces are subspaces of the same space Rn. The two MSSs are characterized by the matrices V1 and V2 that identify the measurement spaces (of dimensions p1 and p2 respectively) and by the two vectors â1 and â2 (with their VCM Sa1 and Sa2). The relationships between these MSSs and the profile x are given by Eq. (3.10):

 

â1 = V1Tx + ɛa1 , (4.1)

â2 = V2Tx + ɛa2     , (4.2)

 

which can be written in the compact form:

1 ,       (4.3)

 

where the notation:

2

means the matrix (vector) obtained arranging the rows of the matrix (vector) Q below the rows of the matrix (vector) P and ɛa1 and ɛa2 contain the errors with which a1=V1Tx and a2=V2Tx are stimated by â1 and â2. Eq. (4.3) implies that the elements of the vector 3 are the scalar products between x and the columns of V1 and of V2 plus the errors. It follows that the knowledge of 3 determines the knowledge of the component of x that lies in the space generated by the columns of V1 and of V2 . This space is the union space of the measurement spaces of the two individual measurements. We are now in a situation similar to that encountered in the case of Eq. (3.2) where the vector  3 plays the role of y-F(x0)+Kx0 and the matrix  4 plays the role of K. Following the procedure described in the section "Theory of MSS calculation" we can calculate the MSS in the union space of the measurement spaces of the two individual measurements.

 

We can represent the profile x in the form of the summation of a vector of the union space of the two measurement spaces and of a vector of the orthogonal complement space to this space (which coincides with the intersection space of the two null spaces of the original measurements). The vector in the union space is the MSS of the data fusion problem and its identification implies the determination of the elements of the following relationship:

â12 = V12T x + ɛa12 , (4.4)

 

where â12 is the estimation of the components of x in the basis of the union space represented by the matrix V12, and ɛa12 is its error.

 

As in section theory we transform the vector 3 in such a way that its VCM becomes the unity matrix. This can be obtained multiplying Eq. (4.3) by the matrix 5 , obtaining:

 

6 ,  (4.5)

 

If we perform the SVD of the kernel of Eq. (4.5) we obtain:

 

7,   (4.6)

 

where U12 is a matrix of dimension (p1+p2)xp12, with p12(p1+p2), Λ12 is a nonsingular diagonal matrix of dimension p12x p12 and V12 is a matrix of dimension nx p12 whose columns are an orthonormal basis of the space generated by the rows of 8. Since Sa1-1/2 and Sa2-1/2 are nonsingular matrices, the space generated by the rows of 8 coincides with the space generated by the columns of V1 and of V2. Consequently the columns of V12 are an orthonormal basis of the union space of the two measurement spaces.

 

 

The knowledge of V12 can be used to determine the vector â12. Substituting Eq. (4.6) in Eq. (4.5) and multiplying both terms on the left by Λ12 -1U12T we obtain:

 

9 ,  (4.7)

 

where εa12 is given by:

 

10 ,   (4.8)

 

and is characterized by the VCM:

 

S a12 =  Λ12 -2 ,    (4.9)

 

â12 obtained from Eq. (4.7) and V12 obtained from Eq. (4.6) provide the MSS of x in the union space of the two measurements. Since the space characterized by V12 exclusively includes the two original measurement spaces, the MSS in the union space has been obtained by utilizing all the information provided by the two sets of observations and without any a priori information. This characteristic makes this method for data fusion optimal and equivalent to the simultaneous analysis of the two sets of observations.

 

This method of data fusion has been applied to the fusion of IASI and MIPAS ozone data to retrieve the tropospheric and stratospheric columns [11] and the vertical profile [12]. In Fig. 4.1 and example of the improvement in retrieval errors obtained using data fusion with respect to using single measurements is reported.

 

 

Fig._4.1

 

Fig. 4.1 Example of the improvement obtained using data fusion with respect to using single measurements. Percentage retrieval errors on ozone volume mixing ratio [12] for the profiles retrieved from simulated measurements using only the IASI measurement (green line), only the MIPAS measurement (blue line) and the IASI-MIPAS data fusion (black line). The standard deviation of the climatological profile (red line) is also reported. Panel (b) shows a blow up of the panel (a) in the low altitude region.