• Aucun résultat trouvé

A Gradient Based Method for Fully Constrained Least-Squares Unmixing of Hyperspectral Images

N/A
N/A
Protected

Academic year: 2021

Partager "A Gradient Based Method for Fully Constrained Least-Squares Unmixing of Hyperspectral Images"

Copied!
5
0
0

Texte intégral

(1)

HAL Id: hal-01966030

https://hal.archives-ouvertes.fr/hal-01966030

Submitted on 27 Dec 2018

HAL is a multi-disciplinary open access

archive for the deposit and dissemination of

sci-entific research documents, whether they are

pub-lished or not. The documents may come from

teaching and research institutions in France or

abroad, or from public or private research centers.

L’archive ouverte pluridisciplinaire HAL, est

destinée au dépôt et à la diffusion de documents

scientifiques de niveau recherche, publiés ou non,

émanant des établissements d’enseignement et de

recherche français ou étrangers, des laboratoires

publics ou privés.

A Gradient Based Method for Fully Constrained

Least-Squares Unmixing of Hyperspectral Images

Jie Chen, Cédric Richard, Henri Lantéri, Céline Theys, Paul Honeine

To cite this version:

Jie Chen, Cédric Richard, Henri Lantéri, Céline Theys, Paul Honeine. A Gradient Based Method

for Fully Constrained Least-Squares Unmixing of Hyperspectral Images. Proc. IEEE workshop on

Statistical Signal Processing (SSP), 2011, Nice, France. pp.301-304, �10.1109/SSP.2011.5967687�.

�hal-01966030�

(2)

A GRADIENT BASED METHOD FOR FULLY CONSTRAINED LEAST-SQUARES

UNMIXING OF HYPERSPECTRAL IMAGES

Jie Chen

⋆†

, C´edric Richard

, Henri Lant´eri

, C´eline Theys

, Paul Honeine

Institut Charles Delaunay, Universit´e de Technologie de Troyes, CNRS, STMR, Troyes, France

Universit´e de Nice Sophia-Antipolis, CNRS, Observatoire de la Cˆote d’Azur, Nice, France

ABSTRACT

Linear unmixing of hyperspectral images is a popular ap-proach to determine and quantify materials in sensed im-ages. The linear unmixing problem is challenging because the abundances of materials to estimate have to satisfy non-negativity and full-additivity constraints. In this paper, we investigate an iterative algorithm that integrates these two requirements into the coefficient update process. The con-straints are satisfied at each iteration without using any extra operations such as projections. Moreover, the mean transient behavior of the weights is analyzed analytically, which has never been seen for other algorithms in hyperspectral image unmixing. Simulation results illustrate the effectiveness of the proposed algorithm and the accuracy of the model.

Index Terms— Hyperspectral imagery, linear unmixing,

estimation under constraints

1. INTRODUCTION

Hyperspectral imagery has been applied to a number of areas, including the environment, land use and agricultural monitor-ing. It can be used for more accurate and detailed information extraction than other types of remotely sensed data, as images provide rich spectral information to identify and distinguish between spectrally similar materials. This technology has re-ceived considerable attention.

In hyperspectral imagery, a pixel is a mixture of spec-tral components associated with a number of pure materials present in the scene. Although nonlinear mixture analysis has begun to find novel applications, linear spectral mixture anal-ysis is still a widely studied approach to determine and quan-tify materials in sensed images, due to its simpler physical in-terpretation. In linear unmixing models, measured pixels can be decomposed as linear combinations of constituent spectra, the so-called endmembers, weighted by corresponding frac-tions named abundances. A key aspect in determining abun-dances is how to deal with the physical constraints, which are non-negativity and full-additivity. This problem is usu-ally resolved by minimizing a cost function such as the least square criterion subject to these two constraints. In [1], the

authors derive an iterative algorithm referred to as fully con-strained least square algorithm, which adopts the sum-to-one constraint in the signature matrix and then performs the active set nonnegative least square algorithm of [2]. This method does not satisfy the full-additivity constraint exactly, and the active set method for non-negative least square includes sub-stantial computational complexity. In [3], another class of algorithm is proposed where the full additivity constraint is ensured by projecting estimates at each iteration.

In this paper, we propose a gradient descent method which does not require any extra projection onto the space of con-straints. It contains requirements for non-negativity and full-additivity into the coefficient update process. These two con-straints are always fulfilled. Moreover, we derive an analyti-cal analysis of the mean weight transient behavior of the pro-posed algorithm. To the best of our knowledge, this is unusual in the hyperspectral imaging community. Simulation results illustrate the effectiveness of the proposed algorithm and the accuracy of the model.

2. LINEAR MIXTURE MODEL

Linear spectral mixing model is largely used in hyperspectral imagery for determining and quantifying materials in sensed images. The physical assumption underlying this model is that each incident photon interacts with one Earth surface component only, and reflected spectra that are collected are not mixed with one another. LetL be the number of spectral

bands, and r= [r1, r2, . . . , rL]⊤ anL-by-1 column pixel of

an hyperspectral image. Assume that M is theL-by-R

tar-get matrix of endmembers, denoted by[m1, m2, . . . , mR],

where each column mi represents an endmember signature.

Let α = [α1, α2, . . . , αR]⊤ be aR-by-1 abundance column

vector associated with the pixel r. The linear mixing model is given by

r= M α + n (1)

where n is an additive white noise sequence. In order to es-timate the abundance vector α, we consider here the least-squares cost function

(3)

To be physically meaningful, this model is subject to two con-straints on α. The non-negativity constraint requires all the abundances to be nonnegative, that is,αi ≥ 0 for all i. The

full additivity constraint, also referred to as sum-to-one con-straint, requiresPR

i=1αi = 1. The linear unmixing problem

is cast as the following constrained optimization problem

α∗= arg min α J(α) (3) subject to αi≥ 0, i = 1, . . . , R (4) R X i=1 αi= 1. (5)

3. CONSTRAINED ALGORITHM WITH GRADIENT DESCENT METHOD

3.1. Solution with non-negativity constraints

In [4], we studied an algorithm for system identification involving only constraint (4), without the full-additivity con-straint (5). For this problem, using Lagrange multiplier method, the Karush-Kuhn-Tucker conditions at the optimum were combined into the following expression

αo

i[−∇αJ(α)]i = 0 (6)

where the minus sign is just used to make a gradient descent of criterionJ(α) apparent. Equations of the form g(u) = 0

can be solved with a fixed-point algorithm, under some condi-tions on functiong, by considering the problem u = u+g(u).

Implementing this fixed-point strategy with (6) leads us to the component-wise gradient descent algorithm

αi(k + 1) = αi(k) + µi(k) αi(k)[−∇αJ(α(k))]i (7)

whereµi(k) > 0 is a positive step size that allows to control

convergence. Suppose thatαi(k) > 0. Non-negativity of

αi(k + 1) is guaranteed if, and only if,

1 + µi(k) [−∇αJ(α(k))]i> 0 (8)

If [∇αJ(α(k))]i < 0, condition (8) is satisfied and

non-negativity constraint does not impose any restriction on the step size. Conversely, if[∇αJ(α(k))]i > 0, non-negativity

ofαi(k + 1) holds if

0 < µi(k) ≤

1 [∇J(α)]i

(9)

3.2. Algorithm for fully constrained problem

If non-negativity of the parameters is guaranteed at each iter-ation, we can make the following variable change to ensure that the sum-to-one constraint (5) is satisfied

αj=

wj

PR

ℓ=1wℓ

(10)

Withwi ≥ 0, the problem becomes unconstrained with

re-spect to constraint (5). The partial derivative of the cost func-tionJ with respect to the new variables wican be expressed

as follows ∂J ∂wi = R X j=1 ∂J ∂αj ×∂αj ∂wi (11) where ∂αj ∂wi = ∂wj ∂wi PR ℓ=1wℓ− ∂PR ℓ=1wℓ ∂wi wj  PR ℓ=1wℓ 2 = PδijR− αj ℓ=1wℓ (12)

The Kronecker symbolδijresults from the derivative∂wj/∂wi.

Replacing (12) into (11), the negative partial derivative ofJ

with respect towican now be written as

−∂J ∂wi =PR1 ℓ=1wℓ  −∂J ∂αi − R X j=1 αj  −∂J ∂αj    (13)

Let us now use the same rule as (7) for updating the non-negative entrieswi(k). The component-wise update equation

is given by wi(k + 1) = wi(k) + . . . + µPMwi(k) ℓ=1wℓ(k)  − ∂J ∂αi(k) − R X j=1 αj(k)  − ∂J ∂αj(k)    (14) It can be easily found thatPR

i=1wi(k + 1) =

PR

i=1wi(k) for

allµ. The factor (PM

ℓ=1wℓ(k))−1is thus constant and can be

absorbed intoµ. This yields wi(k + 1) = wi(k) + . . . + µwi(k)  − ∂J ∂αi(k) − R X j=1 αj(k)  − ∂J ∂αj(k)    (15) Dividing byPR i=1wi(k + 1) and PR

i=1wi(k) the left and

right sides of (15), respectively, and considering the variable change defined by (10), we obtain

αi(k + 1) = αi(k) + . . . + µαi(k)  − ∂J ∂αi(k) − R X j=1 αj(k)  − ∂J ∂αj(k)    (16)

We can verify that PR

i=1αi(k + 1) = PRi=1αi(k), which

(4)

as long as the weight vector is initialized by any vector α(0)

such thatPR

i=1αi(0) = 1.

The gradient of the least-squares criterion (2) is given by

∇αJ(α) = M⊤r− M⊤M α. (17)

Considering now the component-wise update equation (16), we get

α(k + 1) = α(k) + . . .

+ µ diag{α(k)}∇αJ(α) − 1∇αJ(α)⊤α(k)

 (18)

where1 is the all-one vector, and diag{·} a diagonal matrix

4. MEAN WEIGHT TRANSIENT BEHAVIOR

In this section, we are interested in studying the transient be-havior of the iteration governed by (18). As it is well known, it is rather challenging to study the performance of such iter-ative system. Therefore several simplifying assumption will be adopted in the analysis process. However, experiments will show that the simulated and predicted performance match al-most perfectly. Firstly, we rewrite the expression (18) as

α(k + 1) = α(k) + . . .

+µ diag{α(k)}(M⊤r− M⊤M α(k)) +µ diag{α(k)}1(M⊤r− M⊤M α(k))⊤α(k)

(19)

Instead of studying α(k) directly, we introduce the weight

error vector

v(k) = α(k) − α∗ (20)

with α∗ = arg min J(α) the solution of the unconstrained

problem. At each iteration, the residual error can be written as follows

e(k) = r − M α(k)

= n − M v(k) (21)

Subtracting α∗ from both sides of (19) and considering the

above expression, the following iterative equation with re-spect to the weight error vector v(k + 1) is obtained after

some mathematical calculation

v(k + 1) =v(k) + µp1− µp2− µp3+ µp4 (22)

where vectors p1to p4are defined by

p1= diag{v(k)}M⊤n+ diag{α∗}M⊤n p2= diag{v(k)}M⊤M v(k) + diag{α∗}M⊤M v(k) p3= α∗α∗⊤M⊤n+ α∗v(k) M⊤n+ v(k) α∗⊤M⊤n + v(k) v(k)⊤Mn p4= α∗α∗⊤M⊤M v(k) + α∗v⊤(k) M⊤M v(k) + v(k) α∗⊤MM v(k) + v(k)v(k) MM v(k)

Let us now analyze v(k) in terms of its expected value.

Con-sidering that n is zero-mean white noise, and using the inde-pendence of v(k) and n, the expected values of p1and p3are

equal to0. The expected values of p2and p4are given by

E[p2] = M⊤M diag{K(k)} + diag{α∗}M⊤ME[v(k)]

E[p4] = α∗α∗

M⊤ME[v(k)] + α∗trace{K(k) MM}

+ K(k) M M⊤α+ K(k) MME[v(k)]

where K(k) = E[v(k) v⊤(k)] is the covariance matrix of

the weight error vector. Expected valuesE[p2] and E[p4]

re-quire second-order moments defined byK(n) in order to

up-dateE[v(k)]. To simplify the analysis, we use the following

separation approximation

K(k) ≈ E[v(k)] E[v⊤(k)]

This simplification allows us a more detailed study of the mean weight behavior using analytical methods. See [5] for a thorough discussion on this approximation. Extensive simula-tion results have shown that this approximasimula-tion achieves ade-quate accuracy in modeling the mean behavior of the weights. Finally, we obtain the following update equation for the mean weight-error vector

E[v(k + 1)] = E[v(k)] + . . .

− µ M⊤M diag{E[v(k)] E[v⊤(k)]} + diag{α∗}MME[v(k)]})

+ µ(α∗α∗⊤MME[v(k)]

+ α∗trace{E[v(k)] E[v(k)] MM}

+ E[v(k)] E[v⊤(k)] M Mα

+ E[v(k)] E[v⊤(k)] M⊤ME[v(k)]

(23)

The behavior of the estimated abundance vector α(k) is then

described by α(k) = α∗+ v(k). Simulations will be used to

evaluate the model performance.

5. COMPUTER SIMULATIONS 5.1. Synthetic mixture of real spectra

The proposed learning algorithm was used for hyperspectral data unmixing, generated by linear combination of three pure materials with abundance α= [0.3 0.5 0.2]. These materials

are grass, cedar and asphalt, with spectral signatures extracted from the USGC library. These spectra consist of 2151 bands covering wavelengths ranging from 0.35 to 2.5µm. See

Fig-ure 1 (left). The data were corrupted by an additive Gaussian noise with SNR≈ 20 dB. Figure 1 (middle and right)

illus-trates the convergence behavior of the algorithm for two dif-ferent step sizes, averaged over50 runs. The method clearly

(5)

0.5 1 1.5 2 2.5 0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 wavelength (µm) reflecance asphalt grass cedar 0 20 40 60 80 100 0.2 0.25 0.3 0.35 0.4 0.45 0.5 0.55 Iterationts Estimated abundance α1 α2 α3 0 20 40 60 80 100 0.15 0.2 0.25 0.3 0.35 0.4 0.45 0.5 0.55 Iterationts Estimated abundance α1 α2 α3

Fig. 1.Spectra of the three endmembers used for synthesizing data (left). Estimated abundances as a function of the iteration number, with two different step-sizes: µ = 0.1 (middle), µ = 0.05 (right). The red dashed curves represent the model output (23). The blue curves are the result of Monte-Carlo simulations.

Fig. 2.Abundance maps estimated by the proposed algorithm (vegetation, soil and water, respectively)

model defined by equation (23) and simulated data match al-most perfectly. These results clearly show that the derived model can be used for design purposes and, in particular, to derive a condition on the step-size for convergence.

For comparison, we generated a 50-by-50 hyperspectral image with abundance vectors αijuniformly generated in the

simplex defined by the positivity and sum-to-one constraints. The SNR of this image was set to20 dB. We run the proposed

algorithm, the fully-constrained algorithm described in [1], and the projected-gradient algorithm presented in [3]. The root mean square error

RMSE= s

X

ij

kαij− ˆαijk2/N R

was used to compare these algorithms, and the following re-sults were obtained:

RMSEproposed= 0.0107

RMSE[1]= 0.0103 (withδ = 0.1)

RMSE[1]= 0.0120 (withδ = 1)

RMSE[3]= 0.0228

These results show the competitive performance of the pre-sented algorithm compared with state-of-the-art methods.

5.2. Simulation on real data

The studied image is the scene over Moffett Field (CA, USA), captured by the airborne visible infrared imaging spectrome-ter (AVIRIS). A sub-image of size50 × 50 pixels was chosen

to evaluate the proposed algorithm. This scene is mainly com-posed of water, vegetation and soil. The endmembers were extracted by the VCA algorithm withR = 3. Figure 2 shows

the estimated abundance maps. Note that several areas with dominant endmembers are clearly recovered.

6. CONCLUSION

This paper studied a new iterative algorithm for solving the problem of linear unmixing of hyperspectral images. Its tran-sient mean weight behavior was analytically analyzed. Sim-ulations conducted on synthetic data and real data have val-idated the algorithm and the proposed model. Future works will include detailed study of convergence, including conver-gence condition.

7. REFERENCES

[1] D. C. Heinz and C.-I Chang, “Fully constrained least squares linear mixture analysis for material quantification in hyperspec-tral imagery,” IEEE Trans. on Geoscience and Remote Sensing, vol. 39, pp. 529–545, 2001.

[2] C.L. Lawson and R.J. Hanson, Solving least squares problems, Society for Industrial Mathematics, 1995.

[3] C. Theys, N. Dobigeon, J.Y. Tourneret, and H. Lanteri, “Lin-ear unmixing of hyperspectral images using a scaled gradi-ent method,” in Statistical Signal Processing, 2009. SSP’09.

IEEE/SP 15th Workshop on. IEEE, 2009, pp. 729–732.

[4] J. Chen, C. Richard, P. Honeine, H. Lant´eri, and C. Theys, “Sys-tem identification under non-negativity constraints,” in Proc. of

European Conference on Signal Processing, Aalborg, Denmark,

2010.

[5] P.I. Hubscher and J.C.M. Bermudez, “An improved statistical analysis of the least mean fourth (LMF) adaptive algorithm,”

Signal Processing, IEEE Transactions on, vol. 51, no. 3, pp.

Figure

Fig. 1. Spectra of the three endmembers used for synthesizing data (left). Estimated abundances as a function of the iteration number, with two different step-sizes: µ = 0

Références

Documents relatifs

On passe en revue dans le présent digeste les conditions dans lesquelles les travaux d'hiver sont effectués au Canada; on prend en considération les problèmes qui se posent et

Appliqué au contexte des PED, Lin et Darnall (2015) 33 supposent qu’il existe au moins deux motivations qui amènent les entreprises à former des AS. La première est liée à la

In this study, we focused on these proteases in a rat model of Huntington’s disease using the mitochondrial toxin 3-nitropropionic acid (3NP). Re- sults showed that 3NP-induced death

La comparaison entre les deux courbes montre que pour le liant bitumineux, l’auto-échauffement peut expliquer avec une bonne précision la perte et la récupération

This thesis investigates different approaches to in-market distribution and logistics with emphasis on an off- patent blockbuster drug in the Spanish pharmaceutical

In the second scenario we manipulated the SOC and battery size... This result is similar to the previous scenario: the larger battery can provide more V2G services

It implements input/output (IO) routines for reading a variety of MEG/EEG file formats (http://martinos.org/mne/ stable/manual/io.html), advanced preprocessing tools such as the

To conclude, in this paper, undoped, Tb 3+ - and Tb 3+ -Yb 3+ doped SiN x layers deposited by reactive magnetron co-sputtering have been studied in order to obtain a thin DS or DC