382
Views
0
CrossRef citations to date
0
Altmetric
Articles

Reconstruction of refractive indices from spectral measurements of monodisperse aerosols

Pages 323-361 | Received 28 Oct 2016, Accepted 17 Dec 2016, Published online: 19 Jan 2017

Abstract

For the investigation of two-component aerosols, one needs to know the refractive indices of the two aerosol components. One problem is that they depend on temperature and pressure, so one needs for their determination a robust measurement instrument such as the FASP device, which can cope with rigid environmental conditions. In this article, we show that the FASP device is capable of measuring the needed refractive indices, if monodisperse aerosols of the pure components are provided. We determine the particle radii of the monodisperse aerosols needed for this task and investigate how accurate the measurements have to be in order to retrieve refractive indices in a sufficient quality, such that they are suitable for investigations of two-component aerosols.

AMS Subject Classifications:

1. Set-up of the experiment

This paper provides an algorithm for the reconstruction of refractive indices from spectral measurements of monodisperse aerosols. The experiments we are conducting are similar to the experiments presented in [Citation1] with the difference that we are using air as surrounding medium for the aerosol particles and that temperature and pressure may approach 200C and 8 bar, respectively. For these rigid conditions, reliable databases for refractive indices do not exist up to now. These refractive index databases are needed for the measurement of particle size distributions of polydisperse aerosols using the FASP.

As outlined in [Citation2], the FASP measures light intensities Ilong(l) and Ishort(l) having passed a long and a short measurement path length Llong and Lshort, respectively. The evaluations of the FASP measurements are based on the relation(1.1) 0k(r,l)n(r)dr=e(l)withe(l)=-log(Ilong(l))-log(Ishort(l))Llong-Lshort,(1.1)

where k(r,l):=πr2Qext(mmed(l),mpart(l),r,l) is the so-called kernel function, l is the wavelength of the incident light, r is the radius of the spherical scattering particle and mmed(l) and mpart(l) are the refractive indices of the surrounding medium and the particle material depending on the wavelength l. The function Qext(mmed(l),mpart(l),r,l) is the Mie extinction efficiency from [Citation3]. The function n(r) is the size distribution of the scattering particels. The right-hand side e(l) in (Equation1.1) is denoted as the spectral extinction.

Now if n(r) is the size distribution of a monodisperse aerosol, where all particles possess the same radius rm, it is given by n(r)=nδ(r-rm), where n is the total number of particles and δ(r-rm) is a Dirac delta distribution truncated on the positive half-axis. Inserting this into (Equation1.1) gives(1.2) nπrm2Qext(mmed(l),mpart(l),rm,l)=e(l),(1.2)

hence the Mie extinction efficiency is measured directly at the radius rm.

The Mie extinction efficiency is given as an infinite series, i.e.Qext(mmed(l),mpart(l),r,l)=n=1qn(mmed(l),mpart(l),r,l).

The computation of the coefficient functions qn(mmed(l),mpart(l),r,l) will be discussed in Section 2.

It is clear that in practical computations Qext(mmed(l),mpart(l),r,l) can only be approximated by a truncated series, because only the computation of a finite number of the qn(mmed(l),mpart(l),r,l)’s is practically feasible.

We now fix a wavelength l. The complex refractive index mpart(l) for the wavelength l is reconstructed from FASP measurements of several monodisperse aerosols with particle radii r1, ..., rN. Let q(r1,l), ..., q(rN,l) denote the measured spectral extinctions e(l) corresponding to the particle radii r1, ..., rN. We assume that they are contaminated by additive Gaussian noise, i.e. q(ri,l)=qtrue(ri,l)+δi with δiN(0,si2) for i=1,,N. Furthermore, we assume that the standard deviations si can be estimated from measurements sufficiently accurately, such that we can regard them as known. We haveqtrue(ri,l)=niπri2n=1qn(mmed(l),mpart(l),ri,l),

where ni is the number of particles having the same radius ri. Then, a reconstrution of mpart(l) is obtained from the set of solutions M(l) of the nonlinear regression problem(1.3) M(l):=argminmCi=1N12sini2πri2n=1Ntrqn(mmed(l),m,ri,l)-q(ri,l)ni2.(1.3)

Note that M(l) contains in general more than one solution, especially when q(ri,l) is perturbed by measurement noise. We discuss nonlinear regression problems with truncated series expansions such as (Equation1.3) in Section 4.

For solving (Equation1.3), we use a global optimization strategy presented in Section 5 to generate reasonable candidates for start values for a local solver for a regularized version of (Equation1.3). Section 8 provides a selection method to find a unique start value out of the candidates. In order to apply a gradient-based local solver, we must know the derivatives of the Mie extinction efficiency series, which are discussed in Section 3.

2. Mie theory

We recapitulate Mie theory in an absorbing medium as presented in [Citation3]. Our first step is to introduce the complex-valued Riccati–Bessel functions ξn:CC and ψn:CC given by(2.1) ξn(z)=π2zJn+12(z)(2.1)

and(2.2) ψn(z)=π2zJn+12(z)+iπ2zYn+12(z),(2.2)

with the Bessel functions Jn+12:CC and Yn+12:CC of order n+12 of first and second kinds. We define the size parameter ρ=2πrl. Then, we set zmed:=ρ·mmed and zpart:=ρ·mpart. Here and in the following, we omit the wavelength dependence of mmed and mpart for better readability. We introduce the notation mmed=nmed+ikmed and mpart=npart+ikpart.

We introduce the so-called Mie coefficients:(2.3) an:=mpartξn˙(zmed)ξn(zpart)-mmedξn(zmed)ξn˙(zpart)mpartψn˙(zmed)ξn(zpart)-mmedψn(zmed)ξn˙(zpart)bn:=mpartξn(zmed)ξn˙(zpart)-mmedξn˙(zmed)ξn(zpart)mpartψn(zmed)ξn˙(zpart)-mmedψn˙(zmed)ξn(zpart)cn:=mpartψn(zmed)ξn˙(zmed)-mpartψn˙(zmed)ξn(zmed)mpartψn(zmed)ξn˙(zpart)-mmedψn˙(zmed)ξn(zpart)dn:=mpartψn˙(zmed)ξn(zmed)-mpartψn(zmed)ξn˙(zmed)mpartψn˙(zmed)ξn(zpart)-mmedψn(zmed)ξn˙(zpart)(2.3)

With the Mie coefficients, we can express the coefficient functions(2.4) An(ρ,mmed,mpart):=l2πmpartcn2ξn(zpart)ξn˙(zpart)¯-dn2ξn˙(zpart)ξn(zpart)¯(2.4)

and(2.5) Bn(ρ,mmed,mpart):=l2πmmedan2ψn˙(zmed)ψn(zmed)¯-bn2ψn(zmed)ψn˙(zmed)¯,(2.5)

which finally occur in the series expansion(2.6) Qext(r,l,mmed,mpart)=l2cI(r,l)n=1(2n+1)Im(An(ρ,mmed,mpart)+Bn(ρ,mmed,mpart)).(2.6)

Here the quantity I(rl) is the average incident intensity of light with wavelength l for a spherical particle with radius r and c denotes the speed of light in vacuum. The function I(rl) is given by(2.7) I(r,l)=l28π(kmed)2nmed2c(1+4πkmedrl-1e4πkmedrl),forkmed0I(r,l)=πr2nmed2c,forkmed=0.(2.7)

Obviously we cannot evaluate (Equation2.6) exactly, because we cannot compute an infinite sum due to limited processing resources. Therefore, we have to truncate this series expansion. In [Citation4], a commonly used truncation index Ntrunc is presented, which is given by(2.8) Ntrunc=M+4.05·M13+2,withM=maxρ,|ρ·mmed|,|ρ·mpart|.(2.8)

3. Derivatives of the truncated Mie efficiency series

The Bessel functions Jα(z) and Yα(z) for an arbitrary weight α fulfil the recurrence relations(3.1) ddz(zαJα(z))=zαJα-1(z)andddz(zαYα(z))=zαYα-1(z),(3.1)

cf. [Citation5]. For the Bessel functions occurring in the Riccati–Bessel functions, ξn(z) and ψn(z) follow from this with the weight α=n+12 that(3.2) ξn˙(z)=π2zJn-12(z)-nzJn+12(z)(3.2) (3.3) andψn˙(z)=π2zJn-12(z)-nzJn+12(z)+π2zYn-12(z)-nzYn+12(z)i(3.3)

for z0.

We apply (Equation3.1) a second time to get ξn¨(z), which yieldsξn¨(z)=π2zz2n(n+1)Jn+12(z)+(1-2n)Jn-12(z)+z2Jn-32(z).

For an arbitrary weight α, we have the recurrence relationJα-1(z)=2αzJα(z)-Jα+1(z),

see [Citation5], and we use it to eliminate the term Jn-32(z) in the expression for ξn¨(z). Then also Jn-12(z) cancels out, such that we obtain the representation(3.4) ξn¨(z)=π2zz2n(n+1)-z2Jn+12(z)(3.4)

only involving Jn+12(z).

We recapitulate the Cauchy–Riemann equations in its complex form. For a holomorphic function f:CC with f(z)=f(x+iy)=u(x,y)+iv(x,y) holds(3.5) f˙(z)=ddzf(z)=xf(x+iy)=-iyf(x+iy).(3.5)

From this follows(3.6) ux=Re(f˙(z)),uy=-Im(f˙(z)),vx=Im(f˙(z))andvy=Re(f˙(z)).(3.6)

Using the latter relations, we can deduce using zpart=ρnpart+ikpartnpartRe(ξn(zpart))=ρRe(ξn˙(zpart)),kpartRe(ξn(zpart))=-ρIm(ξn˙(zpart)),npartIm(ξn(zpart))=ρIm(ξn˙(zpart)),kpartIm(ξn(zpart))=ρRe(ξn˙(zpart))

and analogouslynpartRe(ξn˙(zpart))=ρRe(ξn¨(zpart)),kpartRe(ξn˙(zpart))=-ρIm(ξn¨(zpart)),npartIm(ξ˙n(zpart))=ρIm(ξn¨(zpart)),kpartIm(ξ˙n(zpart))=ρRe(ξn¨(zpart)).

The Bessel function valuesJ0+12(zmed),,JNtrunc+12(zmed),Y0+12(zmed),,YNtrunc+12(zmed)andJ0+12(zpart),,JNtrunc+12(zpart).

already computed for a function evaluation of the truncated Mie extinction efficiency can be reused for their derivatives.

Now everything is prepared to differentiate the squared magnitude of the Mie coefficient an with respect to npart and kpart. It is sufficient only to discuss an2 in the following because the differentiation of bn2, cn2 and dn2 works analogously.

First we write the squared norm of the Mie coefficient an as an2=anan¯, which givesnpartan2=npartanan¯+annpartan¯andkpartan2=kpartanan¯+ankpartan¯.

We writean=EDwithE:=mpartξn˙(zmed)ξn(zpart)-mmedξn(zmed)ξn˙(zpart)andD:=mpartψn˙(zmed)ξn(zpart)-mmedψn(zmed)ξn˙(zpart),

which yieldsddmpartan=1D2ddmpartED-EddmpartDwithddmpartE=ξn˙(zmed)ξn(zpart)+mpartξn˙(zmed)ρξn˙(zpart)-mmedξn(zmed)ρξn¨(zpart)andddmpartD=ψn˙(zmed)ξn(zpart)+mpartψn˙(zmed)ρξn˙(zpart)-mmedψn(zmed)ρξn¨(zpart).

Furthermore follow from (Equation3.5), the relationsnpartan=ddmpartanandkpartan=ddmpartani.

Although an¯ is not holomorphic with respect to mpart, we can still compute the partial derivatives npartan¯ and kpartan¯. We obtain using (Equation3.6) the relationsnpartan¯=npartan¯andkpartan¯=kpartan¯.

This completes the computations of npartan2 and kpartan2.

For a holomorphic function f(x+iy), we can easily deduce from (Equation3.6) thatxImf(x+iy)=Imxf(x+iy)andyImf(x+iy)=Imyf(x+iy).

This gives with respect to (Equation3.5)npartImAn=ImnpartAnwithnpartAn=l2πnpartcn2U1+cn2npartU1-npartdn2U2-dn2npartU2,whereU1=ξn(zpart)ξn˙(zpart)¯mpart,npartU1=1mpart2ρξn˙(zpart)ξn˙(zpart)¯+ξn(zpart)ρξn¨(zpart)¯mpart-ξn(zpart)ξn˙(zpart)¯,andU2=ξ˙n(zpart)ξn(zpart)¯mpart,npartU2=1mpart2ρξn¨(zpart)ξn(zpart)¯+ξ˙n(zpart)ρξn˙(zpart)¯mpart-ξ˙n(zpart)ξn(zpart)¯.

Analogously we obtainkpartImAn=ImkpartAnwithkpartAn=l2πkpartcn2U1+cn2kpartU1-npartdn2U2-dn2kpartU2,wherekpartU1=1mpart2(ρξn˙(zpart)iξn˙(zpart)¯+ξn(zpart)ρξn¨(zpart)i¯mpart-ξn(zpart)ξn˙(zpart)¯i),andkpartU2=1mpart2(ρξn¨(zpart)iξn(zpart)¯+ξ˙n(zpart)ρξn˙(zpart)i¯mpart-ξ˙n(zpart)ξn(zpart)¯i).

The derivatives of ImBn are much easier to compute, since the dependence on npart and kpart lies only in an2 and bn2 here. So we can deducenpartImBn=ImnpartBnwithnpartBn=l2πnpartan2ψn˙(zmed)ψn(zmed)¯mmed-npartbn2ψn(zmed)ψn˙(zmed)¯mmed.

Analogously we getkpartImBn=ImkpartBnwithkpartBn=l2πkpartan2ψn˙(zmed)ψn(zmed)¯mmed-kpartbn2ψn(zmed)ψn˙(zmed)¯mmed.

4. Nonlinear regression using truncated series expansions

We wish to reconstruct the refractive indices of a particle material from spectral measurements by solving a nonlinear regression problem of the form(4.1) Xt,δ:=argminxRDi=1N12σi2n=1tani(x)-n=1ani(xtrue)-δi2,(4.1)

where tN is a finite truncation index and δiN(0,si2). Remember that N represents the number of particle radii ri of the different monodisperse aerosols we are investigating. We still assume that for each radius ri the standard deviations si are determined well enough from a set of experiments, such that they can be regarded as known. We have to confine ourselves to a finite truncation index t because it is practically not feasible to compute all coefficient functions ani(x) for i=1,,N. Throughout this paper, we assume that the feasible set Ω is compact.

We define the functions ft:RDRN and f:RDRN byft(x):=n=1tan1(x),,n=1tanN(x)Tandf(x):=n=1an1(x),,n=1anN(x)T.

We set e:=f(xtrue)+δ with δ:=δ1,,δNT. Then the observed probability density is given bypobserved(e|x):=(2π)-Nl2|det(Σσ)|-12exp(-12Σσ-12(ft(x)-e)22)

with the covariance matrix Σσ:=diagσ12,,σN2. We know a priori that the vector x specifying our model ft(x) lies within the set Ω. This knowledge can be expressed with the prior probability densitypprior(x):=vol(Ω)-1IΩ(x),

where IΩ is the indicator function of Ω. Now Xt,δ is the set of MAP-estimators of the posterior probability density, i.e.(4.2) Xt,δ:=argmaxxpposterior(x|e)withpposterior(x|e)pobserved(e|x)pprior(x)exp(-12Σσ-12(ft(x)-e)22)IΩ(x).(4.2)

We carry out all the following investigations under the next assumption on the covariance matrix:

Assumption 4.1:

The covariance matrix Σσ has the simple formΣσ=δ2·diag(σ12,,σN2)=:δ2·Σ,

where δ0 is an arbitrary but fixed noise level and σ1, ..., σN are fixed.

To simplify notations, we introduce the two functions ft:RDRN and gt:RDRN depending on the truncation index t and defined byft(x)i:=n=1tani(x)+(t-t)at+1i(x)andgt(x)i:=f(x)i-ft(x)i,fori=1,,N.

In the following, we will investigate how an element xt,δ of the set Xt,δ depends on the truncation index t. We change to a continuous truncation index here, i.e. we change from now on from (Equation4.1) to the new regression problem(4.3) Xt,δ:=argminxRDFt,δ(x)s.t.xΩ,withFt,δ(x):=Σ-12(ft(x)-f(xtrue)-δ)22(4.3)

where the truncation index t0 is allowed to be non-integer.

As a preparation we prove the following technical lemma, which will form the basis of our continuity and convergence results.

Lemma 4.2:

Let the twice continuously differentiable function F:RNR have a strict local minimum x0 inside a compact set SRN. Let the function h:RN×RR have the property limε0h(x,ε)=0 for all xS and let h(x,ε) be twice continuously differentiable with respect to x and continuous in ε. Furthermore, we assume that the local minima xε of Fε(x):=F(x)+h(x,ε) are strict for any ε>0. Then there exists a sequence of local minima xε of Fε(x) with limε0xε=x0.

The strategy of the proof is to construct for given ε a neighbourhood of x0 which must contain a local minimizer xε of the perturbed function Fε(x). By sending ε to 0, this neighbourhood shrinks down to the local minimum x0 itself, thus yielding the convergence of xε to x0. To have this neighbourhood shrink down to x0, it is crucially important that x0 must be a strict local minimum.

We define d(ε):=supxS|h(x,ε)|. From limε0h(x,ε)=0 for all xS follows limε0d(ε)=0. Let us now introduce the function F-(x):=F(x)-d(ε). Obviously, x0 is also a local minimum of F-(x), so for ε sufficiently small there exists a neighbourhood U2d(ε)(x0)S of x0 withF-(x)F-(x0)andF-(x)-F-(x0)2d(ε)for allxU2d(ε)(x0).

In particular, we havexU2d(ε)(x0):F-(x)=F-(x0)+2d(ε)=F(x0)+d(ε).

Let us assume that there exists an xU2d(ε)(x0) withFε(x)<F(x0)+d(ε)=F-(x)=F(x)-d(ε).

Then Fε(x)=F(x)+h(x,ε) implies -d(ε)>h(x,ε), hence -h(x,ε)>d(ε)-h(x,ε) by definition of d(ε), contradiction. Therefore we conclude(4.4) xU2d(ε)(x0):Fε(x)F(x0)+d(ε).(4.4)

Since Fε(x) is continuous and U¯2d(ε)(x0) is compact for ε small enough, there exists an xεU¯2d(ε)(x0) withFε(xε)=minxU¯2d(ε)(x0)Fε(x).

Let us assume Fε(xε)>F(x0)+d(ε). Then by definition of xε, we get in particularF(x0)+h(x0,ε)=Fε(x0)Fε(xε)>F(x0)+d(ε),

i.e. h(x0,ε)>d(ε)h(x0,ε), contradiction. It follows(4.5) Fε(xε)F(x0)+d(ε)andFε(x0)F(x0)+d(ε),(4.5)

where the latter follows with a proof by contradiction as well.

If it happens to hold that Fε(xε)=F(x0)+d(ε), then we also have Fε(x0)=F(x0)+d(ε). Otherwise we have Fε(xε)<F(x0)+d(ε) and then (Equation4.4) implies that xε cannot lie on U2d(ε)(x0), thus it must lie within the interior of U2d(ε)(x0). So in any case (Equation4.5) gives that U2d(ε)(x0) must contain a local minimizer xε of Fε(x).

Now limε0d(ε)=0 gives limε0xε=x0. The existence of the last limit is guaranteed by the fact that x0 is strict and the claim is proved.

Proposition 4.3:

Let all coefficient functions ani(x) be twice continuously differentiable and bounded on Ω. We assume that each local minimum xt,δ of the right-hand side function Ft,δ(x) in (Equation4.3) is strict and lies in the interior of Ω. Then each local minimum depends continuously on the truncation index t.

To prove the claim, one could be tempted to apply the implicit function theorem on the equation d(t,xt,δ)=0 with d(s,x):=fs(x). This would give that the local minima are parameterized by a function m(s) with the property m(t)=xt,δ, where s is from an environment U(t) of t. The problem with this approach is that it requires continuous differentiability of d(s,x) in the truncation parameter s. Thus the continuous truncation we are using would need more complicated methods such as spline interpolation of the partial sums, which would increase the overall computational effort.

Therefore, we use in the following a more direct approach to prove the claim. Let ε>0 be arbitrary. First we consider an integer truncation index tN, i.e. we have t=t. Now for ε small enough, we get t+ε=t and t-ε=t-1. This givesft+ε(x)i=ft(x)i+εat+1i(x)andft-ε(x)i=ft-1(x)i+(1-ε)ati(x)=ft(x)i-εati(x).

As next step we turn to an noninteger truncation index t. In this case, we can always select ε small enough such that t+ε=t and t-ε=t, respectively hold. This yieldsft+ε(x)i=ft(x)i+εat+1i(x)andft-ε(x)i=ft(x)i-εat+1i(x).

Now we introduce the functiona(x):=(at1(x),,atN(x))T,fort-ε,tN(at+11(x),,at+1N(x))T,else.

For Ft,δ(x)=Σ-12(ft(x)-f(xtrue)-δ)22 this yieldsFt+ε,δ(x)=Ft,δ(x)+2εΣ-12a(x),Σ-12(ft(x)-f(xtrue)-δ)+ε2Σ-12a(x)22andFt-ε,δ(x)=Ft,δ(x)-2εΣ-12a(x),Σ-12(ft(x)-f(xtrue)-δ)+ε2Σ-12a(x)22.

Therefore, we obtain both for Ft+ε,δ(x) and Ft-ε,δ(x) a decomposition of the form Ft+ε,δ(x)=Ft,δ(x)+ht,δε(x) and Ft-ε,δ(x)=Ft,δ(x)+ht,δε(x) respectively, where the function ht,δε(x) is appropriately selected according to above findings. We can readily check limε0|ht,δε(x)|=0 for all xΩ from the boundedness of the ani(x)’s. Then the result follows from Lemma 4.2.

Corollary 4.4:

Let t1 and t2 be truncation indices with t1<t2. Let xt1,δ be a local minimizer of (Equation4.1). Let γ[0,1] and define tγ:=t1+γt2-t1. Then beginning at γ=0 one can successively find local minimizers xtγ,δ for the truncation index tγ using numerical continuation, see [Citation6]. Here for γ1<γ2 the minimizer xtγ1,δ is used as a start vector to compute the next minimizer xtγ2,δ. The next parameter γ2 has to be sufficiently close to γ1, such that the start vector xtγ1,δ still lies within the domain of convergence for Newton’s method.

We use Corollary 4.4 to compute xt,δ for increasing truncation index t in a stable way. If we would keep t as integer and increase it in integer steps, we might leave the domain of convergence in the continuation method. Therefore, we increase them using a smaller step width.

In the following, we investigate how well the minimizers xt,δ of the noise-contaminated regression problem (Equation4.3) with truncated series expansions approximate the minimizers x,0 of the noise-free and untruncated problem(4.6) x,0:=argminxRDi=1N12σi2n=1ani(x)-n=1ani(xtrue)2s.t.xΩ.(4.6)

Proposition 4.5:

Let the noise vector δ fulfil limδ0Σ-12δ22=0 and let the functions f(x) and ftδ(x) be bounded on Ω. Assume limδ0Σ-12gtδ(x)22=0 for all xΩ. Then for any strict minimizer x,0 of the right-hand side function of (Equation4.6) in the interior of Ω exist minimizers xtδ,δ of (Equation4.3) with limδ0xtδ,δ=x,0. Here we also assume the xtδ,δ’s to be strict for all δ>0.

With the notation introduced before, we can writex,0X,0:=argminxRDF,0(x)s.t.xΩwithF,0(x):=Σ-12(f(x)-f(xtrue))22.

From the decomposition ftδ(x)=f(x)-gtδ(x) we obtainFtδ,δ(x)=Σ-12f(x)-f(xtrue)-δ22-2Σ-12gtδ(x),Σ-12f(x)-f(xtrue)-δ+Σ-12gtδ(x)22.

Then a further decomposition of the first term on the right-hand side yieldsFtδ,δ(x)=F,0(x)-2Σ-12δ,Σ-12f(x)-f(xtrue)+Σ-12δ22-2Σ-12gtδ(x),Σ-12f(x)-f(xtrue)-δ+Σ-12gtδ(x)22=:F,0(x)+Htδ,δ(x).

From the limit limδ0Σ-12gtδ(x)22=0, the limit limδ0Σ-12δ22=0 and the boundedness of f(x) follows limδ0|Htδ,δ(x)|=0 for arbitrary but fixed xΩ.

Then the existence of the xtδ,δ’s follows from Lemma 4.2.

At last we study how the minimizers xtδ,δ of (Equation4.3) behave for δ0. We begin with a preparing corollary.

Corollary 4.6:

Let the assumptions of Proposition 4.5 hold. Then we have for any local minimizer xtδ,δ of (Equation4.3) approximating a local minimizer x,0 of (Equation4.6) for δ0 with Σ-12(f(x,0)-f(xtrue))2=0 thatlimδ0Σ-12(ftδ(xtδ,δ)-f(xtrue)-δ)2=0.

The assumptions of Proposition 4.5 givelimδ0Σ-12δ22=0andlimδ0Σ-12gtδ(xtδ,δ)22=0.

We have by continuity of f(x) that limδ0Σ-12(f(xtδ,δ)-f(xtrue))2=0. ThenΣ-12(f(xtδ,δ)-f(xtrue))2+Σ-12gtδ(xtδ,δ)2+Σ-12δ2Σ-12(ftδ(xtδ,δ)-f(xtrue)-δ)2

gives the desired result.

Proposition 4.7:

Let the assumptions of Proposition 4.5 hold. Assume that the local minimizers x,0 of (Equation4.6) with Σ-12(f(x,0)-f(xtrue))2=0 form a discrete set S,0. Then the set L,0 consisting of the limits limδ0xtδ,δ of local minimizers xtδ,δ of (Equation4.3) with limδ0Σ-12(ftδ(xtδ,δ)-f(xtrue)-δ)2=0 coincides with S,0 and there exists a noise level δmax such that all minimizers xtδ,δ approximating S,0 are isolated for all δδmax.

On the one hand from Proposition 4.5 we know that there exists a sequence xtδ,δ of minimizers of (Equation4.3) with limδ0xtδ,δ=x,0. Then Σ-12(f(x,0)-f(xtrue))2=0 and Corollary 4.6 give limδ0Σ-12(ftδ(xtδ,δ)-f(xtrue)-δ)2=0, which implies S,0L,0.

On the other hand holds for xtδ,δ with limδ0Σ-12(ftδ(xtδ,δ)-f(xtrue)-δ)2=0 thatΣ-12(ftδ(xtδ,δ)-f(xtrue)-δ)2+Σ-12gtδ(xtδ,δ)2+Σ-12δ2Σ-12(f(xtδ,δ)-f(xtrue))2

which implies limδ0Σ-12(f(xtδ,δ)-f(xtrue))2=0. In particular, this means by continuity of f(x) that the vector limδ0xtδ,δ must be a local minimizer of Σ-12(f(x)-f(xtrue))2. Thus we have also shown L,0S,0.

In the following, we number all elements of S,0 with the index k, i.e. we write x,0k for k=1,,|S,0|. Similarly we number all minimizers xtδ,δ with limδ0Σ-12(ftδ(xtδ,δ)-f(xtrue)-δ)2=0 approximating the x,0k’s with xtδ,δk, i.e. limδ0xtδ,δk=x,0k for k=1,,|S,0|. DefineDmin:=minijx,0i-x,0j2.

Since limδ0xtδ,δk=x,0k, we can find an error levels δmaxk such thatxtδ,δk-x,0k2<12Dminfork=1,,|S,0|,

which holds for all 0δδmaxk for each k. Then for all 0δδmax:=mink{δmaxk} the xtδ,δk’s must have pairwise mutual distances greater than zero.

Now Proposition 4.7 gives that the number of local minima xtδ,δk remains constant if the noise level δ is small enough. It also yields that these local minima then form a set of separated continuous curves parametrized in δ.

At last we wish to have an estimate of the convergence of the local minima xtδ,δk of the truncated and noise contaminated problem to the local minima x,0k of the noise-free and untruncated problem, which is useful for practical computations.

Proposition 4.8:

For the noise-level δ small enough, we can bound for any local minimum xtδ,δk the approximation error x,0k-xtδ,δk2 with a positively weighted linear combination of the residual Σ-12(ftδ(xtδ,δk)-f(xtrue)-δ)2, the truncation error Σ-12gtδ(xtδ,δk)2 and the noise estimate Σ-12δ2.

The first-order necessary conditions for a local minimum of Ftδ,δ(x)=F,0(x)+Htδ,δ(x) at xtδ,δk and a local minimum of F,0(x) at x,0k yield in particularF,0(xtδ,δk)+Htδ,δ(xtδ,δk),x,0k-xtδ,δk0andF,0(x,0k),xtδ,δk-x,0k0,

Adding the last two inequalities yields(4.7) Htδ,δ(xtδ,δk),x,0k-xtδ,δkF,0(x,0k)-F,0(xtδ,δk),x,0k-xtδ,δk.(4.7)

Since F,0(x) is totally differentiable at x,0k, we obtainF,0(xtδ,δk)=F,0(x,0k)+HessF,0(x,0k)xtδ,δk-x,0k+w,0(xtδ,δk,x,0k),

where w,0(x,x,0k) fulfilsw,0(x,x,0k)2x-x,0k2ϵ,0(x,x,0k)withlimxx,0kϵ,0(x,x,0k)=0.

Since HessF,0(x,0k) is positive definite, the expression x,HessF,0(x,0k)x12 gives a norm on RD. Because of the equivalence of all norms in RD, there exists a constant C,0k>0 withx,HessF,0(x,0k)x12C,0kx2for allxRD.

Since limδ0xtδ,δk=x,0k we can find a noise level ρmaxk such that |ϵ,0(xtδ,δk,x,0k)|d,0k for all δρmaxk, where d,0k is a constant with 0d,0k<C,0k2. Then usingw,0(xtδ,δk,x,0k),x,0k-xtδ,δkϵ,0(xtδ,δk,x,0k)x,0k-xtδ,δk22

and (Equation4.7) we can estimateHtδ,δ(xtδ,δk)2x,0k-xtδ,δk2F,0(x,0k)-F,0(xtδ,δk),x,0k-xtδ,δk=HessF,0(x,0k)x,0k-xtδ,δk-w,0(xtδ,δk,x,0k),x,0k-xtδ,δk(C,0k)2-ϵ,0(xtδ,δk,x,0k)x,0k-xtδ,δk22, i.e. this gives(4.8) x,0k-xtδ,δk2(C,0k)2-d,0k-1Htδ,δ(xtδ,δk)2(4.8)

for all δρmaxk.

We haveHtδ,δ(x)=2(JacgtδT(x)Σ-1gtδ(x)-JacgtδT(x)Σ-1(f(x)-f(xtrue)-δ)-JacfT(x)Σ-1gtδ(x)-JacfT(x)Σ-1δ),

i.e. we can find constants K1k and K2k such thatHtδ,δ(xtδ,δk)22(K1kJacgtδT(xtδ,δk)Σ-12Σ-12gtδ(xtδ,δk)2+K1kJacgtδT(xtδ,δk)Σ-12(Σ-12(ftδ(xtδ,δk)-f(xtrue)-δ)2+Σ-12gtδ(xtδ,δk)2)+K2kJacfT(xtδ,δk)Σ-12Σ-12gtδ(xtδ,δk)2+K2kJacfT(xtδ,δk)Σ-12Σ-12δ2),

which gives the result.

Corollary 4.9:

Let the truncation indices tδ depend on the vector of independent Gaussian random variables δ with limδ0E(Σ-12δ22)=0 such that limδ0E(Σ-12gtδ(x)22)=0 holds for all arbitrary but fixed xΩ. Then we have for all minimizers x,0k of (Equation4.6) with Σ-12(f(x,0k)-f(xtrue))2=0 that for δ sufficiently small there exist minimizers xtδ,δk of (Equation4.3) withlimδ0E(x,0k-xtδ,δk2)=0.

Proposition 4.5 establishes the existence of the xtδ,δk’s. Corollary 4.6 giveslimδ0E(Σ-12(ftδ(xtδ,δk)-f(xtrue)-δ)2)=0.

Proposition 4.7 gives that for δ sufficiently small there exists a constant K,0k withE(x,0k-xtδ,δk2)K,0kE(Htδ,δ(xtδ,δk)2).

Set S1k:=supxΩJacgtδT(x)Σ-12< and S2k:=supxΩJacfT(x)Σ-12<. Then the estimate for Htδ,δ(xtδ,δk)2 in the proof of Proposition 4.7 givesE(Htδ,δ(xtδ,δk)2)2(K1kS1kE(Σ-12gtδ(xtδ,δk)2)+K1kS1k(E(Σ-12(ftδ(xtδ,δk)-f(xtrue)-δ)2)+E(Σ-12gtδ(xtδ,δk)2))+K2kS2kE(Σ-12gtδ(xtδ,δk)2)+K2kS2kE(Σ-12δ2)), which proves the claim since Assumption 4.1 gives E(Σ-12δ2)Nδ.

The strategy for our retrieval algorithm is to start with an initial guess for the truncation index tstart and try to find all local minima xtstart,δk. Then the truncation index is gradually increased and starting from xtstart,δk the continuation method is applied to find finally the local minima xtδ,δk. Motivated by Propositions 4.8 and 4.9 only those local minima are considered to be possible approximations to our sought-after refractive index, where the residual Σ-12(ftδ(xtδ,δk)-f(xtrue)-δ)2 and an estimate of the truncation error Σ-12gtδ(xtδ,δk)2 are both reasonably small. The latter serves also as a stopping criterion for the continuation method

The initial guess tstart has to be selected with care. On the one hand if it is to small, the model is to inaccurate and the retrieval of the sought-after local minima cannot be guaranteed. On the other hand if it is too big, computational effort is wasted, since too many Mie coefficient functions with almost vanishing magnitudes and thus essentially not changing the local minima are computed.

Remark 4.10:

We did not explicitly show the assumption that the local minima of (Equation4.3) are strict as demanded in Proposition 4.5. However we could always determine a set of strict local minima for any truncation index in our simulations.

5. The reconstruction algorithm

We now return to our regression problem (Equation1.3). For i=1,,N we see that the measured extinctions normalized by the number of particles ni with radius ri, i.e. the quantity eini, is Gaussian-distributed with mean 1niqtrue(ri,l) and standard deviation σi:=sini. In the following we fix a wavelength l, i.e. we reconstruct the sought-after particle refractive index mpart(l) wavelength by wavelength. In the following, the unit both for particle radii and light wavelengths is μm.

We make use of the function qNtr:R2RN defined by(5.1) qNtr(x):=πr12n=1Ntrqn(x,r1,l),,πrN2n=1Ntrqn(x,rN,l)T,(5.1)

where R={r1,,rN} is the particle radius grid and Ntr the truncation index to be used. We allow non-integer truncation indices Ntr as well, where the non-integer truncation is done like in Proposition 4.3. Here the expression qn(x,rk,l) is a short notation of qn(mmed(l),(x)1+(x)2i,rk,l) from Section 1, where the sought-after refrative index mpart(l) is identified with the vector x here, i.e. mpart(l)=(x)1+(x)2i. So its computation follows Section 2.

In the following, the refractive index search area is given by the rectangle Ω:=[0,20]×[0,40], which means that we only consider refractive indices of particle materials whose real parts lie in the interval [0, 20] and its imaginary parts in the interval [0, 40]. This rather large search area makes the algorithm suitable for a wide range of aerosol materials.

In the first loop from lines 14–25, a search for local minima of the fit function F(x) defined in line 16 for the truncation index Ntr=3 is performed. The loop runs through all grid points (ci,dj)T of the search grid defined in lines 5 and 6. If the Hessian of F(x) at some grid point (ci,dj)T is positive definite, this point might lie in the vicinity of a local minimum. The Hessian is computed exactly, where the second partial derivatives of the Mie extinction efficiency with respect to the real and imaginary part of the scattering material needed here are computed using the product rule approach from Section 3. So we use (ci,dj)T as start point for a local solver in this case. In line 20, we only accept a new local minimum if it is sufficiently different from the local minima already found. Then it is stored in the container Sstart. This simple global search strategy can find all local minima if the search grid is fine enough.

Remark 5.1:

Of course other well-established global search heuristics can be applied here as well. In test runs, we compared genetic algorithms with our sequential search strategy on the two-dimensional refractive index search area, but their computational effort and reliability remained the same or were even worse. In [Citation7], the technique of simulated annealing was used to retrieve aerosol refractive indices, which could be a promising alternative here. In our study, the measurement noise was so high, that a unique global minimum of our fit function could not be determined. Instead our focus lied on effectively finding all local minima with small values of the fit function and we regarded them all as possible approximations to a thought-after refractive index.

The second loop from lines 29–49 uses the local minima found in the first loop as start points for the continuation method following Proposition 4.3 and Corollary 4.4. We found that a step width of 0.1 is for our problem a well-balanced choice between too big step widths rendering the continuation method unstable and too small step widths making it computationally inefiicient. With the stopping criterion DrelTolrel of the while-loop, it is approximately checked if the magnitude of the remainder term is small enough. Finally in line 44, it is checked if the residual is small enough. In our implementation, we did another run of lines 44–48 with τ=5 and τ=7, respectively, if none of the reconstructions had a squared residual smaller than τNrδ2 for the previous τ. This had to be done, because the parameter τ has to be selected carefully in order to estimate the bound on E(x,0k-xtδ,δk2) derived in the proof of Corollary 4.9 correctly.

6. Comparison of the Numerical Continuation Approach with Established Truncation Index Heuristics

As solution of the forward problem, we generated for a discrete set of wavelengths l1, ..., lNl unperturbed spectral extinctions normalized with the number of particles of the monodisperse aerosol by computingetruei,j:=πri2n=1Ntrqn(mmed(lj),mpart(lj),ri,lj),fori=1,,N,j=1,,Nl

with mpart(li) taken as the refractive indices of Ag, H2O and CsI. Here we used the truncation index(6.1) ρ=2πrl,M=max{|ρ|,|ρ·mpart(l)|,|ρ·mmed(l)|},Ntr:=|M+4.05·M13+2|(6.1)

introduced in [Citation4].

For particle size distribution reconstructions as outlined in [Citation2], we need particle refractive indices for five optical windows, see [Citation8], so the wavelength grid of interest consists of five ranges. These ranges are given by 8 linearly spaced wavelengths from 0.6–0.8 μm, 8 from 1.1–1.3μm, 8 from 1.6–1.8μm, 16 from 2.1–2.5μm and 8 from 3.1–3.3μm, so we have in total Nl=48 wavelengths.

For each of the 48 wavelengths, we generated noisy spectral extinctions e by adding zero-mean Gaussian noise to etrue, i.e.ei,j=etruei,j+δi,jwithδi,jN(0,(0.05·etruei,j)2),i=1,,Nj=1,,Nl.

Here the standard deviations were taken to be 5% of the original extinction values. We computed each mean ereali,j of the noisy spectral extinctions with a sample size of Ns=300.

6.1. Run Time Comparison

In the following, Algortihm Equation1 is referred to as method 1. On the same simulated spectral extinctions, we let Algorithm 1 run up to line 25, but with the difference that at each evaluation of qNtr(x) we directly took the trunction index from (Equation6.1). We denote this approach with method 2. We now display the average run times of method 1 and method 2 for 10 sweeps through all 48 wavelenghs.

6.1.1. Results for Ag

6.1.2. Results for CsI

6.1.3. Results for H2O

6.2. Maximal relative deviations

For the 10 simulation runs, we list the maximal relative deviations100·npart1(lj),kpart1(lj)T-npart2(lj),kpart2(lj)T2npart1(li),kpart1(lj)T2

of the refractive index reconstructions npart2(lj),kpart2(lj)T from method 2 from npart1(lj),kpart1(lj)T of method 1 for j=1,,48. At each wavelength, multiple local minima can be detected by both methods. For the relative deviations, we always selected the local minima forming the smoothest reconstructions on each optical window in the sense of Section 8.

6.2.1. Results for Ag

6.2.2. Results for CsI

6.2.3. Results for H2O

6.3. Conclusion

For Ag, the average total run time over all 48 wavelengths for method 1 was 44.0167% less than for method 2, for H2O 43.9808% and for CsI 44.7322%, i.e. method 1 is almost two times faster than method 2. The results are of the same quality, since their relative deviations are just small fractions of percentages.

The continuation method approach saves run time significantly with the same quality of the results compared to using the truncation index (Equation6.1) all the time.

7. Nonlinear Tikhonov regularization

So far we have solved the regression problem (Equation4.3) without any regularization, thus the obtained refractive index reconstructions might still be too error-contaminated to be of practical use. A widely used regularization strategy for nonlinear regression problems is Tikhonov regularization, which yields the regularized regression problem(7.1) xγ:=argminxRDΣ-12(ft(x)-f(xtrue)-δ)22+γx-x22s.t.xΩ(7.1)

when we apply it on (Equation4.3), cf. [Citation9]. Here γ is a regularization parameter and x is an estimate of the sought-after true solution. In many cases, the unregularized problem has a whole set of minimizers, thus the vector x works also as a selection criterion. Now if a reasonable x is found, the regularization parameter γ can be determined with the discrepancy principle, i.e. γ is computed such thatΣ-12(ft(xγ)-f(xtrue)-δ)2=R(δ),

where R(δ) is an estimate of the residual of the ‘true’ solution which depends on the noise level δ. For this task, monotonicity in the residual of xγ is established in [Citation9].

The problem of finding a good estimate x still remains. In [Citation10], an alternative implementable parameter choice strategy without the need of an x is derived. Applied on our problem, it givesγΣ-12(ft(xγ)-f(xtrue)-δ),Jγ-1(Σ-12(ft(xγ)-f(xtrue)-δ)=R(δ),withJγ:=γI+Σ-12Jacft(xγ)Jacft(xγ)TΣ-12.

This method has the drawback that the matrix Jγ needs to be inverted, which may lead to instabilities.

Nevertheless the quality of the regularized solutions still depends strongly on the start values for solving (Equation7.1). We know about our sought-after refractive indices that they form smooth curves on each of the five optical windows. The complex refractive index curves of most materials can be described using the so-called Lorentz-oscillator-model, cf. [Citation11]. Here points with bigger curvatures only occur at so-called resonance frequencies corresponding to some isolated resonance wavelengths. Motivated by these facts, we derive in the following a method to find reasonable start values for Phillips–Twomey regularization out of the results of Algorithm 1, which will be outlined in Section 9.

8. Finding the smoothest coupled solutions

We have the problem of identifying the best approximation to the sought-after true particle material refractive index xtrue out of a set of multiple solutions obtained with Algorithm 1. This problem can be solved by coupling the solutions, which means that we combine solutions from neighbouring wavelengths l in each of the five optical windows in order to obtain a unique solution for every optical window. We know about the complex refractive index curves to be retrieved that they are smooth; hence, we expect their sum of the squared second finite differences both in the real and imaginary parts to be small.

Let l1, ..., ls denote the wavelengths of any of our five wavelength ranges. Let N1, ..., Ns be the number of solutions found for all the wavelengths. We denote with xji the jth solution found for wavelength li for i=1,,s and j=1,,Ni. Now we wish to find the smoothest combined solution from all possible combinations xj11, ..., xjss for ji=1,,Ni, hence we have a total number of i=1sNi combinations. Here we measure smoothness of a combination xj11, ..., xjss with the sumS:=i=2s-1(xji-1i-1)1-2(xjii)1+(xji+1i+1)12+(xji-1i-1)2-2(xjii)2+(xji+1i+1)22

of its second finite differences both in the real parts (xjii)1 and its imaginary parts (xjii)2, which means that we regard a combination the smoother the smaller its sum S is.

We encounter the problem that the total number of possible combinations i=1sNi might get too big to iterate through all combinations in the search for the smoothest one in acceptable time. Therefore we propose a greedy algorithm, which uses each second finite difference as start point to find a smooth combination.

The main loop spanning over the lines 5–52 iterates through all positions z=2,,s-1 of a middle point for a second finite difference both in the real and imaginary part. At each position z, the inner loops beginning in lines 6–8 iterate through all possible second finite differences which can be formed out of the vectors xc1z-1, xc2z and xc3z+1 for c1=1,,Nz-1, c2=1,,Nz and c3=1,,Nz+1, i.e. they loop through all of their middle points and left and right neighbours at position z. In lines 9–12, the variable Scur is initialized with the sum of the squared second finite differences in the real and imaginary parts of the current vectors xc1z-1, xc2z and xc3z+1 and the positions z-1, z and z+1 of the array Comb are filled with the current vectors. For z3, the loop in lines 13–28 successively fills the positions k=z-2,z-3,,1 of the array Comb. At each new position k, the minimal sum of the two squared second finite differences in the real and imaginary part Dmin is determined in lines 19–25, where the middle and right points are fixed and taken as the leftmost two vectors from the array Comb and the right point runs through all xjk for j=1,,Nk. After the vector xmin giving out of all of the xjk’s giving the minimal sum Dmin is found, the Dmin is added to Scur and xmin is stored in kth entry Comb(k). In a similar way, the loop in lines 29–44 succesively fills the positions k=z+2,z+3,,s for zs-2. This time the left and middle points are fixed and taken as the rightmost two points of the array Comb, whereas the left point iterates through all xjk for j=1,,Nk. Again the vector xmin is that one of the xjk’s giving the minimal sum Dmin and it is stored in Comb(k). As well the sum Dmin is added to Scur.

In the above procedure, every triple of neighbouring vectors from the results of Algorithm 1 is considered to possibly lie on the sought-after smoothest combination with the smallest sum of all squared second differences. The three vectors are used as start points to find a smooth combination with a greedy strategy, where only a vector is added to the current combination, if it gives the smallest sum Dmin at the left or right end of the growing set with vectors already added, until the first and last position are reached.

Finally, from all of the combinations constructed this way the smoothest one with the smallest sum Smin out of all Scur’s is selected in lines 45–48 to be the final output SmoothestCombination.

Define Ntotal:=j=1sNj. Then the total number of operations needed for Algorithm 2 can be estimated by O(Ntotalj=2s-1Nj-1NjNj+1) which is considerably less than the O(j=1sNj) operations needed by the naive method of iterating through all possible combinations.

9. Further regularization of coupled solutions

Not only for determining the smoothest refractive index curve reconstructions formed from the results of Algorithm 1 the coupled view on the solutions is beneficial – it also leads to further improvement of the results by Twomey-regularization. Let us investigate the coupled approach in a probability theoretical setting. Here we reuse the notations introduced in Section 4, i.e. we let x1, ..., xs denote a set of solution for any of the five optical windows. Then the joint posterior probability density of x1, ..., xs is given by(9.1) p(x1,,xs|e1,,es)=j=1sp(xj|ej)exp-12j=1sΣj-12(ftj(xj)-ej)22×j=1sIΩ(xj),(9.1) where ej is the data vector for the jth wavelength lj having N entries with N being the size of the radius grid. Moreover, Σj is the scaled covariance matrix for lj and ftj(x) is the applied model depending on lj. Note that we initially have differing truncation indices t1, ..., ts. Since the coefficient functions of the truncated model function ftj(x) are decaying fast for each tj, it is convenient to change to the same truncation index t:=max{t1,,ts} for all wavelengths l1, ..., ls. The errors introduced by doing so are negliglible. It is easy to show that maximizing the joint density (Equation9.1) is equivalent to maximize all single densities p(xj|ej) independently, i.e. a joint MAP-estimatorxopt1,...,xoptsXopt:=argmaxx1,...,xsp(x1,...,xs|e1,...,es)

consists of the single MAP-estimatorsxoptjXoptj:=argmaxxp(x|ej)

for j=1,,s. This means the the results of Algorithm 1 can be used to construct MAP-estimators for the joint posterior probabilty density.

This behaviour changes when we replace the joint prior probabilty densitypprior(x1,,xs)=vol(Ω)-sj=1sIΩ(xj)

withpprior(x1,,xs)exp-12γS(x1,,xs)j=1sIΩ(xj),

whereS(x1,,xs):=i=2s-1((xi-1)1-2(xi)1+(xi+1)1)2+((xi-1)2-2(xi)2+(xi+1)2)2+ρi=1s((xi)12+(xi)22),

where γ is a regularization parameter and ρ is a parameter specifying the amount Tikhonov regularization.

In the new prior distribution, we use a combination of Tikhonov and Phillips–Twomey regularization both in the real and imaginary parts. Here we apply a small amount of Tikhonov-regularization by setting ρ=10-8, such that the resulting regularization operator gets regular. This means that the regularized regression problem (Equation9.2) can be transformed into standard Tikhonov form and that the monotonicity results from [Citation9] are still valid. Each second finite difference is clearly a function of three neighbouring points; therefore, a decoupled computation of the joint MAP-estimator(9.2) xopt1,...,xoptsXopt:=argmaxx1,...,xsp(x1,...,xs|e1,...,es)(9.2)

withp(x1,,xs|e1,,es)exp-12j=1sΣj-12(ftj(xj)-ej)22-12γS(x1,,xs)×j=1sIΩ(xj)

for each wavelength seperately is not possible anymore after changing to the new prior density. However the result vectors x1, ..., xs from Algorithm 2 form a good start vector to solve the nonlinear regression problem (Equation9.2).

We selected the regularization parameter γ using the discrepancy principle, i.e. we compute γ such that the regularized solutionxγ1,...,xγsXγ:=argminx1,...,xsj=1sΣj-12(ftj(xj)-ej)22+γS(x1,...,xs)s.t.xjΩ,j=1,...,s

fulfils a relation of the formj=1sΣj-12(ftj(xγj)-ej)22=R(δ),

where R(δ) is a proposed residual value depending on the noise level δ. In [Citation2], several different residual values are proposed for a fixed model discretization and a set of regularization parameters is obtained from those using the discrepancy principle. The pairings of model discretizations and regularization parameters obtained this way are compared by their Bayesian posterior probabilities. For these probabilities, a set of integrals of the different model posterior densities is needed to be computed, which can be done with Monte Carlo integration methods. Due to the highly nonlinear behaviour of our model ft(x) such integration methods are not available here. Therefore, we simplified the posterior exploration in such a way that only one residual value is proposed.

Since each observed probability density p(ej|xj) for j=1,,s is Gaussian, the joint observed density p(e1,,e1|x1,,xs)=j=1sp(ej|xj) is Gaussian as well. We have xjR2, thus the sum of residuals j=1sΣj-12(ftj(xj)-ej)22 running through all wavelengths in the optical window is χ2(2s)-distributed. This yieldsE(j=1sΣj-12(ftj(xj)-ej)22)=2s.

Now a widely proposed residual value for the discrepancy principle is τ·2s, where τ=1.1 is the so-called Morozov safety factor. This choice is prone to under- or overregularization since the residual value corresponding to the ‘true’ solution might be much smaller or bigger than 2τs. Therefore, we proposed a residual value which depends more dynamically on the observed behoviour of the residual.

Let x01, ..., x0s denote the unregularized solutions, i.e. the results of Algorithm 2. Then their squared residual is given by R0:=j=1sΣj-12(ftj(x0j)-ej)22. We first proposedR(δ)=max2τ1s,τR0,

where we selected τ1=1.1. This means that the residual of the regularized solution is beginning at R0 at least increased by the factor τ1, which avoids underregularization. If it then happens that R(δ)R0>θ with θ=1.5 the proposed residual is most likely too big and overregularization occurs. In this case, we corrected R(δ) by settingR(δ)=max2τ2s,θR0

with τ2=0.9.

10. Numerical results

To see how reliable our proposed reconstruction algortihm is, we performed for each of the scatterer materials Ag, H2O a numerical study with 100 sweeps through all 48 wavelenghts of the five optical windows with the same settings as in Section 6. We found out that the radii r1:=0.1μm, r2:=0.2μm and r3:=0.3μm contain the most information about the refractive indices. This was found by keeping our 48 wavelengths fixed and comparing the quality of inversion results under varying aerosol particle radii. Bigger radii did not improve the results in our simulations and refractive index reconstructions only using bigger radii even turned out to be too unstable. A more thorough treatment of this problem can be found in [Citation7], where a covariance eigenvalue analysis is used. Although not directly comparable with our study of uncoated particles, the coated radii 0.0975μm, 0.2305μm and 0.11μm carrying the most information content found in this study are roughly comparable to our radii.

We computed original spectral extinctionsetruei,j:=πri2n=1Ntrqn(mmed(li),mpart(li),rj,li),i=1,,48,j=1,,3

for Ag, H2O and CsI and added zero-mean Gaussian noise to it in order to obtain the simulated noisy spectral extinctionsei,j=etruei,j+δi,jwithδi,jN(0,(0.05·etruei,j)2),i=1,,48,j=1,,3.

The standard deviations were taken to be 5% of the true spectral extinctions. Real experiments using 500 wavelengths were contaminated by Gaussian noise with 30% of the true spectral extinctions as standard deviations. We expect that switching to 48 wavelengths and thus increasing the time resolution of the measurements will lower the standard deviations to a small percentage. We used a sample size of Ns=300 to compute each mean ereali,j of noisy spectral extinctions.

In the following, the results are presented separately for each of the three materials. The uppermost plot presents the relative errors of the unregularized solutions obtained with Algorithm 2 from the original scatterer refractive indices. The next plot displays the run times of Algorithm 1, which returned all local minima of (Equation4.3). These candidate solutions served as input for Algorithm 2. Then the relative errors of the regularized solutions are presented. Finally, the relative errors of the average of the regularized solutions are shown.

10.1. Results for Ag

10.1.1. Results of Algorithms Equation1 and Equation2

10.1.2. Relative errors of the regularized solutions

10.1.3. Relative errors of the average of the regularized solutions

10.2. Results for CsI

10.2.1. Results of Algorithms Equation1 and Equation2

10.2.2. Relative errors of the regularized solutions

10.2.3. Relative errors of the average of the regularized solutions

10.3. Results for H2O

10.3.1. Results of Algorithms Equation1 and Equation2

10.3.2. Relative errors of the regularized solutions

10.3.3. Relative errors of the average of the regularized solutions

10.4. Conclusion

The severest relative errors can be observed for Ag. For the initial unregularized solutions, they lie between 1 and 5% on average and can go up to ca. 53% in the extreme cases as one can see in the leftmost subplot for the first optical window. The run times of Algorithm 1 lie between 30 and 50 s in the average case and can rise up to 200 s in the extreme cases. A typical sweep through all 48 wavelengths needed ca. 30 minutes in total and this value was very much the same for all three materials. For Ag, the regularization procedure effectively reduced the relative errors such that they are in the range between 0.5 and 2.2% on average and are below 10% in the extreme cases. Finally, one can see in the last plot that the relative errors of the average of all 100 regularized solutions are all below 0.4%.

For CsI, the relative errors of the unregularized solutions are already quite small and lie between 0.03 and 0.065% on average and rise only up to 0.3% in the extreme cases. The run times of Algorithm 1 are typically in the range from 25 to 55 s and are always below 95 s. The regularization of the solutions brought only a small improvement of the results here such that the relative errors did not change much. They are still in the same range from 0.03 to 0.065% on average but only reach up to ca. 0.2% now. The relative errors of the average of the 100 regularized solutions are between 0.01 and 0.055%.

Also for H2O, the relative errors of the unregularized solutions are comparably small and below 0.35% on average and still below 1.3% in the extreme cases. Especially, the rightmost subplot for the last optical window shows the biggest relative errors, whereas for all the other optical windows the relative errors are below 0.03% on average and below 0.15% in the extreme cases. A similar behaviour can be observed for the run times of Algorithm 1. For the first four optical windows, they are between 20 and 45 s on average and below 100 s in the extreme cases, whereas for the last optical window they are between 30 and 170 s on average and can even rise up to 350 s. For H2O, the regularization procedure improves the relative errors only slightly for the first four optical windows and even increases them for the last optical window such that they can rise up to ca. 0.06% on average and 1.4% in the extreme cases. The relative errors of the average of the 100 regularized solutions are virtually zero for the first four optical windows and below 0.55% for the last optical window.

11. Higher noise levels

To see how our proposed reconstruction algortihm behaves for higher noise levels, we performed for each of the scatterer materials Ag, H2O and CsI two numerical studies with 10 sweeps through all 48 wavelenghts of the five optical windows with the same settings as in Section 6. We computed original spectral extinctionsetruei,j:=πri2n=1Ntrqn(mmed(li),mpart(li),rj,li),i=1,,48,j=1,,3

for all wavelengths and added zero-mean Gaussian noise to it in order to obtain the simulated noisy spectral extinctionsei,j=etruei,j+δi,jwithδi,jN(0,(0.15·etruei,j)2),i=1,,48,j=1,,3.

for the first study andei,j=etruei,j+δi,jwithδi,jN(0,(0.3·etruei,j)2),i=1,,48,j=1,,3.

for the second. The standard deviations were taken to be 15% and 30%, respectively, of the true spectral extinctions. We used a sample size of Ns=300 to compute each means ereali,j of noisy spectral extinctions.

For brevity, we only present the relative errors of the average of the 10 regularized solutions.

11.1. Results for Ag

11.1.1. Standard deviation of 15%

11.1.2. Standard deviation of 30%

11.2. Results for CsI

11.2.1. Standard deviation of 15%

11.2.2. Standard deviation of 30%

11.3. Results for H2O

11.3.1. Standard deviation of 15%

11.3.2. Standard deviation of 30%

11.4. Conclusion

Whereas the relative errors for CsI and H2O are still below 1%, they can rise up to ca. 53% for Ag. Therefore, the reconstructed refractive indices for Ag under this noise level are most likely not of practical use. This shows that the FASP measurements of monodisperse aerosols must be sufficiently accurate in order to retrieve the scatterer refractive indices from them.

12. Numerical study

We performed four numerical studies for two-component aerosols with log-normal, RRSB and Hedrih model size distributions as outlined in [Citation2]. The aerosol particles were assumed to be homogeneously internally mixed, such that only one effective refractive index was retrieved. One component of the simulated aerosols was H2O with volume fractions of 0, 11, 22, 33, 44, 56, 67, 78, 89 and 100%. In the first two studies, we simulated mixtures of H2O and CsI, where we used the original aerosol component refractive indices for the first study. For the second study, we used the average of the 100 regularized solutions from Section 10. We did the same for the third and fourth studies, but here we simulated mixtures of H2O and Ag. In the third study, we utilized the original aerosol component refractive indices and for the fourth the average of the 100 regularized solutions from Section 10.

We applied the same reconstruction methods described in [Citation2] under the same settings, i.e. for each reconstruction we generated 300 artificial noisy measurements for all 48 wavelengths, where the measurement error was simulated as additive zero-mean Gaussian noise. For each wavelength, the standard deviations were taken as 5% of the solutions of the forward problem. In [Citation2], three different regularization methods, namely Tikhonov, minimal first differences and Twomey regularization, were compared and their results turned out to be very similar. Therefore, we only used Tikhonov regularization in the following. The results for the first study were directly adopted from [Citation2].

For every inversion, we computed the L2-error of the obtained reconstruction relative to the original size distribution and measured the total run time needed for the inversion. The computations were performed on a notebook with a 2.27 GHz CPU and 3.87 GB accessible primary memory.

13. Results for mixtures of H2O and CsI

13.1. Noise-free refractive indices

13.2. Noisy refractive indices

13.3. Results for mixtures of H2O and Ag

13.3.1. Noise-free refractive indices

13.3.2. Noisy refractive indices

13.4. Conclusion

The resuts of the first and second study only differ by ca. 4% at most and behave very similarly. The same is for the third and fourth studies. These numerical results indicate that 100 FASP measurement sweeps consisting of 300 single measurements with an accuracy as in Section 10 are sufficient to determine aerosol refractive indices in such a quality, that they are suitable for particle size distribution reconstructions for two-component homogeneously internally mixed aerosols using the FASP. The particle radii of the three monodisperse aerosols generated for the refractive indices retrieval need to be 0.1μm, 0.2μm and 0.3μm, respectively.

14. Outlook

It is of interest to investigate if the methods derived in this study can be extended to the case of core-plus-shell aerosols.

Acknowledgements

I thank Graham Alldredge, Ph.D. for proofreading the manuscript, useful recommendations and fruitful discussions on this topic.

Additional information

Funding

This work is sponsored by the German Federal Ministry of Education and Research (BMBF) [contract number 02NUK022A0]. Responsibility for the content of this report lies with the authors.

Notes

No potential conflict of interest was reported by the author.

References

  • Riziq AA, Erlick C, Dinar E, et al. Optical properties of absorbing and non-absorbing aerosols retrieved by cavity ring down (CRD) spectroscopy. Atmos Chem Phys. 2007;7:1523–1536.
  • Alldredge G, Kyrion T. Robust inversion methods for aerosol spectroscopy. Inverse Probl. Sci. Eng. 2016.
  • Fu Q, Sun W. Mie theory for light scattering by a spherical particle in an absorbing medium. Appl. Opt. 2001;40(9):1354–1361.
  • Wiscombe W. Improved Mie scattering algorithms. Appl Opt. 1980;19(9):1505–1509.
  • Abromovitz M, Stegun IA. Handbook of mathematical functions with formulas, graphs, and mathematical tables. Washington (DC): Dover; 1972.
  • Deuflhard P, Hohmann A. Numerical mathematics I: an algorithmically oriented introduction. Berlin: de Gruyter; 2002.
  • Erlick C, Haspel M, Rudich Y. Simultaneous retrieval of the complex refractive indices of the core and shell of coated aerosol particles from extinction measurements using simulated annealing. Appl Opt. 2011;50(22):4393–4402.
  • Goody RM, Yung YL. Atmospheric radiation. Theoretical basis. 2nd ed. New York (NY): Oxford University Press; 1989.
  • Engl HW, Kunisch K, Neubauer A. Convergence rates for Tikhonov regularisation of non-linear ill-posed problems. Inverse Prob. 1989;5:523–540.
  • Scherzer O, Engl HW, Kunisch K. Optimal a posteriori parameter choice for Tikhonov regularization for solving nonlinear Ill-posed problems. SIAM J Numer Anal. 1993;6:1796–1838.
  • Quinten M. Optical properties of nanoparticle systems. Weinheim: WILEY-VCH Verlag; 2010.

Reprints and Corporate Permissions

Please note: Selecting permissions does not provide access to the full text of the article, please see our help page How do I view content?

To request a reprint or corporate permissions for this article, please click on the relevant link below:

Academic Permissions

Please note: Selecting permissions does not provide access to the full text of the article, please see our help page How do I view content?

Obtain permissions instantly via Rightslink by clicking on the button below:

If you are unable to obtain permissions via Rightslink, please complete and submit this Permissions form. For more information, please visit our Permissions help page.