<<

Extinction spectra of suspensions of microspheres: Determination of spectral and particle size distribution with nanometer accuracy

Jonas Gienger,∗ Markus Bär, and Jörg Neukammer Physikalisch-Technische Bundesanstalt (PTB), Abbestraße 2–12, 10587 Berlin, Germany (Dated: Compiled October 5, 2017) A method is presented to infer simultaneously the wavelength-dependent real refractive index (RI) of the material of microspheres and their size distribution from extinction measurements of particle suspensions. To derive the averaged spectral optical extinction of the microspheres from such ensemble measurements, we determined the particle concentration by flow cytometry to an accuracy of typically 2% and adjusted the particle concentration to ensure that perturbations due to multiple are negligible. For analysis of the extinction spectra we employ Mie theory, a series-expansion representation of the refractive index and nonlinear numerical optimization. In contrast to other approaches, our method offers the advantage to simultaneously determine size, size distribution and spectral refractive index of ensembles of microparticles including uncertainty estimation.

I. INTRODUCTION can be inferred from measurements of the scattering and absorption of light by the particles. The refractive index (RI) describes the refraction of a A reference case is that of homogeneous spheres de- beam of light at a (macroscopic) interface between any scribed by a single refractive index, since an analytical two materials. Consequently, a variety of experimental solution for the mathematical problem of light scatter- methods exist for measuring the RI of a material that rely ing exists for this class of particles (Mie theory) [12, 13]. on the refraction or reflection of light at a planar interface This makes the analysis of light scattering data feasible between the sample and some other known material, such and at the same time is a good approximation for many as air, water or an optical glass. This approach is feasible real-world situations. Thus far, several techniques to in- for materials that can exhibit macroscopically large de- fer the RI or the size of spherical or small particles from fined interfaces, such as bulk liquids, homogeneous solids measurements of the scattering or extinction of light have or thin films and permits RI measurements with high ac- been discussed in the literature [3–9, 14–16] These tech- curacy. For example, the refractive index of synthetic niques can be divided into those relying on the angular calcium fluoride has been determined with high accuracy scattering pattern at a single wavelength [3, 4, 7, 14], between 138 nm and 2326 nm [1] and for distilled water and into those techniques relying on spectra of extinc- [2] applying the minimum deviation method. However, tion or diffuse transmittance [5, 8, 9, 15, 16]. Of course, in many cases one is interested in the optical properties of a combination of these techniques is also possible [6]. materials that do not have a homogeneous macroscopic The inference problem or inverse problem “What is form, such as atmospheric [3–5], soot particles in the RI and size distribution for a given extinction spec- flames [6], biological cells or tissues [7, 8] or man-made trum?” is generally challenging. Because the samples nanoparticles [9]. Often times, these materials cannot be used and quantities measured for the various approaches condensed into a homogeneous bulk sample. If the size described in the literature are so different, researchers is comparable to the wavelength of visible light, media have come up with tailor-made solutions for their par- containing these particles (e. g., a suspension of cells or ticular fields of application. For example, in Ref. 8, the a ) are turbid, i. e., light is scattered in a complex authors used combined collimated and diffuse transmit- process instead of propagating in straight rays. Never- tance measurements to determine the RI of polystyrene theless, the interaction of light waves with such materials spheres of known size. In Ref. 5, the authors analyzed at- is still determined by the optical and geometrical prop- mospheric aerosols of variable chemical composition. The erties of the constituting particles. It is thus reasonable complex refractive index was restricted to those functions represented by a multi-band damped-harmonic-oscillator arXiv:1710.01658v1 [physics.] 4 Oct 2017 to ask the question “What is the RI of the small parti- cles contained in the inhomogeneous sample?”. This RI model. This allowed to retrieve the RI of the aerosols. can yield information about the chemical composition of However, the approach was unable to represent asymmet- the particles or can be used for further mathematical ric dispersion features. The method described in Ref. 9 modeling of light scattering processes. For similar pur- was developed for nanoparticles and consequently could poses, one is interested in the size of the particles under be applied only to particles of sub-wavelength size. Mea- consideration, e. g., to calculate the force exerted by an surements in different suspending media were required for optical trap [10, 11], For certain cases both properties RI determination. The authors of Ref. 16 explicitly ana- lyzed the location of ripples in the extinction spectrum of single microspheres using synchrotron radiation. While the RI could be determined quantitatively from the the ∗Electronic address: [email protected] relative shape of these ripples, an accurate particle sizing 2

Figure 1: Schematic of experimental setup to measure spectral extinction cross sections of suspensions of spheres. was not possible. Hamamatsu Photonics Deutschland GmbH, Germany) In this work we present a method to infer the significantly drops. The slit width was set to obtain a wavelength-dependence of the real RI and the particle spectral resolution of 0.5 nm and the wavelength incre- size distribution, characterized by mean value and stan- ment amounted to 1 nm. Spectra were recorded with an dard deviation, from a single measurement of spectral integration time of 1 s and a scanning speed of 0.5 nm s−1. extinction cross section for spherical particles in the Mie Hence, the total measurement time was 18 min. scattering regime, i. e., the size is comparable to the The light emitted from the monochromator was colli- wavelength. We exemplify the method for suspensions mated by a spherical mirror of 20 cm focal length (M1) of microscopic synthetic polystyrene (PS) beads that are and reflected to the sample cell by a plane mirror (M2). widely used in colloidal and optical research, e. g., as a The system of apertures served to minimize the diver- calibration material for cell measurements in optical and gence of the light beam and to reduce stray light and impedance flow cytometry [17] or attached as “handles” ambient light. The adjustable apertures were set to a to optically manipulate biological cells [18]. From the diameter of 2 mm. Taking into account the distance of measurement of an extinction spectrum in the UV-VIS 90 cm between the first aperture and the aperture directly range with a tabletop experimental setup, one can infer mounted in front of the cuvette with the index matching the refractive index with an accuracy of 2 to 4 decimal fluid, the maximum cone of light that can enter the cu- places, depending in the wavelength, as well as the size vette is characterized by a half-angle of 0.13°. The sus- parameters with accuracies in the range of one nanome- pension of microspheres was pipetted in a quartz glass ter. cuvette (type 110-QS, Hellma GmbH & Co. KG, Ger- many) used for photometric absorption measurements. The path length of the cuvette is d = 10 mm with an II. MATERIALS, METHODS AND MODELS uncertainty of u(d) = 10 µm. To avoid the influence of slightly different angles when inserting the cuvette in the beam path, which would result in changes in back A. Experiment reflection and transmission, a container filled with in- dex matching oil (Immersol 518N, Carl Zeiss Microscopy 1. Experimental Setup GmbH, Germany) was permanently mounted. The size of the container (inner dimensions 20 mm × 30 mm) was To determine spectral extinction cross sections of chosen in such a way that the quartz glass cuvettes, fea- polystyrene microspheres we measured the collimated turing outer dimensions of 12.5 mm × 12.5 mm, could be transmission of particle suspensions using the setup de- easily changed. The long distance of 120 cm between the picted in Fig. 1. As light source a 1 kW Xenon high pres- cuvette and the detector, together with a system of ad- sure discharge lamp is mounted to a double monochroma- justable apertures set to 2 mm diameter, served to effec- tor (model 1680 double spectrometer, SPEX industries, tively suppress light scattered at small angles. The half- Inc., USA, NJ), equipped with a 1200 gr mm−1 grating observation angle amounted to 0.1°. As stated above, the for the wavelength range from 185 nm to 900 nm. The fo- transmitted light was recorded using a R928 photomul- cal length of the double monochromator is 22 cm. When tiplier tube. measuring the collimated transmission of particle suspen- sions, the scanning range was reduced to 260 nm–800 nm. Outside this wavelength region, the signal to noise ratio 2. Properties of polystyrene microspheres was insufficient, since below 260 nm the output power of the Xe high pressure lamp is too low and above 800 nm a. Diameter and size distribution To validate the the sensitivity of the photomultiplier tube (model R928, experimental procedure and the associated mathematical 3 model we measured the collimated transmission of two different types of PS microspheres. The mean diameter 1.7 of the first type (dynospheres™ 50.010.SS-021 P, LOT Q262, Dyno Particles A.S., Norway) – the results are de- nominated as dataset 1 in the following – was specified 1.6 as 2.0 µm. The stated coefficient of variation of 1.2% cor- ) λ responds to a standard deviation of 0.024 µm. The mass ( of PS amounted to 0.1 g in 10 mL. Taking into account n 1.5 PS [Nikolov, Ivanov (Appl. Opt. 2000)] PS Sellmeier equation the density of PS of 1.05 g mL−1 and the particle volume H2O [Daimon, Masumura (Appl. Opt. 2007)] of 4.19 fL the concentration was c = 2.3 × 109 mL−1. As second sample (dataset 2) we chose microspheres (Flow 1.4 Check™ microparticles, Cat. No. 23526-10, Polysciences Europe GmbH, Germany) with a slightly different diame- 1.3 ter of 2.076 µm and larger standard deviation of 0.053 µm 300 400 500 600 700 800 (CV = 2.6%). Since these particles are generally used for vacuum wavelength λ/nm calibration purposes in flow cytometry without diluting 7 −1 the sample, the concentration of about 10 mL is sig- Figure 2: Refractive index of water [2] and polystyrene [19]. nificantly lower compared to the other sample used. To create synthetic data, the data for PS was extrapolated b. Sedimentation The influence of sedimentation of from the measured range 436 nm–1052 nm to the near UV by the particles during the measurement time of 18 min was the Sellmeier equation shown as a blue solid line. estimated by calculating their velocity using Stokes’ law. In equilibrium, the frictional force and the excess force 75 due to the difference in density of water and PS are equal. −1 dataset 1 Using the density of PS ρPS = 1.05 g mL and the dy- −4 −1 −1 70 dataset 2 namic viscosity of water µH2O = 9.321 × 10 kg m s %

◦ / at 23 C the sinking velocity of PS particles with 2 µm di- )

−1 λ 65 ameter amounts to 0.12 µm s . It follows that the sedi- ( mentation path in 18 min is about 0.13 mm. This value is T negligible, since the light beam is positioned in the mid- 60 dle of the cuvette, the filling height of which is typically 2 cm, and the depletion zone as well as the enrichment 55 zone are in a distance of about 1 cm. transmittance c. RI of bulk polystyrene and water For the analy- 50 sis of the experimental spectral extinction cross sections initial values are needed for the wavelength dependence 45 of the refractive index of the PS microspheres. To this 300 400 500 600 700 800 end, we employ literature values for the RI of bulk PS vacuum wavelength λ/nm and water, shown in Fig. 2. The data for the PS RI from Ref. 19 can be fitted very well using a one-term Sellmeier Figure 3: Transmission spectrum of polystyrene microspheres equation containing a single absorption pole.[30] We used with a diameter of approximately 2 µm. this extrapolation curve to generate synthetic data for the spectral range [260, 800] nm. For the wavelength- dependent RI of pure water we used a four-term Sellmeier the concentration of particles is chosen correspondingly equation [2] that is accurate to a few 10−6 in the wave- and measured by a flow cytometer, designed to deter- length range 182 nm–1129 nm. The absorbance of water mine reference values for concentrations of blood cells in the wavelength range under consideration is very low and particles in suspensions [21, 22]. The mean free path and can thus be neglected, hence nH O ∈ R. The same 2 length between two scattering events was estimated by basically holds for PS, however, there is the onset of a taking two times the geometrical cross section as op- (weak) UV absorption line at 4.5 eV or 275 nm [20]. Nev- tical extinction cross section. For a transmittance of, ertheless, we treat its RI as real-valued. e. g., 70% the mean free path length of 30 mm exceeds the cuvette path length by a factor of 3. For the sam- ples the dilution was adjusted to yield concentrations c 3. Sample preparation and transmittance measurements of (4.95 ± 0.09) × 106 mL−1 and (5.23 ± 0.10) × 106 mL−1 for datasets 1 and 2, respectively. In Fig. 3, we depict the The mathematical model for the analysis of extinc- result of measurements of the spectral transmission for tion spectra requires that only single scatter events occur the two datasets. These spectra represent the transmit- when the light passes the cuvette. Multiple scattering tance of the particle suspensions, since blank values were is not included. To ensure that this condition is met, corrected. To this end, we first measured the transmit- 4 tance I0(λ) of the cuvette with pure water. Subsequently, the cuvette was cleaned and filled with the particle sus- repeatability 10% local fluctuations pension. The transmittance I(λ) of the suspension was model

measured and the spectral transmittance derived as ratio ) λ (

between both measurements: ∗ C I λ /

( ) )]

T (λ) = . (1) λ ( I0(λ) ∗ 1% C

It should be noted that all measurements were normal- std[ ized to the spectral output power of the monochromator using the quantum absorber [23] integrated in the instru- ment. The spectral transmittance can also be expressed as 0.1% 300 400 500 600 700 800  Censemble  vacuum wavelength λ/nm T = exp − ext , (2) Ω ∗ Figure 4: Estimates of the coefficient of variation σˆ[C (λ)] ensemble where Ω is the area illuminated by the beam. Cext is due to measurement noise: Maximal estimate from repeated measurements (black dots), minimal estimate from local fluc- the extinction cross section of the particle ensemble. The ∗ particles in the sample are specified by the manufacturer tuations of C (λ) (red circles) and model curve [Eq. (6), blue as monodisperse, since generally the variation in size is line]. negligible for the intended purpose, e. g., calibration of flow cytometers. However, spectral extinction is sensitive to the size variation and consequently we have to consider We assume the detector noise to be white (i. e., un- correlated between different wavelengths), to have a the particles as polydisperse. Because we are dealing with ∗ incoherent single scattering events, the total extinction wavelength-dependent amplitude std[C (λ)], and a coef- cross section of the ensemble of ν particles is the sum of ficient of variation the single particles’ extinction cross sections ∗ ∗ std[C (λ)] σ C λ . ν ˆ[ ( )] = ∗ (5) ensemble X j C (λ) Cext = Cext. (3) ∗ j=1 We estimated σˆ[C (λ)] from the data in two different ways: With the optical path length of d = 10 mm, the par- ticle concentration of c ≈ 5 × 106 mL−1 and the area 1. From the variation of repeated measurements for Ω = 3.1 mm2 given by a 2 mm aperture, the number of the same sample in overlapping spectral regions, particles ν = Ω d c is on the order of 105. This allows for which yields an upper bound of the statistical the measurement of an ensemble average, denoted by measurement error, because it also contains sys-   tematic contributions resulting from drift of mea- 1 ensemble I 1 surement conditions and other sources. The es- Cext = Cext = − ln . (4) ∗ ν I0 d c timated σˆ[C (λ)] has a characteristic “bathtub shape” (Fig. 4), which we approximate by a func- d The path length is known to a relative accuracy of tion 10−3 and the particle concentration c was measured us- ∗ p1 ing flow cytometry to a relative standard deviation of σˆ[C (λ)] = + p4 λ + p5. (6) p 2 2 2%. The latter uncertainty contribution thus introduces p2 − (λ − p3) an uncertainty into the scaling factor for the measured with five fit parameters. Such a shape indicates ensemble average in Eq. (4). In the following, we drop increased noise contributions due to low output the subscript ext, i. e., Cext = C and denote experimental ∗ power of the light source in the UV and a decreas- data by an asterisk: C . ing sensitivity of the detector for long wavelengths.

2. From the local fluctuations of Cext in regions of 4. Noise model for cross section measurements vanishing slope, which yields an estimate of the measurement error. Due to the pronounced ripple There are different sources of measurement uncer- structure of the data, no estimate is possible at the tainty: (1) concentration uncertainties as discussed blue end of the spectrum and the bathtub shape above, (2) detector noise, and (3) possible other system- is not resolved. We use the same model function atic influences yet to be determined. The latter point is [Eq. (6)] as for the estimate based on the variation discussed in subsection II B 3. of repeated measurements above and simply rescale 5

it by a factor < 1 to optimally fit the minimal er- 2. Polydisperse ensembles ror estimate. The resulting curve is shown in Fig. 4 and is used for weighting and uncertainty analysis Since the number of particles measured simultaneously in the following. in the experiment is large, the ensemble average in Eq. (4) is modeled by simply integrating over the corresponding size distribution B. Mathematical model Z ∞ C(λ; n|r) = C(λ; n,R) r(R) dR, (7) 1. Mie scattering 0

where r(R) is the probability density function (pdf) of The scattering particles are assumed to be spheri- the radius R. We model it by a normal distribution cal and multiple scattering effects are negligible due to the low particle concentration and low sample thickness.  2  1 [R/Rc − µR] Mathematically, the experiment can thus be described r(R) ∝ exp − . (8) 2 σ2 using the time-harmonic Maxwell equations for the scat- R tering of a plane electromagnetic wave by a single homo- R geneous sphere. This problem has an analytic solution, Here c is a typical length scale (e. g., a rough guess which is known as Mie theory or Mie scattering [12, 13]. for the mean particle radius), rescaling the parameters µ Mie scattering yields the full electric field for the scat- to dimensionless quantities of the order of 1. R is the σ tering problem as a series expansion in vector spherical mean and R the standard deviation of the distribution of R harmonics, from which, among other quantities, the to- radii, relative to the characteristic scale c. We combine these two parameters of the distribution into the vector tal extinction cross section Cext can be computed. The problem is fully characterized by two parameters: T θ := (µR, σR) . (9) 1. The size parameter A normal distribution was chosen as a model func- 2π nm tion since it is known to describe the size distribution X := R, λ of polystyrene microspheres very well [25]. Alternatively, we also tested a log-normal distribution. For the narrow where R is the radius of the sphere, λ is the wave- size distributions of the samples used here, the difference length in vacuo and nm ∈ R is the RI of the (non- in shape between a normal distribution an a log-normal absorbing) medium (water), respectively. distribution of identical mean value and standard devia- tion is only minor. Consequently, the results were virtu- 2. The relative refractive index ally the same in both cases. n To implement the integrals in the ensemble average m := ∈ C, numerically, we used trapezoidal sums over uniform grids, nm with appropriate spacing and range. where n = n + iκ ∈ C is the RI of the sphere. The imaginary part κ describes the absorption of light in the sphere’s material. 3. Correction for systematic influences

Consequently, besides the known quantities vacuum Including concentration errors in the forward model wavelength λ and RI of the medium nm(λ), the two quan- for the measured curve would result in the ensemble av- tities sphere radius R and sphere’s RI n(λ) are the two eraged C(λ; n|r) being multiplied with a prefactor 1 − c, free parameters of the system. where c corresponds to the relative concentration error Using a numerical implementation of the Mie scatter- with E(c) = 0, std(c) ≈ 2%. However, it turns out that ing formulae [24], accurate values for the extinction cross the forward model is much more flexible and the qual- section of a single particle are easily obtained for any ity of the solution of the inverse problem can be greatly given parameters n(λ) and R in the relevant range. We enhanced when a simple parametric wavelength depen- denote these numerically obtained values by the function dence of this prefactor is introduced, i. e., instead of 1−c, we have 1 − c + o(λ; η), where o(λ; η) corrects for in- C(λ; n,R). fluences other than the concentration. Mathematically, the influences of concentration and other factors can not This notation shall indicate that we are considering a be distinguished, hence we merge them into a single wavelength-dependent function with two parameters, one wavelength-dependent function (λ; η) = −c + o(λ; η). of which, the RI n(λ), is a function of the wavelength This compensation curve is characterized by the param- T itself. eters η = (η1, . . . , ηmη ) and defined as a linear interpo- 6 lation between a few grid points affect the result of the RI reconstruction and Eq. (12)  may not even have solutions any more for some wave- ηj λ = lj lengths. Since the prior knowledge about the parameters  linearly interpolated between the lj of the size distribution is not accurate enough, they need (λ; η) := , η λ < l to be inferred from the data as well. At first glance, this  1 1 η λ > l leaves one with more parameters to be reconstructed than Mη Mη data points. But the problem can be restated in a non- (10) T pointwise sense as a least-squares optimization problem. with l1 < ··· < lMη , η = (η1, . . . , ηMη ) . A small num- ber of points (Mη = 4 in our case) is sufficient to signifi- cantly improve the quality of the fit compared to the one 1. Representation of n(λ) obtained with a constant function. Hence, our model for ∗ the measured C (λj) is We restate the problem by implying a constraint on the admissible functions n, namely that they can be ex- M(λ; n, θ, η) := [1 + (λ; η)] C(λ; n|r). (11) pressed as a finite sum of smooth basis functions gj, i. e.,

From a physical point of view, this wavelength-dependent M factor serves to correct for one or several influences, which X n(λ) = ajgj(λ) ∀ λ ∈ R (13) are difficult to quantify here, in particular aspects such j=1 as non-sphericity as well as the finite-size detector aper- ture and the beam divergence. From a mathematical with unknown coefficients aj. Hence, we are working point of view, the introduction of (λ; η) is justified by in the subspace spanned by the M basis functions with the much higher quality of the fit. As will be discussed M < N. in section III, we obtain more convincing results for the Simple oscillator models for light-matter interactions fitted RI and size distribution using this correction. reveal that particles with a resonance at wavelength Λ > l , . . . , l For a choice of the grid points 1 Mη we tested 0 and of width γ > 0 exhibit an anomalous dispersion uniform grids as well as a random placement using Mη = feature described by the expression 0..10 points. However, the best results (measured by the quality of the fit) were obtained when the local minima λ − Λ ∗ n(λ) − 1 ∝ , (14) and maxima of the slow oscillations of C (λ) were se- (λ − Λ)2 + γ2 lected, which are at 300 nm, 350 nm, 450 nm and 770 nm for 2 µm PS spheres in water (Fig. 6). which is thus a generic shape of a feature of the real part of the RI. Hence, we represent the refractive index n(λ) in a sparse form by working in the M-dimensional space C. Inverse problem

n ∈ { , f , . . . , f }, We now have a mathematical model for the mea- span 1 1 M−1 (15) surement and are able to compute the average extinc- where tion cross section for a given ensemble of spheres, i. e., for a given size distribution and RI n(λ). We assume λ − y f λ j . n λ < n λ n λ j( ) := 2 2 (16) ( ) = [ ( )] = ( ) here and in the following due to (λ − yj) + γ polystyrene’s negligible absorbance. The question we now address is: Can one infer the The centers of the peaks yj are uniformly spaced, i. e., h iN ∗ yj − yj−1 = ∆y. We chose a grid spacing of ∆y = 30 nm RI n(λ) from measurements C (λj) ? First of all, j=1 and a constant width of γ = 100 nm for all functions. it is clear that one should not try to infer more than The first (last) grid point was set to be one grid spacing N parameters from N given measurement data. This ∆y smaller (larger) than the lowest (highest) wavelength. leaves the possibility to obtain n(λj) for all λj, j = 1..N, These functions fi are linearly independent, but they given knowledge about the size distribution and particle are not useful for a practical implementation, since they concentration, which is a relatively simple problem: At are far from being orthogonal, i. e., the scalar products each wavelength λj, find nj = n(λj), such that between them are not close to zero, which would lead to problems in the numerics. To avoid this problem, ! ∗ C(λj; nj|r) = C (λj). (12) we orthonormalize the set of functions {1, f1, . . . , fM−1}, or rather the set the values of these functions at The problem is hence stated as finding roots of a non- (λ1, . . . , λN ), using the (modified) Gram-Schmidt pro- linear equation for all wavelengths separately. Hence, cess, yielding an orthonormal set {g1, . . . , gM }, i. e., this approach is pointwise. However, as it turns out, the pointwise approach fails: Slight, sub-percent errors in the 1 hgi, gji = δij and g1(λj) = √ ∀ j = 1, .., N, (17) size distribution or in the particle concentration strongly N 7

sections M(λ; ψ) from the parameter vector ψ is de- picted in Fig. 5. We then set up a quadratic cost function consisting of the summed squared residuals

∗ Fi := M(λi; ψ) − C (λi) (20) N 2 X T χ (ψ) := wijFiFj = F W F , (21) i,j=1

∗ where W = V(C )−1 is a symmetric weight matrix with

∗ 2 2 V(C ) = diag(σ1, . . . , σN ). (22)

∗ The standard deviations σi = std[C (λi)] are estimated according to Eq. (6) for each dataset. The necessary con- dition for optimal parameters that minimize χ2 is then

Figure 5: Schematic representation of the mathematical model of the experiment. Linear transformations are shown ! 2 T 2 T 0 = ∇ψχ (ψ) = J ∇F χ (ψ) = 2 J WF , (23) as solid lines, nonlinear dependencies as dashed lines. where we have introduced the Jacobian matrix h.,.i RN   where denotes the standard inner product in . ∂Fi N×L This representation of n(λj) as a series expansion in J = ∈ R . (24) ∂ψj adequately-chosen basis vectors results in a reduction of ij the dimensionality by a factor of 30: Instead of one co- Further details for the implementation of the forward efficient every 1 nm corresponding to the spectral reso- model are given in Appendix A. Numerically, the nec- lution, only one coefficient every 30 nm is needed. At essary condition can be solved by iterative local op- the same time, it is possible to represent the RI of bulk timization methods. We tested both the Levenberg- PS (Fig. 2) to machine precision. But also curves less Marquardt algorithm and the trust-region-reflective al- generic than this Sellmeier equation (see footnote [30]), gorithm as they are implemented in Matlab (MATLAB e. g., the feature-rich spectral RI of the protein complex R2016a, The MathWorks, Inc.). Both algorithms basi- hemoglobin [26, 27] can be represented with errors well cally yield the same result when converged. Since the below the respective measurement uncertainties using an trust-region-reflective algorithm proved more stable with appropriate grid spacing ∆y. respect to the choice of initial values, all the results pre- sented here were obtained with this algorithm. When converged, the numerical routine yield an optimal pa- 2. Nonlinear least-squares optimization rameter set ψˆ corresponding to a local minimum and χ2 normalized to the degrees of freedom The series expansion of the real RI in Eq. (13) can also be written as a matrix-vector product χ2(ψˆ) χ2 := (25) dof N − L n = G a (18) + 1

N×M provides a measure for the quality of the fit. with G := {gj(λi)}i,j ∈ R , n = T a. Influence of initial parameter values The local (n(λ1), . . . , n(λN )) . The coefficient vector a de- nonlinear optimization employed for the same dataset scribing the RI, the vector θ describing the size can result in different estimates of the parameter vec- distribution of the spheres and vector η describing the tor ψ when different initial values are used. Firstly, it is compensation curve are the unknown quantities which possible that the found local minimum is far away from we need to recover from the spectral extinction data. true parameter values (for synthetic data) or plausible Thus we combine them into a single parameter vector parameter values (for experimental data). This is usu-   2 2 a ally indicated by a large value of χdof (i. e., χdof  1). L ψ := θ ∈ R , (19) Secondly, one can also find different local minima rather η close to each other and – in the case of synthetic data – similarly close to the correct value, indicating multiple where L = 27 with the parameters chosen for measure- minima of the least-squares problem. For these minima ment data discussed here (N = 540 points). The math- further iteration does not improve the fit. We estimated ematical forward model to compute the extinction cross the resulting uncertainty stemming from this effect as 8 follows: Different initial parameter values were created found size distribution parameters were within the mar- by adding normally distributed random numbers to the gins expected from the corresponding estimated uncer- same mean initial parameter vector. We tuned the am- tainties. We computed the difference between the RI val- plitude of these random numbers such that the range ues n(λj) obtained by optimization and those according of initial conditions is sufficiently wide and that, on the to the target curve. The percentages of values within one other hand, the optimization converged for the major- (two) standard deviation(s) estimated from the uncer- 2 ity of samples. However, very high values of χdof do tainty analysis were found to be close to the 68% (95%) occur in some cases. In order to include all of these sam- one expects from a normal distribution, indicating a rea- ples appropriately, we introduced weights proportional sonable estimate of the uncertainties. Generally, these 2 to exp(−χdof ) in the averages and covariances over the percentages were even higher, indicating that the uncer- ensemble of optimization runs. This penalty term atten- tainties are slightly overestimated. The compensation uates samples with poor agreement between model and curve also returned the target (η = 0) within the uncer- 2 data (e. g., χdof = 10). The result for the parameter vec- tainties. tor is then obtained as the weighted average hψiinit over the ensemble of random initial parameter values. The experimental cross sections, i. e., datasets 1 and 2 The uncertainty analysis for the estimated parameters (Fig. 6), were analyzed in the same way. The resulting is presented in Appendix B. RI is depicted in Fig. 7 (a) and the corresponding un- certainty (Fig. 7 (b)) is shown to be less than 1% almost everywhere and to be less than 0.1% between 300 nm and III. RESULTS 650 nm. The compensation curve is shown in Fig. 7 (c). The wavelength-dependence of (λ; η) is significant with the estimated uncertainties. Tab. I lists the mean val- ues and standard deviations of the particle size distribu- tions. All these results correspond to the MC-averaged optimized parameters.

For wavelengths larger than 436 nm, we compared the RI uncertainties with the deviations between the found RI curves and the Sellmeier curve for the literature RI of bulk PS (see footnote [30]). For the first dataset 52% (65%) of the values were within one (two) estimated stan- dard deviations of the literature values. For the second dataset 37% (57%) were within one (two) standard de- viations. Although these percentages are somewhat too low, it indicates that the estimated uncertainties are not too far off. Furthermore, these numbers do not take into account any uncertainties of the RI literature values or possible deviations of the RI of PS microspheres in our samples from that of bulk PS. Figure 6: Experimental extinction cross sections (“target”) and model fits with optimized parameters as described in sub- As an additional consistency test, we computed the sections II B and II C. See Data Files 1 and 2 for underlying difference between the RI curves obtained independently values of the respective datasets. from the two experimental datasets. This difference was within the combined estimated uncertainties, thus indi- cating that the uncertainty estimates are reliable.

A. Inverse problem results

To evaluate the method, we created synthetic datasets Table I: Optimization results for size distribution parameters, for polystyrene particles suspended in pure water. These cf. Fig. 7. The numbers in parentheses are the standard un- synthetic data were created as follows: The optical prop- certainties referred to the last digits of the respective results. erties assumed are shown in Fig. 2. The compensation mean(R)/nm std(R)/nm curve was fixed to (λ; η) ≡ 0 in the forward model and dataset 1 Gaussian white noise was added according to the coef- result 1015.1(8) 4.7(2) ficient of variation estimated in Eq. (6) and shown in specification 1000 12 Fig. 4. We analyzed synthetic data with mean(R) = dataset 2 1 µm, std(R) = 5 nm for different realizations of mea- result 1038.5(9) 4.4(2) surement noise and observed that the deviations of the specification 1038 27 9

B. Effect of the compensation curve

In Fig. 8 (a) we plot the local extinction-cross-section ∗ residuals F (λj; ψ) = M(λj; ψ) − C (λj) as a function of wavelength and refractive index. We performed an addi- tional optimization with a forward model not including the compensation curve, i. e., η ≡ 0 in the optimization. The result is shown in Fig. 8 (b). Comparing Figs. 8 (a) and (b), ones sees how the incorporation of (λ; η) into the model enables it to close the gaps that are other- ∗ wise present near the local extrema of C (λ) (compare ∗ Fig. 6). The ripples in C (λ) are not matched by the model fit (data not shown) and for dataset 1 we ob- 2 tained hχdof i = 7.3 (indicating a poor fit) instead of (a) 2 hχdof i = 0.78 (indicating a very good fit) with the for- ward model including (λ; η). For dataset 2 these values 2 2 were hχdof i = 10.6 and hχdof i = 1.57, respectively. Fur- thermore, the curves for n(λ) obtained for datasets 1 and 2 do not match each other within the estimated uncer- tainties in the former case and neither of them matches the literature data. In conclusion, the optimization with the forward model including (λ; η) (i. e., in our exam- ple with L = 27 free parameters instead of L = 23) yield much more consistent and plausible results than with the one not including (λ; η) and thus was chosen even though the physical origin of the effect cannot be identified unambiguously.

C. Reliability of solutions for wider ensembles (b)

The procedure for solving the inverse problem and related uncertainty analysis works well for the cases discussed here with mean(R) ≈ 1 µm and std(R)/mean(R) ≈ 0.5%, i. e., for a narrow size distri- bution. Using synthetic data, however, we found that the parameter retrieval becomes increasingly difficult for wider size distributions. This is because the characteris- tic ripples in C(λ) get smoothed out for wider size distri- butions. This ripple structure is very sensitive to parti- cle size and RI [16]. Hence, when this characteristic fine structure is suppressed by ensemble averaging essential information is lost. The effects of RI, mean radius and particle concentration in the model can then compensate each other, which can lead to a good fit of C(λ) with systematically incorrect parameters. To quantify this, we analyzed synthetic data with mean(R) = 1 µm and (c) distribution widths of 0.1% ≤ std(R)/mean(R) ≤ 5.0%. Without added measurement noise, the correct parame- Figure 7: Optimization results for the model parameters for ters can be found in all cases, i. e., this is not a problem 2 µm PS spheres for two experimental datasets: (a) Refrac- of (numerical) sensitivity in the forward model. In the tive index. Literature values are from Ref. 19. The oscil- presence of measurement noise, however, systematic pa- n λ lations of ( ) at the red and UV end of the spectrum are R / R ≥ within the estimated uncertainties (Fig. 7 (b)). See Data Files rameter errors occur for values of std( ) mean( ) 1%, 1 and 2 for underlying values of the respective datasets. (b) because in these cases the sensitivity is obscured by mea- Estimated uncertainties (one standard deviation) of the RI surement noise. These deviations typically increase with (Fig. 7 (a)). (c) Compensation curve. The piecewise linear distribution width. A similar effect can also be expected function (λ; η) is spanned by 4 grid points at 300 nm, 350 nm, for low-contrast particles, because this attenuates the rip- 450 nm and 770 nm. 10

low-dimensional set of orthonormal basis functions, we 4 developed a forward model M(λ; ψ) to mathematically describe the measurements for a given parameter set ψ. 2 Applying standard nonlinear least-squares optimization, the L model parameters contained in the vector ψ were 0 fitted to N measurement data points. In our example, we had N = 540 and L = 27. This yields the spectral -2 refractive index of the microspheres, the mean value of their size and the standard deviation of their size distri- -4 bution. The uncertainties of these results were estimated using linearized propagation of covariance matrices and -6 Monte Carlo sampling. The wavelength dependence of the RI of PS micro- spheres derived by modeling experimental extinction (a) cross sections including uncertainties yields results con- sistent with literature values. This good accordance proves that the microspheres are produced with a homo- 4 geneity in density and optical properties that is compa- rable to bulk material. On the other hand, both samples 2 feature about the same size distribution with a coeffi- cient of variation of about 0.5%, which is significantly 0 lower than the specification of the vendors of 1.2% and 2.6%, respectively. It follows that the particles are closer -2 to monodispersity than declared, presumably since the specified distribution widths are estimates of the corre- -4 sponding upper limit during production.

-6 The method for measurement and data analysis pre- sented here allows to determine the mean diameter of an ensemble of 2 µm PS spheres with an uncertainty of 2 nm or a relative uncertainty of 10−3. This uncertainty was estimated from the measurement noise and the limited (b) precision of the inverse problem’s solution and verified ∗ with synthetic data. Model errors might have an ad- F λ ψ M λ ψ −C λ Figure 8: Dataset 1: Residuals ( j ; ) = ( j ; ) ( j ) ditional effect not analyzed here, e. g., deviations from of the extinction cross section as a function of wavelength surface roughness and assumed refractive index. The parameters θ and η are a spherical shape or . However, since fixed to the optimization result. (a) With forward model even the most intricate details of the extinction spec- incorporating the compensation curve (λ; η) (Fig. 5). The tra, i. e., their ripple structure, can be fitted with the black dotted lines denote the wavelength grid spanning (λ; η) model, it seems unlikely that surface roughness is rele- (Fig. 7 (c)). Note the false solutions intersecting at the local vant. The scattering properties from ensembles of rough ∗ extrema of C (λ) (Fig. 6) and the blurred valley at the red spheres have been examined [28, 29] and it was found end due to measurement noise. (b) With forward model not that the deviations between spheres and rough spheres incorporating the compensation curve (λ; η). are largest for side scattering and backscattering ampli- tudes, whereas forward and near-forward scattering is least sensitive to irregularities. Due to the optical the- ple structure as well. This problem could in principle orem [13], this means that the extinction cross section be resolved by achieving a sufficiently low measurement Cext is very insensitive to irregularities of the particles’ noise. surfaces. Hence the Mie scattering formulae may be used even for somewhat irregular particles. We have demonstrated the potential of spectral ex- IV. SUMMARY AND DISCUSSION tinction measurements in combination with an adapted mathematical model to determine microparticle proper- We presented measurements of the spectral extinction ties. To further improve the model, in particular for new cross sections of polystyrene microspheres suspended in applications, the investigations will be extended to par- ∗ water. Theses measurements C (λ) represent averages ticles in the size range of several hundred nanometers up over the size distribution of the polydisperse ensemble. to about 10 µm and to other (non-absorbing and absorb- Using a numerical implementation of Mie scattering, in ing) materials. In addition, comparison with complemen- combination with an appropriate description of the size tary methods used to determine size and size distribu- distribution and the spectral refractive index n(λ) in a tion, i. e., dynamic light scatter, secondary electron mi- 11 croscopy, nanoparticle tracking and sedimentation anal- now analyze the uncertainties of the resulting parameters yses will be carried out. obtained by nonlinear optimization. There are two types Potential applications of our spectral extinction mea- of uncertainty contributions: surement and analysis method include quality assurance 1. A contribution due to the measurement uncertain- to monitor size, size distribution and composition (i. e., ∗ spectral RI) when producing suspensions of nano- and ties of the experimental data C (λ). Here we focus micro-particles or emulsions. The method is also sen- on detector noise since concentration uncertainties sitive towards the homogeneity of the particles (e. g., and the compensation curve are included explicitly quartz spheres) and allows to identify irregularities by in the forward model. We will denote the terms comparing the spectral RI of the particles to the RI of related to the measurement uncertainty by a su- the bulk material. Complementarily, the roles of medium perscript “meas”. and particle RI can be interchanged in the optimiza- 2. A contribution from the local least-squares opti- tion. In this way, the spectral RI of the solvent, e. g., mization, denoted by a superscript “init”. It de- a protein solution, can be deduced using (monodisperse) scribes the uncertainty arising from the ambiguity micro-spheres with known optical and morphological pa- of the solution when starting the parameter opti- rameters as a probe. mization from different initial values. We estimated both contributions using the direct sam- Appendix A: Expressions for the numerical pling Monte Carlo (MC) method as described in the fol- implementation of optimization lowing.

The ensemble-averages with trapezoidal sums replac- ing integrals read 1. Propagation of measurement uncertainties

I X The contribution of measurement uncertainties is es- C(λi; ψ) = C(λi; ni(a),Rt)r ˜t(θ). (A1) t=1 timated by linearized covariance-matrix propagation ac- cording to For the Jacobian matrix one finds −1 meas h T meas ∗ −1 i ( ∂C(λi;a,θ) V (ψ) = J V (C ) J . (B1) ∂F [1 + (λ; η)] j = 1..L − Mη i ∂ψj Jij = = ∂(λ;η) , ∂ψj C(λi; a, θ) else ∂ψj To verify the validity of this linearization of the for- (A2) ward model, we performed MC simulations for synthetic data: A number NMC of synthetic datasets was cre- with ated by adding independent random realizations of the measurement noise [Eq. (6)] to the same synthetic curve I   ∂C(λi; a, θ) X ∂ ∂ni (mean(R) = 1 µm, std(R) = 5 nm,  ≡ 0). For each of = C(λi; ni,Rt) r˜t ∂a ∂n ∂a j t=1 i j these random datasets, the inverse problem was solved numerically using the same initial parameter values. The I   X ∂ covariance matrix of the optimized parameter vector ψ = C(λi; ni,Rt) r˜t Gij (A3) ∂n was estimated for each of these numerical solutions ac- t=1 i cording to Eq. (B1) and an average was taken over the I meas ∂C(λi; a, θ) X ∂ NMC samples. In addition, we estimated V (ψ) using = C(λi; ni(a),Rt) r˜t(θ) (A4) ∂θ ∂θ the statistical sample covariance over the set of MC sam- j t=1 j ples. In this setup, the two independent estimates are ∂ A partial derivative like C(λi; ni,R) can be computed consistent, thus legitimating the linearized propagation ∂ni numerically by finite differences at the same computa- of measurement uncertainties, also for experimental data ∂(λ;η) with similar parameters. tional complexity as C(λi; ni,R). The derivatives ∂ηj are trivial because of the linear dependence of  on η. The Jacobian matrix J is thus implemented efficiently 2. Uncertainties from initial values using the above equations and was used explicitly in the numerical optimization. As described in subsection II C 2 the result for the pa- rameter vector is estimated as the weighted average from Appendix B: Uncertainty analysis Ninit optimization runs with random initial values

N ∗ init C λ 1 X (i) The sources of measurement uncertainties of ( ) and hψiinit := βi ψ (B2) Z estimates thereof were given in subsection II A 4. We 1 i=1 12

 2 (i) with weights βi := exp −χdof ψ and Zs := a. Initial parameter values For analyzing 2 µm PS P s (i) ± i βi , s ∈ N. ψ is the optimization result of the ith particles, the initial values were chosen as follows (mean run. From these statistical ensembles, we further esti- 1std of Gaussian white noise): mated the covariance 1. The RI was initialized as a piecewise-linear function init V (hψiinit) := spanned over the points n[(260, 460, 800) nm] = N . , . , . init T (1 75 1 605 1 585), which is a crude approximation Z2 1 X     β ψ(i) − hψi ψ(i) − hψi , of the Sellmeier curve (Fig. 2). For the MC sam- 2 Z i init init Z1 − Z2 1 i=1 pling, the coefficient vector varied a with a stan- (B3) dard deviation of 0.6 for a1 and 0.04 for aj, j = 2, ..., M, respectively. which can be applied to both, experimental and synthetic data (keeping the realization of measurement noise iden- 2. The mean radius mean(R) was initialized to tical in all samples). The total parameter variance cor- 1000 nm ±20 nm and the distribution width std(R) responding to the combined uncertainty of measurement to 10 nm ± 4 nm. and data analysis is then

meas init 3. The compensation curve was initialized with ηj = V(hψiinit) = hV (ψ)i + V (hψiinit). (B4) init 0.00 ± 0.05. For a sufficiently large number of samples,√ the “initial” init term will ultimately scale like V ∝ 1/ Ninit. Using Ninit = 50, the computing time was a few minutes on a year 2014 desktop PC and Vinit indeed became almost negligible. Acknowledgements From the parameter covariance V(hψiinit) we can easily extract the uncertainty of the spectral refractive index n = G a as The authors thank Volker Ost and Uwe Sukowski for their support when designing the experiment, assembling T V(hniinit) = GV(haiinit) G , (B5) the experimental setup and performing measurements. We would also like to thank Sebastian Heidenreich and where V(haiinit) corresponds to the first M × M entries Hermann Groß for helpful discussions. of V(hψiinit), and analogously for the other resulting pa- rameters.

[1] M. Daimon and A. Masumura, Appl. Opt. 41, 5275 [9] B. N. Khlebtsov, V. A. Khanadeev, and N. G. (2002), URL http://ao.osa.org/abstract.cfm?URI= Khlebtsov, Langmuir 24, 8964 (2008), pMID: 18590302, ao-41-25-5275. http://dx.doi.org/10.1021/la8010053, URL http://dx. [2] M. Daimon and A. Masumura, Appl. Opt. 46, 3811 doi.org/10.1021/la8010053. (2007). [10] B. , Y. Roichman, and D. G. Grier, Opt. Express [3] M. R. Jones, B. P. Curry, M. Q. Brewster, and K. H. 16, 15765 (2008), URL http://www.opticsexpress. Leong, Appl. Opt. 33, 4025 (1994), URL http://ao. org/abstract.cfm?URI=oe-16-20-15765. osa.org/abstract.cfm?URI=ao-33-18-4025. [11] A. B. Stilgoe, T. A. Nieminen, G. Knöner, N. R. Heck- [4] M. R. Jones, K. H. Leong, M. Q. Brewster, and B. P. enberg, and H. Rubinsztein-Dunlop, Opt. Express 16, Curry, Appl. Opt. 33, 4035 (1994), URL http://ao.osa. 15039 (2008), URL http://www.opticsexpress.org/ org/abstract.cfm?URI=ao-33-18-4035. abstract.cfm?URI=oe-16-19-15039. [5] G. E. Thomas, S. F. Bass, R. G. Grainger, and A. Lam- [12] G. Mie, Annalen der Physik 330, 377 (1908). bert, Appl. Opt. 44, 1332 (2005), URL http://ao.osa. [13] C. F. Bohren and D. R. Huffman, Absorption and Scat- org/abstract.cfm?URI=ao-44-7-1332. tering of Light by Small Particles (Wiley, 1983). [6] J. D. Felske and J. C. Ku, Combustion and [14] A. P. Nefedov, O. F. Petrov, and O. S. Vaulina, Flame 91, 1 (1992), ISSN 0010-2180, URL Appl. Opt. 36, 1357 (1997), URL http://ao.osa.org/ http://www.sciencedirect.com/science/article/ abstract.cfm?URI=ao-36-6-1357. pii/0010218092901237. [15] Y. Zhang, Y. Zhang, X. Han, P. Tuersun, and K. F. [7] E. K. Naumenko, T. V. Oleinik, and A. Y. Khairullina, Ren, Procedia Engineering 102, 315 (2015), ISSN 1877- Optics and Spectroscopy 53, 288 (1982). 7058, new Paradigm of Particle Science and Technology [8] X. Ma, J. Q. Lu, R. S. Brock, K. M. Jacobs, P. Yang, Proceedings of The 7th World Congress on Particle Tech- and X.-H. Hu, Physics in Medicine and Biology 48, 4165 nology, URL http://www.sciencedirect.com/science/ (2003), URL http://stacks.iop.org/0031-9155/48/i= article/pii/S1877705815001496. 24/a=013. [16] R. Blümel, M. Bağcioğlu, R. Lukacs, and A. Kohler, J. 13

Opt. Soc. Am. A 33, 1687 (2016), URL http://josaa. (2010), URL http://stacks.iop.org/0957-0233/21/i= osa.org/abstract.cfm?URI=josaa-33-9-1687. 7/a=074006. [17] M. Frankowski, P. Simon, N. Bock, A. El-Hasni, [23] W. H. Melhuish, J. Res. Nat. Bur. Stand. USA 76, 547 U. Schnakenberg, and J. Neukammer, Engineering in (1972). Life Sciences 15, 286 (2015), ISSN 1618-2863, URL [24] For numerical Mie computations we used the http://dx.doi.org/10.1002/elsc.201400078. 77 code by W. J. Wiscombe (wis- [18] K. Svoboda and S. M. Block, Annual Re- [email protected]), NASA Goddard Space view of Biophysics and Biomolecular Struc- Flight Center and wrapped it in Matlab. ture 23, 247 (1994), pMID: 7919782, [25] T. R. Lettieri, A. W. Hartman, G. G. Hembree, and https://doi.org/10.1146/annurev.bb.23.060194.001335, E. Marx, Journal of research of the National Institute URL https://doi.org/10.1146/annurev.bb.23. of Standards and Technology 96, 669 (1991). 060194.001335. [26] M. Friebel and M. Meinke, Appl. Opt. 45, 2838 (2006). [19] I. D. Nikolov and C. D. Ivanov, Appl. Opt. 39, 2067 [27] J. Gienger, H. Groß, J. Neukammer, and M. Bär, (2000), URL http://ao.osa.org/abstract.cfm?URI= Appl. Opt. 55, 8951 (2016), URL http://ao.osa.org/ ao-39-13-2067. abstract.cfm?URI=ao-55-31-8951. [20] T. Inagaki, E. T. Arakawa, R. N. Hamm, and M. W. [28] R. Schiffer, J. Opt. Soc. Am. A 6, 385 (1989), URL http: Williams, Phys. Rev. B 15, 3243 (1977), URL https: //josaa.osa.org/abstract.cfm?URI=josaa-6-3-385. //link.aps.org/doi/10.1103/PhysRevB.15.3243. [29] R. Schiffer, Appl. Opt. 29, 1536 (1990), URL http:// [21] V. Ost, J. Neukammer, and H. Rinneberg, Cytometry ao.osa.org/abstract.cfm?URI=ao-29-10-1536. 32, 191 (1998), ISSN 1097-0320, URL http://dx.doi. 2 B λ2 [30] The one-term Sellmeier equation reads n(λ) = 1+ λ2−C org/10.1002/(SICI)1097-0320(19980701)32:3<191:: and fits the literature data for PS [19] with B = 1.4432 AID-CYTO5>3.0.CO;2-N. and C = (142.1 nm)2. [22] S. Reitz, A. Kummrow, M. Kammel, and J. Neukam- mer, Measurement Science and Technology 21, 074006