research papers\(\def\hfill{\hskip 5em}\def\hfil{\hskip 3em}\def\eqno#1{\hfil {#1}}\)

Journal logoJOURNAL OF
APPLIED
CRYSTALLOGRAPHY
ISSN: 1600-5767

Measurable structure factors of dense dispersions containing polydisperse optically inhomogeneous particles

crossmark logo

aInstitut für Chemie, Universität Rostock, 18051 Rostock, Germany
*Correspondence e-mail: joachim.wagner@uni-rostock.de

Edited by E. P. Gilbert, Australian Centre for Neutron Scattering, ANSTO, Australia (Received 14 February 2024; accepted 12 August 2024; online 25 September 2024)

Here, it is investigated how optical properties of single scatterers in interacting multi-particle systems influence measurable structure factors. Both particles with linear gradients of their scattering length density and core–shell structures evoke characteristic deviations between the weighted sum 〈S(Q)〉 of partial structure factors in a multi-component system and experimentally accessible measurable structure factors SM(Q). While 〈S(Q)〉 contains only the structural information of self-organizing systems, SM(Q) is additionally influenced by the optical properties of their constituents, resulting in features such as changing amplitudes, additional peaks in the low-wavevector region or splitting of higher-order maxima, which are not related to structural reasons. It is shown that these effects can be systematically categorized according to the qualitative behaviour of the form factor in the Guinier region, which enables assessing the suitability of experimentally obtained structure factors to genuinely represent the microstructure of complex systems free from any particular model assumption. Hence, a careful data analysis regarding size distribution and optical properties of single scatterers is mandatory to avoid a misinterpretation of measurable structure factors.

1. Introduction

Colloidal dispersions attract wide interest in condensed matter physics as highly tunable model systems, mimicking atoms and molecules on the much larger mesoscopic scale with typical length scales between 10 and 1000 nm. Studying these systems enabled major advances in the comprehension of the characteristics of simple fluids and solids, which stimulated the progress of significant theoretical developments towards the understanding of complex systems and materials (Lu & Weitz, 2013[Lu, P. J. & Weitz, D. A. (2013). Annu. Rev. Condens. Matter Phys. 4, 217-233.]).

Scattering experiments serve as essential methods for structural and dynamical investigations in colloidal many-particle systems (Li et al., 2016[Li, T., Senesi, A. J. & Lee, B. (2016). Chem. Rev. 116, 11128-11180.]). Small-angle scattering (SANS with neutrons or SAXS with X-rays as a probe) enables the characterization of colloidal suspensions across the entire range of relevant scattering vector magnitudes Q (Glatter, 2018[Glatter, O. (2018). Scattering Methods and their Application in Colloid and Interface Science. Amsterdam: Elsevier.]). By employing visible light, which is also a natural choice since its wavelength is of the same order of magnitude as the typical size of a colloidal particle, the same type of analysis is in principle also possible in a simpler laboratory setup. This is however connected with the cost of a limited resolution and, as a consequence thereof, the restriction to comparatively large structures (Bohren & Huffmann, 2008[Bohren, C. F. & Huffman, D. R. (2008). Absorption and Scattering of Light by Small Particles. New York: John Wiley.]).

In non-interacting systems, the positions and orientations of the colloidal particles are completely uncorrelated. Thus, the scattered intensity results solely from the superposition of the scattering functions of the single constituents. However, when the particles do interact, higher-level structures emerge from inherent self-organization due to interparticle forces, such as electrostatic and steric interactions or van der Waals attractions. The intensity is then influenced both by the optical properties of the scatterers themselves and by the spatial correlations between them. For idealized radially symmetric and monodisperse systems, where all particles are assumed to be identical, the two contributions can be rigorously separated into the form factor P(Q), containing the single-particle properties, and the structure factor S(Q), which encodes the structural correlations, employing the well known factorization I(Q) ∝ P(Q) S(Q) (Hansen et al., 1991[Hansen, J.-P., Levesque, D. & Zinn-Justin, J. (1991). Liquids, Freezing and Glass Transition. Amsterdam: North Holland.]).

Realistic dispersions typically exhibit a distribution of characteristics, prominently through particle size. In polydisperse interacting systems, the characterization via scattering experiments is in general significantly more complicated, as the factorization of the intensity into form factor and structure factor can no longer be employed in a straightforward way (Salgi & Rajagopalan, 1993[Salgi, P. & Rajagopalan, R. (1993). Adv. Colloid Interface Sci. 43, 169-288.]). Additionally, the observed diffraction patterns become increasingly featureless for broader size distributions, further obstructing the interpretation of experimental intensities. The analysis of multi-component systems thus requires a thorough understanding of the underlying distributions of scattering properties and particle interactions. Insights can be gained through contrast-variation techniques (Ballauff, 2001[Ballauff, M. (2001). Curr. Opin. Colloid Interface Sci. 6, 132-139.]); selectively altering the contrast between specific particle types or between particles and the surrounding medium allows for the isolation and probing of distinct species, aiding the validation of theoretical models.

From an experimental standpoint, it is useful to analyse the measurable structure factor SM(Q), which is defined in such a way that the factorization property I(Q) ∝ P(Q) SM(Q) is also recovered in the polydisperse case (Hansen et al., 1991[Hansen, J.-P., Levesque, D. & Zinn-Justin, J. (1991). Liquids, Freezing and Glass Transition. Amsterdam: North Holland.]). SM(Q) is comparatively easy to access experimentally from the ratio between the intensities of an interacting suspension and a highly diluted non-interacting one. It is as such also widely used as a measure for structural correlations in polydisperse systems, where the height of the principal peak is especially well established as an order parameter (Banchio et al., 1998[Banchio, A. J., Nägele, G. & Ferrante, A. (1998). J. Colloid Interface Sci. 208, 487-499.]). Under specific circumstances, this type of analysis can however turn into a serious pitfall: SM(Q) is also fundamentally affected by optical properties of the particles and not only by their interactions (Salgi & Rajagopalan, 1993[Salgi, P. & Rajagopalan, R. (1993). Adv. Colloid Interface Sci. 43, 169-288.]).

For certain types of dispersions, some simplifying assumptions can be employed. In dilute suspensions of strongly interacting charged particles, for example, the interparticle distances typically are about an order of magnitude larger than the particle sizes because of the large electrostatic repulsion (Hayter & Penfold, 1981[Hayter, J. B. & Penfold, J. (1981). Mol. Phys. 42, 109-118.]). In such a case, the correlation between the particle positions and the scattering amplitudes can be neglected. This neglect of correlations leads to the `decoupling approximation', under which SM(Q) can be decomposed into a structure factor that genuinely represents the averaged structural correlations and weighting factors solely dependent on the scattering amplitudes (Pusey et al., 1982[Pusey, P. N., Fijnaut, H. M. & Vrij, A. (1982). J. Chem. Phys. 77, 4270-4281.]; Kotlarchyk & Chen, 1983[Kotlarchyk, M. & Chen, S.-H. (1983). J. Chem. Phys. 79, 2461-2469.]). This type of analysis was also recently used in a SANS study of moderately concentrated poly(N-isopropylacrylamide) microgels (Zhou et al., 2023[Zhou, B., Gasser, U. & Fernandez-Nieves, A. (2023). Phys. Rev. E, 108, 054604.]), where again the importance of an accurate treatment of polydispersity was stressed. Non-spherical particles cause effects in scattering patterns that appear quite similar to those introduced by size dispersity in systems of spherical particles (Pusey et al., 1982[Pusey, P. N., Fijnaut, H. M. & Vrij, A. (1982). J. Chem. Phys. 77, 4270-4281.]). These effects can to a certain degree also be treated within the decoupling approximation (Kotlarchyk & Chen, 1983[Kotlarchyk, M. & Chen, S.-H. (1983). J. Chem. Phys. 79, 2461-2469.]), whose range of validity has been extensively characterized (Greene et al., 2016[Greene, D. G., Ferraro, D. V., Lenhoff, A. M. & Wagner, N. J. (2016). J. Appl. Cryst. 49, 1734-1739.]). In highly concentrated suspensions, where particles are in close contact, the accuracy of the decoupling approximation is strongly diminished, as in these systems the correlation lengths of the particles' centres of mass are comparable to the correlation lengths inside the particles themselves (Pedersen, 1997[Pedersen, J. S. (1997). Adv. Colloid Interface Sci. 70, 171-210.]). At such high particle volume fractions, excluded volume effects are the predominant contribution to the total interaction potential.

The fundamental interactions in dense colloidal dispersions consisting of spherical particles can to a good approximation be theoretically described with the hard-sphere model (Kirkwood & Boggs, 1942[Kirkwood, J. G. & Boggs, E. M. (1942). J. Chem. Phys. 10, 394-402.]; Widom, 1967[Widom, B. (1967). Science, 157, 375-382.]). The radial distribution functions of an n-component mixture of hard spheres can be calculated within the Percus–Yevick closure of the Ornstein–Zernike equation (Percus & Yevick, 1958[Percus, J. K. & Yevick, G. J. (1958). Phys. Rev. 110, 1-13.]) using Baxter's technique (Baxter, 1970[Baxter, R. J. (1970). J. Chem. Phys. 52, 4559-4562.]), giving access to the corresponding partial structure factors (Vrij, 1978[Vrij, A. (1978). J. Chem. Phys. 69, 1742-1747.], 1979[Vrij, A. (1979). J. Chem. Phys. 71, 3267-3270.]; Blum & Stell, 1979[Blum, L. & Stell, G. (1979). J. Chem. Phys. 71, 42-46.], 1980[Blum, L. & Stell, G. (1980). J. Chem. Phys. 72, 2212.]). Building on Vrij's work (Vrij, 1979[Vrij, A. (1979). J. Chem. Phys. 71, 3267-3270.]; van Beurten & Vrij, 1981[Beurten, P. van & Vrij, A. (1981). J. Chem. Phys. 74, 2744-2748.]), Griffith et al. (1987[Griffith, W. L., Triolo, R. & Compere, A. L. (1987). Phys. Rev. A, 35, 2200-2206.]) presented an analytical scattering function of a polydisperse hard-sphere fluid with a Schulz–Flory distribution (Flory, 1936[Flory, P. J. (1936). J. Am. Chem. Soc. 58, 1877-1885.]; Schulz, 1939[Schulz, G. V. (1939). Z. Phys. Chem. 43B, 25-46.]) of particle diameters. Despite their helpfulness, these expressions are not widely used because of their perceived complexity. Nayeri et al. (2009[Nayeri, M., Zackrisson, M. & Bergenholtz, J. (2009). J. Phys. Chem. B, 113, 8296-8302.]) later extended this approach to core–shell structured hard spheres and used their expressions to describe experimental intensities of a hard-sphere-like microemulsion system. Only recently, Botet et al. (2020[Botet, R., Kwok, S. & Cabane, B. (2020). J. Appl. Cryst. 53, 1570-1582.]) provided expressions for SM(Q) in a simple accessible form and for a number of commonly encountered size distributions. Their analytical expressions are valid for hard optically homogeneous spheres.

This resurgence of interest is an incentive to systematically examine how different form-factor models affect the characteristics of measurable structure factors. It is a well known fact that, especially in dense dispersions, SM(Q) is generally not equal to the structure factorS(Q)〉 representing the averaged spatial correlations of the entire system (Salgi & Rajagopalan, 1993[Salgi, P. & Rajagopalan, R. (1993). Adv. Colloid Interface Sci. 43, 169-288.]; Frenkel et al., 1986[Frenkel, D., Vos, R. J., de Kruif, C. G. & Vrij, A. (1986). J. Chem. Phys. 84, 4625-4630.]). The mismatch between SM(Q) and 〈S(Q)〉 is precisely the reason why in many past studies an extraction of SM(Q) is deliberately not attempted. By describing the scattered intensity of concentrated suspensions with sophisticated models, a thorough characterization of particle properties is possible without calculating SM(Q), as performed, for example, by Stieger, Pedersen et al. (2004[Stieger, M., Pedersen, J. S., Lindner, P. & Richtering, W. (2004). Langmuir, 20, 7283-7292.]); Stieger, Richtering et al. (2004[Stieger, M., Richtering, W., Pedersen, J. S. & Lindner, P. (2004). J. Chem. Phys. 120, 6197-6206.]); Zackrisson et al. (2005[Zackrisson, M., Stradner, A., Schurtenberger, P. & Bergenholtz, J. (2005). Langmuir, 21, 10835-10845.]); Balogh et al. (2007[Balogh, J., Olsson, U. & Pedersen, J. S. (2007). J. Phys. Chem. B, 111, 682-689.]); or Scotti (2021[Scotti, A. (2021). Soft Matter, 17, 5548-5559.]).

Certainly, however, for many applications in condensed matter science, gaining an accurate approximation of the structure factor is still highly desired. This is especially the case when employing computer simulations or many-body theory to model structural phenomena and wanting to compare detailed facets of particle self-organization directly with experimental outcomes (Dekker et al., 2020[Dekker, F., Kuipers, B. W. M., González García, Á., Tuinier, R. & Philipse, A. P. (2020). J. Colloid Interface Sci. 571, 267-274.]; Peláez-Fernández et al., 2011[Peláez-Fernández, M., Moncho-Jordá, A. & Callejas-Fernández, J. (2011). J. Chem. Phys. 134, 054905.]; Anta & Madden, 1999[Anta, J. A. & Madden, P. A. (1999). J. Phys. Condens. Matter, 11, 6099-6111.]; Krause et al., 1991[Krause, R., D'Aguanno, B., Mendez-Alcaraz, J. M., Nägele, G., Klein, R. & Weber, R. (1991). J. Phys. Condens. Matter, 3, 4459-4475.]; Stellbrink et al., 2002[Stellbrink, J., Allgaier, J., Richter, D., Moussaid, A., Schofield, A. B., Poon, W. C. K., Pusey, P. N., Lindner, P., Dzubiella, J., Likos, C. N. & Löwen, H. (2002). Appl. Phys. Mater. Sci. Process. 74, s355-s357.]). Beyond using structure factors to deduct structural patterns from experimental scattering data (Mohanty et al., 2017[Mohanty, P. S., Nöjd, S., van Gruijthuijsen, K., Crassous, J. J., Obiols-Rabasa, M., Schweins, R., Stradner, A. & Schurtenberger, P. (2017). Sci. Rep. 7, 1487.]; Phalakornkul et al., 1996[Phalakornkul, J. K., Gast, A. P., Pecora, R., Nägele, G., Ferrante, A., Mandl-Steininger, B. & Klein, R. (1996). Phys. Rev. E, 54, 661-675.]; Scheffold & Mason, 2009[Scheffold, F. & Mason, T. G. (2009). J. Phys. Condens. Matter, 21, 332102.]; Mason et al., 2006[Mason, T. G., Graves, S. M., Wilking, J. N. & Lin, M. Y. (2006). J. Phys. Chem. B, 110, 22097-22102.]), these quantities are used in theoretical approaches to calculate short- and long-time dynamics in many-particle systems. Two notable examples in this context are the δγ-expansion by Beenakker & Mazur (1983[Beenakker, C. W. J. & Mazur, P. (1983). Physica A, 120, 388-410.], 1984[Beenakker, C. W. J. & Mazur, P. (1984). Physica A, 126, 349-370.]) used to model hydrodynamic effects during short-time diffusion (Genz & Klein, 1991[Genz, U. & Klein, R. (1991). Physica A, 171, 26-42.]) and the widely known mode-coupling theory of the glass transition (Janssen, 2018[Janssen, L. M. C. (2018). Front. Phys. 6, 97.]), which both need the static structure factor as an input for calculations. In these theories, a popular method is to circumvent computationally expensive multi-component calculations by considering an effective one-component analysis using the experimentally obtained structure factor SM(Q) directly, assuming this quantity to be an accurate representation of the average structure factorS(Q)〉 (Robert et al., 2008[Robert, A., Wagner, J., Härtl, W., Autenrieth, T. & Grübel, G. (2008). Eur. Phys. J. E, 25, 77-81.]; Di Cola et al., 2009[Di Cola, E., Moussaïd, A., Sztucki, M., Narayanan, T. & Zaccarelli, E. (2009). J. Chem. Phys. 131, 144903.]).

For a number of specialized cases, deviations between 〈S(Q)〉 and SM(Q) and the general influence of varying optical properties on SM(Q) have already been assessed (Banchio et al., 1998[Banchio, A. J., Nägele, G. & Ferrante, A. (1998). J. Colloid Interface Sci. 208, 487-499.]; Pedersen, 2001[Pedersen, J. S. (2001). J. Chem. Phys. 114, 2839-2846.]). The purpose of this contribution is to raise further awareness on how particles' optical properties influence the shape of SM(Q) while the underlying interactions remain unchanged and to show that these observations can be model-independently systematized on the basis of quite universal principles. This enables practitioners to make informed judgements under which circumstances such an experimentally obtained structure factor can still serve as a valid order parameter. We show typical examples of shapes that can be realistically encountered during contrast-variation experiments, so even without explicitly employing established theoretical models [see e.g. Pedersen (1997[Pedersen, J. S. (1997). Adv. Colloid Interface Sci. 70, 171-210.]) for a large collection of scattering functions], a qualitative assessment of experimental findings is possible. We also analyse two simplified models for optically inhomogeneous particles: those with a linear gradient of the scattering contrast and spheres with a core–shell structure. Nevertheless, the approach is readily adaptable to any model and provides a toolbox for the modelling of measurable structure factors for hard-sphere suspensions with arbitrary form factors, as demonstrated in several past studies (Vrij, 1979[Vrij, A. (1979). J. Chem. Phys. 71, 3267-3270.]; Frenkel et al., 1986[Frenkel, D., Vos, R. J., de Kruif, C. G. & Vrij, A. (1986). J. Chem. Phys. 84, 4625-4630.]; Pedersen, 2001[Pedersen, J. S. (2001). J. Chem. Phys. 114, 2839-2846.]; Nayeri et al., 2009[Nayeri, M., Zackrisson, M. & Bergenholtz, J. (2009). J. Phys. Chem. B, 113, 8296-8302.]; Botet et al., 2020[Botet, R., Kwok, S. & Cabane, B. (2020). J. Appl. Cryst. 53, 1570-1582.]; Diaz Maier & Wagner, 2024[Diaz Maier, J. & Wagner, J. (2024). Soft Matter, 20, 1309-1319.]).

2. Scattering of hard-sphere mixtures

We consider a mixture of spherical particles, where each particle can be categorized into one of n species. The composition of the mixture is specified by the number fractions xα = Nα/N, where N is the total number of particles and Nα is the number of particles belonging to species α. We further restrict ourselves to elastic single scattering events where the Born approximation is applicable. In such a case, the mean intensity

[\displaystyle I(Q)\propto{\textstyle\sum\limits_{\alpha,\beta}}(x_{\alpha}x_{\beta})^{1/2} f_{\alpha}(Q)f_{\beta}(Q)S_{\alpha\beta}(Q) \eqno (1)]

is proportional to the weighted sum of the single-particle scattering amplitudes fα(Q) and the partial structure factors Sαβ(Q) (Salgi & Rajagopalan, 1993[Salgi, P. & Rajagopalan, R. (1993). Adv. Colloid Interface Sci. 43, 169-288.]). Herein, the scattering amplitude

[\displaystyle f_{\alpha}(Q) = 4\pi\int\limits_{0}^{\infty}\rho_{\alpha}(r)r^{2} {{\sin(Qr)} \over {Qr}}\ {\rm d}r \eqno (2)]

is the Fourier–Bessel transform of the scattering contrast ρα(r), whereas the partial structure factors Sαβ(Q) are obtained from the solution of the multi-component Ornstein–Zernike equation. Expressions for Sαβ(Q) of the hard-sphere fluid within the Percus–Yevick closure are given by Vrij (1979[Vrij, A. (1979). J. Chem. Phys. 71, 3267-3270.]) but, for the convenience of the reader, the solution is re-articulated in Appendix A[link], and presented in a manner that is accessible and easily applicable. This genuine multi-component approach involving partial structure factors is a general and versatile formalism, from which both the decoupling approximation (Kotlarchyk & Chen, 1983[Kotlarchyk, M. & Chen, S.-H. (1983). J. Chem. Phys. 79, 2461-2469.]) and the local monodisperse approximation (Pedersen, 1994[Pedersen, J. S. (1994). J. Appl. Cryst. 27, 595-608.]), serving as a localized effective one-component approach suited for very polydisperse systems, can be derived.

For non-interacting particles, the partial structure factors are simply Sαβ(Q) = δαβ, where δαβ denotes the Kronecker symbol. Equation (1[link]) then reduces to the size average of the squared scattering amplitudes,

[\displaystyle I(Q)\propto\langle f^{2}(Q)\rangle = {\textstyle\sum\limits_{\alpha}}x_{\alpha }\,f^{2}_{\alpha}(Q). \eqno (3)]

The average form factor

[\displaystyle P(Q) = {{\langle f^{2}(Q)\rangle} \over {\langle f^{2}(0)\rangle}}\eqno (4)]

is familiarly obtained from the normalization to forward scattering. As the measurable structure factor should satisfy the relation I(Q) ∝ P(Q) SM(Q), the expression

[\displaystyle S_{{\rm M}}(Q) = \left[\langle f^{2}(Q)\rangle\right]^{-1}{\textstyle\sum \limits_{\alpha,\beta}}(x_{\alpha}x_{\beta})^{1/2}f_{\alpha}(Q)f_{\beta}(Q)S_{ \alpha\beta}(Q) \eqno (5)]

results from the combination of equations (1[link]) and (3[link]). The averaged structure factor

[\displaystyle\langle S(Q)\rangle = {\textstyle\sum\limits_{\alpha,\beta}}(x_{\alpha}x_{\beta })^{1/2}S_{\alpha\beta}(Q) \eqno (6)]

provides information about the total spatial correlations between all present particles, regardless of their species labels. It represents a true thermodynamic average independent of any optical properties. Any deviation between SM(Q) and 〈S(Q)〉 is thus a measure for the perturbation of 〈S(Q)〉 caused by the scattering amplitudes.

We now want to explore the influence of the scattering amplitudes on the shape of SM(Q). The aim is to gain a qualitative understanding of generic patterns; so to keep the analysis tractable, only a single representative size distribution is considered. For this purpose, the Schulz–Flory distribution with probability density

[\displaystyle c(R) = {{1} \over {\Gamma(Z+1)}}\left({{Z+1} \over {\langle R\rangle}} \right)^{Z+1}R^{Z}\exp\left(-{{Z+1} \over {\langle R\rangle}}R\right) \eqno (7)]

is chosen. Here, R is the particle radius with mean 〈R〉 and Γ(x) represents the gamma function. The polydispersity p of the system is specified by the shape parameter Z via p2 = (〈R2〉 − 〈R2)/〈R2 = 1/(Z + 1). The idea is now to discretize the distribution to a representative n-component mixture. For the Schulz–Flory distribution, an efficient way to achieve this is by exploiting the generalized Gauss–Laguerre quadrature rule, specifically used to calculate integrals with a weighting function like equation (7[link]) (D'Aguanno & Klein, 1992[D'Aguanno, B. & Klein, R. (1992). Phys. Rev. A, 46, 7652-7656.]; D'Aguanno, 1993[D'Aguanno, B. (1993). Phys. Scr. T49A, 84-88.]; Olver et al., 2010[Olver, F. W. J., Lozier, D. W., Boisvert, R. F. & Clark, C. W. (2010). The NIST Handbook of Mathematical Functions. Cambridge University Press.]). The nodes and weights generated by such a procedure are equivalent to the particle radii and number fractions of a discrete mixture which shares the first 2n − 1 moments 〈Rn〉 with the original continuous distribution. For each calculated scattering function, we carefully checked that the number of nodes necessary for convergence was reached. The numerical scheme was further tested against the analytical SM(Q) for homogeneous spheres provided by Botet et al. (2020[Botet, R., Kwok, S. & Cabane, B. (2020). J. Appl. Cryst. 53, 1570-1582.]), where excellent agreement was found.

3. Measurable structure factors of polydisperse systems

3.1. General remarks

Fig. 1[link] provides a general introductory overview of the influence of polydispersity on P(Q), SM(Q) and 〈S(Q)〉, discussed for a dense suspension of homogeneous spheres, serving as a reiteration of well known phenomenology (Botet et al., 2020[Botet, R., Kwok, S. & Cabane, B. (2020). J. Appl. Cryst. 53, 1570-1582.]). Concerning the form factors, only those corresponding to polydispersities of less than 10% appear structured. Familiarly, the characteristic minima in P(Q) become increasingly smeared out for broader size distributions.

[Figure 1]
Figure 1
Comparative analysis of scattering functions for an ensemble of optically homogeneous hard spheres with varying degrees of polydispersity. (a) Probability density function illustrating the Schulz–Flory distributed radius R. (b) Form factor P(Q). (c) Average structure factorS(Q)〉. (d) Measurable structure factor SM(Q). All evaluated at a total volume fraction of φ = 0.5, spanning polydispersities from 5 to 25%. 〈R〉 denotes the mean radius of the spheres.

Polydispersity also causes a change in the initial slope of P(Q) in the Guinier region. Reflecting the distribution of particle sizes when calculating the Taylor expansion of P(Q), the slope is now given by [-Q^{2}\langle R_{{\rm G}}^{2}\rangle/3], where the familiar radius of gyration RG is substituted by an apparent radius of gyration [\langle R_{{\rm G}}^{2}\rangle^{1/2}] (Glatter, 2018[Glatter, O. (2018). Scattering Methods and their Application in Colloid and Interface Science. Amsterdam: Elsevier.]; Tomchuk et al., 2014[Tomchuk, O. V., Bulavin, L. A., Aksenov, V. L., Garamus, V. M., Ivankov, O. I., Vul', A. Y., Dideikin, A. T. & Avdeev, M. V. (2014). J. Appl. Cryst. 47, 642-653. ]). For homogeneous spheres,

[\displaystyle\langle R_{{\rm G}}^{2}\rangle = {{3}\over{5}}{{\langle R^{8 }\rangle}\over{\langle R^{6}\rangle}}\eqno (8)]

is obtained, which reduces to the well known result of [R_{{\rm G}}^{2} = (3/5)R^{2}] for monodisperse systems.

Similarly to P(Q), both the measurable structure factor SM(Q) and the average structure factorS(Q)〉 become increasingly featureless at high polydispersities, which is distinctively noticeable as the principal peak's amplitude decreases and the secondary oscillations gradually disappear. Shifting the focus to direct comparison between the two structure factors SM(Q) and 〈S(Q)〉, multiple observations are apparent. While the amplitude of the principal peak is similar for both functions, differences appear at larger wavevectors, where secondary peaks in SM(Q) appear at roughly the locations of the form-factor minima, as similarly noticed by Ginoza & Yasutomi (1999[Ginoza, M. & Yasutomi, M. (1999). J. Phys. Soc. Jpn, 68, 2292-2297.]). With increasing polydispersity, these maxima evolve into broad shoulders that get smeared out eventually. As also noted by Ginoza & Yasutomi (1999[Ginoza, M. & Yasutomi, M. (1999). J. Phys. Soc. Jpn, 68, 2292-2297.]), sharp secondary maxima are hard to observe experimentally because a very narrow size distribution in combination with a homogenous distribution of the scattering length density (SLD) inside the particles is required. On the other hand, shoulder-like features in experimentally determined structure factors are well documented [see, as an example, Di Cola et al. (2009[Di Cola, E., Moussaïd, A., Sztucki, M., Narayanan, T. & Zaccarelli, E. (2009). J. Chem. Phys. 131, 144903.])]. In the low-Q region, a striking observation is the significant increase of 〈S(0)〉 at elevated polydispersities in comparison with SM(0). According to the fluctuation-dissipation theorem from statistical mechanics, the isothermal compressibility κT is for monodisperse systems connected to the zero-wavevector limit of S(Q) via S(0) = ρkBTκT, where ρ denotes the number density and kBT is the thermal energy. The extension of this concept to mixtures must however be treated with caution because, for multi-component systems, the connection between structure and thermodynamics is not simply given by the size average 〈S(0)〉. According to the Kirkwood–Buff theory of solutions, it is instead given by the relation [(\rho k_{{\rm B}}T\kappa_{{\rm T}})^{-1} =] [\sum x_{\alpha}x_{\beta}S^{-1}_ {\alpha\beta}(0)], where [S^{-1}_{\alpha\beta}(Q)] is the αβ element of the inverse structure-factor matrix (Hansen & McDonald, 2013[Hansen, J.-P. & McDonald, I. R. (2013). Theory of Simple Liquids with Applications to Soft Matter. Amsterdam: Elsevier.]).

3.2. Linear contrast gradient

As a prototypical example for particles with inhomogeneous scattering strength, particles with a linear gradient of the SLD are investigated. This is particularly relevant for swellable particles into which the suspension medium can diffuse. This can occur with microgel particles (Karg et al., 2019[Karg, M., Pich, A., Hellweg, T., Hoare, T., Lyon, L. A., Crassous, J. J., Suzuki, D., Gumerov, R. A., Schneider, S., Potemkin, I. I. & Richtering, W. (2019). Langmuir, 35, 6231-6255.]), for example. Under certain reaction conditions, an inhomogeneous degree of cross-linking arises, which also leads to inhomogeneous scattering properties. Particles with intrinsic material gradients are also plausible, obtained for example by continuously changing the monomer composition in a feed process during synthesis. Then, in principle, a suspension in which the contrast within a particle changes its sign can also be realized. The form of a linear gradient is assumed for the sake of simplicity in order to investigate the phenomenology of continuous contrasts as an example.

The scattering contrast as a function of the distance r from the centre can for a single particle be parametrized as

[\displaystyle\rho(r) = \cases{\displaystyle\rho_{0}+(\rho_{R}-\rho_{0}){{r}\over{R}} & $ {\rm if } \ 0 \leq r\leq R,$\cr 0 & $ \ {\rm otherwise},$} \eqno (9)]

where R is the particle radius, ρ0 is the contrast in the centre and ρR is the contrast at the interface to the surrounding medium. Accordingly, the resulting single-particle scattering amplitude is given by

[\eqalignno{ f(Q) =\,\, &4\pi\biggl\{\rho_{0}{{\sin(QR)-QR\cos(QR)} \over {Q^{3}}}+{{ \rho_{R}-\rho_{0}} \over {R}}\cr &\times{{2QR\sin(QR)-[(QR)^{2}-2]\cos(QR)-2} \over {Q^{4}}} \biggr\}, & (10)}]

which reduces to

[\displaystyle f(0) = \pi R^{3}\left({{\rho_{0}} \over {3}}+\rho_{R}\right) \eqno (11)]

in the forward-scattering limit. A closer look at equation (11[link]) reveals that the forward-scattering contribution disappears if the condition ρR/ρ0 = −1/3 is fulfilled. In particular, when the maximum accessible scattering vector is limited, as in the case of light scattering, forward scattering contributes significantly to the total scattering cross section. If the forward scattering is zero, the sample appears almost optically transparent. Refractive-index matching can be achieved for particles with a homogeneous scattering capacity if the SLD of the suspension medium is adapted to that of the particles. If the scattering capacity is inhomogeneous, index matching can only minimize the total scattering cross section, which is often achieved by making the forward scattering almost zero. In the following, the condition when the forward scattering power is minimal is referred to as the index match point.

To gain a systematic understanding of the behaviour of the measurable structure factor SM(Q) as a function of the contrast ratio ρR/ρ0, it will prove advantageous to investigate the Guinier region of the form factor. Using the contrast profile from equation (9[link]), for a single particle with radius R,

[\displaystyle R_{{\rm G}}^{2} = {{2}\over{5}}{{\rho_{0}+5\rho_{R}}\over{\rho _{0}+3\rho_{R}}} R^{2} \eqno (12)]

is obtained for the effective squared radius of gyration, which depends not only on the particle's radius but also on the two contrast parameters ρ0 and ρR. For polydisperse suspensions, a similar expression emerges:

[\displaystyle\langle R_{{\rm G}}^{2}\rangle = {{2}\over{5}}\,{{\rho_{0}+5 \rho_{R}}\over{\rho_{0}+3\rho_{R}}}\,{{\langle R^{8}\rangle}\over{\langle R^{6} \rangle}}. \eqno (13)]

As such, the contrast dependence of the prefactor is not altered by polydispersity and the qualitative discussion can instead be based on monodisperse suspensions. We will thus refer to the prefactor simply as [R_{{\rm G}}^{2}/R^{2}], even in the polydisperse case.

Inspecting equation (12[link]), several characteristic ratios ρR/ρ0 are apparent: [R_{{\rm G}}^{2}] becomes zero for ρR/ρ0 = −1/5; exhibits a pole at ρR/ρ0 = −1/3, incident with the index match point; and has an asymptotic limit of [R_{{\rm G}}^{2}/R^{2} = 2/3] for ρR/ρ0 → ±∞. Here, it will be shown that the behaviour of the scattering functions can be divided into three qualitatively distinct classes, and that form factors and measurable structure factors within each domain share unique features. The classification based on the behaviour of [R_{{\rm G}}^{2}], together with form factors P(Q) and measurable structure factors SM(Q) representative of each region, is visualized in Fig. 2[link]. The regions are characterized as follows:

[Figure 2]
Figure 2
An illustrative breakdown of the classification of the scattering functions of spheres with a linear contrast gradient into the three regimes discussed in the main text, where each column corresponds to a unique region. In the top row, the reduced squared radius of gyration [R_{{\rm G}}^{2}/R^{2}] as a function of the contrast ratio ρR/ρ0 is depicted. The location of the respective regions labelled (I), (II) or (III) is indicated by the darker shaded area. The middle and bottom rows display selected form factors P(Q) and measurable structure factors SM(Q) that exemplify each region's variability in the shape observed during contrast variation. 〈R〉 indicates the mean radius of the particles. Note the shared axes of P(Q) and SM(Q) between rows and columns.

(I) For ρR/ρ0 > −1/5, [R_{{\rm G}}^{2}] is positive and the form factors have the familiar decaying shape known from homogeneous spheres. With decreasing contrast ratio, the decay becomes increasingly gradual until [R_{{\rm G}}^{2} = 0] is reached for ρR/ρ0 = −1/5. Around the principal peak of SM(Q) and for lower wavevectors, changes in the contrast have a negligible influence on the measurable structure factors. However, at wavevectors beyond the principal peak's location, SM(Q) is greatly affected by contrast variation. Depending on the specific location of the first form-factor minimum, which shifts to larger wavevectors with lower contrast ratios, the shoulder-like artefact also visible in Fig. 1[link] moves through SM(Q) towards larger wavevectors and therein most prominently affects the shape of the first local minimum and the following secondary maximum.

(II) For the contrast ratios −1/3 < ρR/ρ0 < −1/5, [R_{{\rm G}}^{2}] becomes negative, which implies an imaginary radius of gyration RG leading to a positive initial slope of P(Q). Form factors in this region therefore initially increase from P(0) = 1 until a global maximum is reached at QR ≃ 4, after which they decay. The height of the maximum increases as the contrast ratio moves towards the index match point at ρR/ρ0 = −1/3. Curiously, the measurable structure factors in this domain are almost indistinguishable, even though the variation of [R_{{\rm G}}^{2}] is much more pronounced in comparison with region (I), where the span of [R_{{\rm G}}^{2}] is small but SM(Q) shows a much more diverse behaviour. Also, the distorting artefacts from region (I) disappear almost completely.

(III) Contrast ratios of ρR/ρ0 < −1/3 again result in positive [R_{{\rm G}}^{2}] and negative initial slopes. Close to the index match point, where [R_{{\rm G}}^{2}] is comparatively large, P(Q) exhibits an intriguing shape. At small wavevectors, a pronounced minimum occurs even in very polydisperse suspensions. Beyond the minimum, P(Q) rises to a global maximum reminiscent of region (II). For contrasts in this range, an additional local maximum in SM(Q) appears at low wavevectors, caused by the presence of the first form-factor minimum. Such secondary maxima are often discussed in the literature as an indication of self-organization on length scales beyond the distance of nearest neighbours, i.e. the formation of correlated clusters (Sciortino et al., 2004[Sciortino, F., Mossa, S., Zaccarelli, E. & Tartaglia, P. (2004). Phys. Rev. Lett. 93, 055701.]; Liu et al., 2005[Liu, Y., Chen, W.-R. & Chen, S.-H. (2005). J. Chem. Phys. 122, 044507.]). The secondary maxima occurring here are exclusively caused by the scattering amplitudes and cannot be attributed to structural properties of the sample. This constitutes a valuable example of a situation where a careless inspection of experimentally determined SM(Q) can in the worst case lead to unjustified assumptions about the structure of a system. Moving further away from the index match point, the first form-factor minimum moves towards larger wavevectors and gets shallower. At the same time, the following maximum declines and, as such, the shape of P(Q) morphs back into the familiar decaying shape from region (I). Simultaneously, the location of the secondary maximum in SM(Q) drifts towards higher wavevectors. Fig. 2[link] also displays a situation where the form-factor minimum exactly coincides with the location where the principal peak of SM(Q) would normally occur. In this case, the main peak is drastically diminished, which is again not an indicator for a less pronounced short-range order in this particular instance, but can certainly be mistaken as such.

The principal-peak height of a structure factor is an often employed structural order parameter. Scheffold & Mason (2009[Scheffold, F. & Mason, T. G. (2009). J. Phys. Condens. Matter, 21, 332102.]) noticed in their investigation of highly concentrated nanoemulsions that the peak amplitude in SM(Q) is deeply affected by polydispersity. As such, the evolution of this height during contrast variation is also of special interest. Fig. 3[link] compares the peak height of the average structure factorS(Qmax)〉 with the value of SM(Qmax) at the same wavevector as a function of the contrast ratio ρR/ρ0 and for different degrees of polydispersity. Overall, it is clearly shown that SM(Qmax) is deeply affected by changes in the contrast. There exist two contrast ratios where SM(Qmax) and 〈S(Qmax)〉 coincide. One of them is to a good approximation given by ρR/ρ0 ≃ −1/5, the location where the apparent radius of gyration disappears and P(Q) decays very slowly. The other location is at a positive contrast ratio and drifts towards higher ρR/ρ0 with increasing polydispersity. Bounded by those two ratios is a regime where SM(Qmax) exceeds 〈S(Qmax)〉, while for all other contrast ratios, the peak height from SM(Qmax) underestimates the actual height. For comparatively small polydispersities around 5%, the deviation from 〈S(Qmax)〉 is small and only amounts to a few per cent, as long as the contrast ratio is larger than ρR/ρ0 ≃ −1/5. For lower ratios, SM(Qmax) is strongly diminished, most pronouncedly at contrast ratios of ρR/ρ0 ≃ −1. For higher polydispersities, the deviations become even more severe, as best visualized in Fig. 3[link](b), where the relative deviation between 〈S(Qmax)〉 and SM(Qmax) is depicted. Even in the immediate vicinity of 〈S(Qmax)〉 = SM(Qmax), already deviations of the order of 5–10% appear for the highest shown polydispersities. This demonstrates that, no matter what the actual degree of polydispersity, SM(Qmax) can only serve as a reliable order parameter for very specific contrast ratios.

[Figure 3]
Figure 3
(a) The influence of the contrast ratio ρR/ρ0 on the principal-peak value SM(Qmax) for spheres with a linear contrast gradient, for polydispersities in a range between 5 and 25% at a total volume fraction of φ = 0.5. The horizontal dashed lines mark, for comparison, the height of the principal peak of the average structure factorS(Qmax)〉. The distinction between the different introduced contrast regimes from Fig. 2[link] is indicated by the vertical dashed lines. (b) Relative deviation between SM(Qmax) and 〈S(Qmax)〉 for an enlarged region.

3.3. Core–shell particles

Core–shell models are commonly employed to describe particles consisting of different layers of material, e.g. nanoparticles with grafted stabilizer shells (Hallett et al., 2020[Hallett, J. E., Grillo, I. & Smith, G. N. (2020). Langmuir, 36, 2071-2081.]; Diaz Maier & Wagner, 2024[Diaz Maier, J. & Wagner, J. (2024). Soft Matter, 20, 1309-1319.]) or micellar structures (Szymusiak et al., 2017[Szymusiak, M., Kalkowski, J., Luo, H., Donovan, A. J., Zhang, P., Liu, C., Shang, W., Irving, T., Herrera-Alonso, M. & Liu, Y. (2017). ACS Macro Lett., 6, 1005-1012.]). As core and shell naturally differ in their material properties, in principle both positive and negative contrast differences with respect to the surrounding medium can occur, similarly to particles with continuous material gradients. For Schulz–Flory-distributed core–shell particles, analytical expressions for the form factor P(Q) exist in the case of a polydisperse core and a shell of constant thickness (Bartlett & Ottewill, 1992[Bartlett, P. & Ottewill, R. H. (1992). J. Chem. Phys. 96, 3306-3318.]), for a polydisperse total diameter and a constant core-to-shell ratio (Wagner, 2004[Wagner, J. (2004). J. Appl. Cryst. 37, 750-756.]), and for both core radius and shell thickness independently distributed (Wagner, 2012[Wagner, J. (2012). J. Appl. Cryst. 45, 513-516.]). Moreover, an analytical solution for the problem of correlated hard-sphere core–shell systems was provided by Nayeri et al. (2009[Nayeri, M., Zackrisson, M. & Bergenholtz, J. (2009). J. Phys. Chem. B, 113, 8296-8302.]).

The scattering amplitude of a single core–shell particle,

[\eqalignno{ f(Q) =\,\, &4\pi\biggl[(\rho_{{\rm c}}-\rho_{{\rm s}}){{\sin(QR_{{\rm c}})-QR_{{\rm c}}\cos(QR_{{\rm c}})}\over{Q^{3}}} \cr &+ \rho_{{\rm s}}{{\sin(QR)-QR\cos(QR)}\over{Q^{3}}}\biggr] ,&(14)}]

is the sum of the amplitudes of a sphere and a spherical shell, weighted by their respective contrasts, ρc and ρs. Rc and R are the core radius and the total radius of the particle, respectively, and we specifically consider the case where the core radius and the total radius are connected by a constant species-independent size ratio δ = Rc/R.

Similar to the gradient model, the forward-scattering contribution

[\displaystyle f(0) = {4\over 3}\pi R^{3}\left[\delta^{3}(\rho_{{\rm c}}- \rho_{{\rm s}})+\rho_{{\rm s}}\right] \eqno (15)]

disappears for specific contrast combinations of the ratio of contrasts ρs/ρc = δ3/(δ3 − 1), which now additionally depends on the size ratio δ. For the effective radius of gyration of a polydisperse system, an expression with similar structure to equation (13[link]) emerges:

[\displaystyle\langle R_{{\rm G}}^{2}\rangle = {{3}\over{5}}\,{{\delta^{5} \rho_{{\rm c}}+(1-\delta^{5})\rho_{{\rm s}}}\over{\delta^{3}\rho_{{\rm c}} +(1-\delta^{3})\rho_{{\rm s}}}}\,{{\langle R^{8}\rangle}\over{\langle R^{6}\rangle}}. \eqno (16)]

That, again, a prefactor containing the contrasts can be decoupled from the size average is a peculiarity of this model with constant size ratio and a key reason why this assumption was made for this investigation.

In Fig. 4[link], the contrast dependence of [R_{{\rm G}}^{2}/R^{2}] is visualized for different size ratios δ. As in the case of spheres with a linear gradient of the SLD, this results in hyperbola-like curves, where the location of the pole is now influenced by δ; an increasing ratio of core diameter to total diameter shifts the location of the pole to more negative contrast ratios (ρs/ρc). The contrast ratio where [R_{{\rm G}}^{2} = 0] is, in comparison, only slightly altered by δ. This leads to a larger range of contrast ratios with negative [R_{{\rm G}}^{2}] as the shell thickness decreases.

[Figure 4]
Figure 4
The influence of the contrast ratio ρs/ρc on the reduced squared radius of gyration [R_{{\rm G}}^{2}/R^{2}] for core–shell particles with different ratios δ between core radius and total radius, along with three representative sets of form factors, each sharing the same radius of gyration for different size ratios.

This shows that core–shell particles exhibit qualitatively comparable optical characteristics to particles with a linear density gradient. As such, the form factors P(Q) of core–shell systems can likewise be categorized into three classes based on their behaviour at low wavevectors. Example form factors for each class are also visualized in Fig. 4[link].

Because of these similarities, we focus the remainder of the discussion on aspects that are unique to core–shell particles, i.e. how measurable structure factors are influenced by different core-to-shell ratios. For this purpose, structure factors corresponding to two important edge cases, particles with a small core and particles with a thin shell, are compared in Fig. 5[link] for different degrees of polydispersity and chosen contrast ratios ρs/ρc. Core–shell models with thin shells are often encountered when characterizing particles stabilized by a grafted polymer layer, which are prototypical colloidal model particles displaying hard-sphere behaviour (Royall et al., 2013[Royall, C. P., Poon, W. C. K. & Weeks, E. R. (2013). Soft Matter, 9, 17-27.]). The case of hard spheres with a strongly scattering small core and a weakly scattering comparatively large shell is equally of interest. Under these conditions, essentially, the behaviour of highly charged strongly repelling particles whose interparticle distance is several times larger than their diameter is artificially mimicked. For these systems, the measurable structure factor SM(Q) should in theory to a good approximation coincide with the average structure factorS(Q)〉. To reasonably compare models with different size ratios (δ), two specific contrast ratios (ρs/ρc) are depicted: the ratio ρs/ρc = δ3/(δ3 − 1) at the index matching point, where for­ward scattering is minimized; and the ratio ρs/ρc = δ5/(δ5 − 1), where [\langle R_{{\rm G}}^{2}\rangle = 0] and P(Q) shows the weakest decay. In the case of [\delta\ll 1], both conditions basically lead to the same result: the shell is virtually hidden with ρs ≃ 0.

[Figure 5]
Figure 5
Comparison between measurable structure factor SM(Q) and size-averaged structure factorS(Q)〉 of core–shell particles for different core-to-total ratios (δ), contrast ratios (ρs/ρc) and polydispersities as indicated in the figure. The total volume fraction for all shown structure factors is φ = 0.5.

As can be observed in Fig. 5[link], for moderate polydispersities of 5–10%, the small core-to-total ratio δ = 0.1 indeed yields measurable structure factors SM(Q) that are indistinguishable from 〈S(Q)〉 for both depicted contrast ratios. For particles with thin shells (δ = 0.9), SM(Q) and 〈S(Q)〉 also agree well in the vicinity of the principal peak. However, differences arise around the secondary maxima, where the peak amplitudes in SM(Q) are diminished because of the interference of the scattering amplitudes. With increasing polydispersity, this deviation becomes more pronounced. Still, even for particles that are seemingly quite close to homogeneous spheres, artefacts in SM(Q) can be significantly reduced by careful contrast variation.

Looking at highly polydisperse systems, it is evident that, even for rather small cores with δ = 0.1, 〈S(Q)〉 cannot be accurately represented by any SM(Q). Only the height of the principle peak is correctly estimated. This stresses again the importance of an accurate treatment of very broad size distributions, where any kind of approximation must be carefully checked for validity.

4. Conclusions

Colloidal dispersions generally exhibit a particle size distribution, which needs to be taken into account when interpreting results from scattering experiments. The measurable structure factor SM(Q) is an experimental, comparatively easily accessible measure for the interparticle structure in interacting systems. However, in polydisperse systems, SM(Q) is, beyond the structural correlations, also decisively affected by the optical properties of the individual particles. To this end, we systematically investigated the influence of different form-factor models on the shape of SM(Q) of dense dispersions with hard-sphere interactions. The characterization of measurable structure factors was extended to two classes of spherical particles with inhomogeneous scattering capacity: first, spheres with a linear SLD profile as a general model for particles with continuous contrast gradients and, second, a core–shell system as a prototype for particles with layered structures.

For both models, we find that the structure factors can be categorized into three distinctive classes of shared qualitative features, based on the behaviour of the form factor P(Q) in the Guinier region. SM(Q) can, for these optically inhomogeneous model particles, be significantly influenced by the variation of the scattering contrasts relative to the surrounding medium. Depending on the specific contrast combination, shoulder-like features emerge, maxima are diminished or split, and even secondary maxima in the low-wavevector region, reminiscent of cluster peaks, can be observed. These effects are solely due to the optical properties of the particles and are not caused by structural changes in the sample. We further showed that the height of the principal peak of SM(Q) can only be regarded as a representative order parameter in a very restricted range of contrasts, especially for broad size distributions.

These observations emphasize the need to properly address the distribution of particle size (and possibly also other characteristics) in the interpretation of static scattering experiments. Actually, for many applications, deliberately broad size distributions are a desired feature; an academically relevant example is studies of deeply supercooled glass-forming systems (Ninarello et al., 2017[Ninarello, A., Berthier, L. & Coslovich, D. (2017). Phys. Rev. X, 7, 021039.]), where crystallization needs to be suppressed and where polydispersity effects in any form certainly cannot be neglected (Zaccarelli et al., 2015[Zaccarelli, E., Liddle, S. M. & Poon, W. C. K. (2015). Soft Matter, 11, 324-330.]; Pihlajamaa et al., 2023[Pihlajamaa, I., Laudicina, C. C. L. & Janssen, L. M. C. (2023). Phys. Rev. Res. 5, 033120.]).

Beyond providing an enhanced qualitative understanding of features that can possibly be encountered when analysing experimentally extracted measurable structure factors, the numerical scheme presented in this contribution in principle provides a means to model the scattered intensity of any polydisperse hard-sphere system, provided a model for the single-particle scattering amplitude and an appropriate size distribution is available. Performing fits with such advanced models directly on experimentally observed intensities gives access to the underlying partial structure factors, enabling a characterization and possible further theoretical analysis on a genuine multi-component foundation, rather than employing effective one-component approaches. The current restriction to hard-sphere interactions is a major incentive to promote advancements in the analytical evaluation of partial structure factors for other interaction potentials, since numerically solving integral equations or employing computer simulations with reasonable statistics are currently only realistically feasible for a restricted number of components, especially in mixtures with large size disparities (Allahyarov et al., 2022[Allahyarov, E., Löwen, H. & Denton, A. R. (2022). Phys. Chem. Chem. Phys. 24, 15439-15451.]).

APPENDIX A

Percus–Yevick structure factors for hard-sphere mixtures

The analytical solution of the Ornstein–Zernike equation for the hard-sphere potential within the Percus–Yevick closure in terms of the partial structure factors Sαβ(Q), presented by Vrij (1979[Vrij, A. (1979). J. Chem. Phys. 71, 3267-3270.]) and reformulated by Voigtmann (2003[Voigtmann, T. (2003). Mode Coupling Theory of the Glass Transition in Binary Mixtures, Technische Universität München.]), is restated here. In short, an expression for the partial direct correlation functions cαβ(r) in real space can be found using Baxter's factorization technique (Baxter, 1970[Baxter, R. J. (1970). J. Chem. Phys. 52, 4559-4562.]). The transformed solution in wavevector space cαβ(Q) can subsequently be used to obtain the partial structure factors Sαβ(Q).

Let φ be the total volume fraction of all spheres, dα be the diameter and xα be the number fraction of the spheres of species α. The total number density ρ of the system is related to the volume fraction by [\varphi = (\pi/6)\rho\sum x_{\alpha}d_{\alpha}^{3}]. With the abbreviations

[\displaystyle d_{\alpha\beta} = {{d_{\alpha}+d_{\beta}}\over{2}} \eqno (17)]

and

[\displaystyle\xi_{x} = {{\pi}\over{6}}\rho\sum_{\gamma}x_{\gamma}d_{\gamma}^{x},\eqno (18)]

the set of coefficients

[\displaystyle a_{\alpha} = {{1-\xi_{3}+3d_{\alpha}\xi_{2}} \over {\left(1-\xi_{3} \right)^{2}}},\eqno (19)]

[\displaystyle\tilde{a}_{2} = {\textstyle\sum\limits_{\gamma}}\rho_{\gamma}a_{\gamma}^{2},\eqno (20)]

[\displaystyle\hat{\beta_{0}} = {{9\xi_{2}^{2}+3\xi_{1}(1-\xi_{3})} \over {(1-\xi_{3 })^{3}}},\eqno (21)]

[\displaystyle A_{\alpha\beta} = {{d_{\alpha\beta}(1-\xi_{3})+{{3} \over {2}}d_{ \alpha}d_{\beta}\xi_{2}} \over {(1-\xi_{3})^{2}}},\eqno (22)]

[\displaystyle B_{\alpha\beta} = {{1} \over {1-\xi_{3}}}-\hat{\beta}_{0}d_{\alpha}d_{\beta}\eqno (23)]

and

[\displaystyle D_{\alpha\beta} = {{6\xi_{2}+12d_{\alpha\beta}[\xi_{1}+3\xi_{2 }^{2}/(1-\xi_{3})]} \over {(1-\xi_{3})^{2}}} \eqno (24)]

can be determined. Furthermore, by introducing [S_{\alpha} = \sin(Qd_{\alpha}/2)] and [C_{\alpha} = \cos(Qd_{\alpha}/2)], the terms

[\displaystyle\mu_{A} = A_{\alpha\beta}{{S_{\alpha}S_{\beta}-C_{\alpha}C_{ \beta}} \over {Q^{2}}}, \eqno (25)]

[\displaystyle\mu_{B} = B_{\alpha\beta}{{C_{\alpha}S_{\beta}+C_{\beta}S_{ \alpha}} \over {Q^{3}}}, \eqno (26)]

[\displaystyle\mu_{D} = D_{\alpha\beta}{{S_{\alpha}S_{\beta}} \over {Q^{4}}} \eqno (27)]

and

[\displaystyle\tilde{\mu} = {{4\pi} \over {Q^{4}}}\tilde{a}_{2}\biggl({{C_{ \alpha}C_{\beta}d_{\alpha}d_{\beta}} \over {4}}+{{S_{\alpha}S_{\beta}} \over {Q^{2}}}-{{C_{\alpha}S_{\beta}d_{\alpha}+C_{\beta}S_{\alpha}d_{\beta} } \over {2Q}}\biggr) \eqno (28)]

can be calculated, which finally leads to

[\displaystyle c_{\alpha\beta}(q) = -4\pi\,(\mu_{A}+\mu_{B}+\mu_{D}+\tilde{\mu}).\eqno (29)]

The partial direct correlation functions form the matrix C with elements Cαβ = (xαxβ)1/2cαβ, which is related to the matrix of partial structure factors S by the Ornstein–Zernike relation

[\displaystyle{\bf S} = \left[{ \bf 1}-\rho\,{\bf C}\right]^{-1}. \eqno (30)]

The partial structure factors here are defined within the convention [\lim_{Q\to\infty}S_{\alpha\beta}(Q) = \delta_{\alpha\beta}], where δαβ is the Kronecker delta.

Acknowledgements

Open access funding enabled and organized by Projekt DEAL.

References

First citationAllahyarov, E., Löwen, H. & Denton, A. R. (2022). Phys. Chem. Chem. Phys. 24, 15439–15451.  CrossRef CAS PubMed Google Scholar
First citationAnta, J. A. & Madden, P. A. (1999). J. Phys. Condens. Matter, 11, 6099–6111.  CrossRef CAS Google Scholar
First citationBallauff, M. (2001). Curr. Opin. Colloid Interface Sci. 6, 132–139.  Web of Science CrossRef CAS Google Scholar
First citationBalogh, J., Olsson, U. & Pedersen, J. S. (2007). J. Phys. Chem. B, 111, 682–689.  CrossRef PubMed CAS Google Scholar
First citationBanchio, A. J., Nägele, G. & Ferrante, A. (1998). J. Colloid Interface Sci. 208, 487–499.  CrossRef CAS PubMed Google Scholar
First citationBartlett, P. & Ottewill, R. H. (1992). J. Chem. Phys. 96, 3306–3318.  CrossRef CAS Web of Science Google Scholar
First citationBaxter, R. J. (1970). J. Chem. Phys. 52, 4559–4562.  CrossRef CAS Web of Science Google Scholar
First citationBeenakker, C. W. J. & Mazur, P. (1983). Physica A, 120, 388–410.  CrossRef Google Scholar
First citationBeenakker, C. W. J. & Mazur, P. (1984). Physica A, 126, 349–370.  CrossRef Web of Science Google Scholar
First citationBeurten, P. van & Vrij, A. (1981). J. Chem. Phys. 74, 2744–2748.  Google Scholar
First citationBlum, L. & Stell, G. (1979). J. Chem. Phys. 71, 42–46.  CrossRef CAS Web of Science Google Scholar
First citationBlum, L. & Stell, G. (1980). J. Chem. Phys. 72, 2212.  CrossRef Google Scholar
First citationBohren, C. F. & Huffman, D. R. (2008). Absorption and Scattering of Light by Small Particles. New York: John Wiley.  Google Scholar
First citationBotet, R., Kwok, S. & Cabane, B. (2020). J. Appl. Cryst. 53, 1570–1582.  CrossRef CAS IUCr Journals Google Scholar
First citationD'Aguanno, B. (1993). Phys. Scr. T49A, 84–88.  CAS Google Scholar
First citationD'Aguanno, B. & Klein, R. (1992). Phys. Rev. A, 46, 7652–7656.  CAS PubMed Google Scholar
First citationDekker, F., Kuipers, B. W. M., González García, Á., Tuinier, R. & Philipse, A. P. (2020). J. Colloid Interface Sci. 571, 267–274.  CrossRef CAS PubMed Google Scholar
First citationDiaz Maier, J. & Wagner, J. (2024). Soft Matter, 20, 1309–1319.  CrossRef CAS PubMed Google Scholar
First citationDi Cola, E., Moussaïd, A., Sztucki, M., Narayanan, T. & Zaccarelli, E. (2009). J. Chem. Phys. 131, 144903.  CrossRef PubMed Google Scholar
First citationFlory, P. J. (1936). J. Am. Chem. Soc. 58, 1877–1885.  CrossRef CAS Google Scholar
First citationFrenkel, D., Vos, R. J., de Kruif, C. G. & Vrij, A. (1986). J. Chem. Phys. 84, 4625–4630.  CrossRef CAS Web of Science Google Scholar
First citationGenz, U. & Klein, R. (1991). Physica A, 171, 26–42.  CrossRef Google Scholar
First citationGinoza, M. & Yasutomi, M. (1999). J. Phys. Soc. Jpn, 68, 2292–2297.  CrossRef CAS Google Scholar
First citationGlatter, O. (2018). Scattering Methods and their Application in Colloid and Interface Science. Amsterdam: Elsevier.  Google Scholar
First citationGreene, D. G., Ferraro, D. V., Lenhoff, A. M. & Wagner, N. J. (2016). J. Appl. Cryst. 49, 1734–1739.  Web of Science CrossRef CAS IUCr Journals Google Scholar
First citationGriffith, W. L., Triolo, R. & Compere, A. L. (1987). Phys. Rev. A, 35, 2200–2206.  CrossRef CAS PubMed Web of Science Google Scholar
First citationHallett, J. E., Grillo, I. & Smith, G. N. (2020). Langmuir, 36, 2071–2081.  Web of Science CrossRef CAS PubMed Google Scholar
First citationHansen, J.-P., Levesque, D. & Zinn-Justin, J. (1991). Liquids, Freezing and Glass Transition. Amsterdam: North Holland.  Google Scholar
First citationHansen, J.-P. & McDonald, I. R. (2013). Theory of Simple Liquids with Applications to Soft Matter. Amsterdam: Elsevier.  Google Scholar
First citationHayter, J. B. & Penfold, J. (1981). Mol. Phys. 42, 109–118.  CrossRef CAS Web of Science Google Scholar
First citationJanssen, L. M. C. (2018). Front. Phys. 6, 97.  CrossRef Google Scholar
First citationKarg, M., Pich, A., Hellweg, T., Hoare, T., Lyon, L. A., Crassous, J. J., Suzuki, D., Gumerov, R. A., Schneider, S., Potemkin, I. I. & Richtering, W. (2019). Langmuir, 35, 6231–6255.  CrossRef CAS PubMed Google Scholar
First citationKirkwood, J. G. & Boggs, E. M. (1942). J. Chem. Phys. 10, 394–402.  CrossRef CAS Google Scholar
First citationKotlarchyk, M. & Chen, S.-H. (1983). J. Chem. Phys. 79, 2461–2469.  CrossRef CAS Web of Science Google Scholar
First citationKrause, R., D'Aguanno, B., Mendez-Alcaraz, J. M., Nägele, G., Klein, R. & Weber, R. (1991). J. Phys. Condens. Matter, 3, 4459–4475.  CrossRef CAS Google Scholar
First citationLi, T., Senesi, A. J. & Lee, B. (2016). Chem. Rev. 116, 11128–11180.  Web of Science CrossRef CAS PubMed Google Scholar
First citationLiu, Y., Chen, W.-R. & Chen, S.-H. (2005). J. Chem. Phys. 122, 044507.  Web of Science CrossRef Google Scholar
First citationLu, P. J. & Weitz, D. A. (2013). Annu. Rev. Condens. Matter Phys. 4, 217–233.  CAS Google Scholar
First citationMason, T. G., Graves, S. M., Wilking, J. N. & Lin, M. Y. (2006). J. Phys. Chem. B, 110, 22097–22102.  CrossRef PubMed CAS Google Scholar
First citationMohanty, P. S., Nöjd, S., van Gruijthuijsen, K., Crassous, J. J., Obiols-Rabasa, M., Schweins, R., Stradner, A. & Schurtenberger, P. (2017). Sci. Rep. 7, 1487.  CrossRef PubMed Google Scholar
First citationNayeri, M., Zackrisson, M. & Bergenholtz, J. (2009). J. Phys. Chem. B, 113, 8296–8302.  CrossRef PubMed CAS Google Scholar
First citationNinarello, A., Berthier, L. & Coslovich, D. (2017). Phys. Rev. X, 7, 021039.  Google Scholar
First citationOlver, F. W. J., Lozier, D. W., Boisvert, R. F. & Clark, C. W. (2010). The NIST Handbook of Mathematical Functions. Cambridge University Press.  Google Scholar
First citationPedersen, J. S. (1994). J. Appl. Cryst. 27, 595–608.  CrossRef Web of Science IUCr Journals Google Scholar
First citationPedersen, J. S. (1997). Adv. Colloid Interface Sci. 70, 171–210.  CrossRef CAS Web of Science Google Scholar
First citationPedersen, J. S. (2001). J. Chem. Phys. 114, 2839–2846.  Web of Science CrossRef CAS Google Scholar
First citationPeláez-Fernández, M., Moncho-Jordá, A. & Callejas-Fernández, J. (2011). J. Chem. Phys. 134, 054905.  PubMed Google Scholar
First citationPercus, J. K. & Yevick, G. J. (1958). Phys. Rev. 110, 1–13.  CrossRef CAS Web of Science Google Scholar
First citationPhalakornkul, J. K., Gast, A. P., Pecora, R., Nägele, G., Ferrante, A., Mandl-Steininger, B. & Klein, R. (1996). Phys. Rev. E, 54, 661–675.  CrossRef CAS Web of Science Google Scholar
First citationPihlajamaa, I., Laudicina, C. C. L. & Janssen, L. M. C. (2023). Phys. Rev. Res. 5, 033120.  CrossRef Google Scholar
First citationPusey, P. N., Fijnaut, H. M. & Vrij, A. (1982). J. Chem. Phys. 77, 4270–4281.  CrossRef CAS Web of Science Google Scholar
First citationRobert, A., Wagner, J., Härtl, W., Autenrieth, T. & Grübel, G. (2008). Eur. Phys. J. E, 25, 77–81.  Web of Science CrossRef PubMed CAS Google Scholar
First citationRoyall, C. P., Poon, W. C. K. & Weeks, E. R. (2013). Soft Matter, 9, 17–27.  CrossRef CAS Google Scholar
First citationSalgi, P. & Rajagopalan, R. (1993). Adv. Colloid Interface Sci. 43, 169–288.  CrossRef CAS Google Scholar
First citationScheffold, F. & Mason, T. G. (2009). J. Phys. Condens. Matter, 21, 332102.  CrossRef PubMed Google Scholar
First citationSchulz, G. V. (1939). Z. Phys. Chem. 43B, 25–46.  CrossRef Google Scholar
First citationSciortino, F., Mossa, S., Zaccarelli, E. & Tartaglia, P. (2004). Phys. Rev. Lett. 93, 055701.  Web of Science CrossRef PubMed Google Scholar
First citationScotti, A. (2021). Soft Matter, 17, 5548–5559.  CrossRef CAS PubMed Google Scholar
First citationStellbrink, J., Allgaier, J., Richter, D., Moussaid, A., Schofield, A. B., Poon, W. C. K., Pusey, P. N., Lindner, P., Dzubiella, J., Likos, C. N. & Löwen, H. (2002). Appl. Phys. Mater. Sci. Process. 74, s355–s357.  CrossRef CAS Google Scholar
First citationStieger, M., Pedersen, J. S., Lindner, P. & Richtering, W. (2004). Langmuir, 20, 7283–7292.  Web of Science CrossRef PubMed CAS Google Scholar
First citationStieger, M., Richtering, W., Pedersen, J. S. & Lindner, P. (2004). J. Chem. Phys. 120, 6197–6206.  Web of Science CrossRef PubMed CAS Google Scholar
First citationSzymusiak, M., Kalkowski, J., Luo, H., Donovan, A. J., Zhang, P., Liu, C., Shang, W., Irving, T., Herrera-Alonso, M. & Liu, Y. (2017). ACS Macro Lett., 6, 1005–1012.  CrossRef CAS PubMed Google Scholar
First citationTomchuk, O. V., Bulavin, L. A., Aksenov, V. L., Garamus, V. M., Ivankov, O. I., Vul', A. Y., Dideikin, A. T. & Avdeev, M. V. (2014). J. Appl. Cryst. 47, 642–653.   CrossRef CAS IUCr Journals Google Scholar
First citationVoigtmann, T. (2003). Mode Coupling Theory of the Glass Transition in Binary Mixtures, Technische Universität München.  Google Scholar
First citationVrij, A. (1978). J. Chem. Phys. 69, 1742–1747.  CrossRef CAS Web of Science Google Scholar
First citationVrij, A. (1979). J. Chem. Phys. 71, 3267–3270.  CrossRef CAS Web of Science Google Scholar
First citationWagner, J. (2004). J. Appl. Cryst. 37, 750–756.  Web of Science CrossRef CAS IUCr Journals Google Scholar
First citationWagner, J. (2012). J. Appl. Cryst. 45, 513–516.  CrossRef CAS IUCr Journals Google Scholar
First citationWidom, B. (1967). Science, 157, 375–382.  CrossRef PubMed CAS Google Scholar
First citationZaccarelli, E., Liddle, S. M. & Poon, W. C. K. (2015). Soft Matter, 11, 324–330.  CrossRef CAS PubMed Google Scholar
First citationZackrisson, M., Stradner, A., Schurtenberger, P. & Bergenholtz, J. (2005). Langmuir, 21, 10835–10845.  Web of Science CrossRef PubMed CAS Google Scholar
First citationZhou, B., Gasser, U. & Fernandez-Nieves, A. (2023). Phys. Rev. E, 108, 054604.  CrossRef PubMed Google Scholar

This is an open-access article distributed under the terms of the Creative Commons Attribution (CC-BY) Licence, which permits unrestricted use, distribution, and reproduction in any medium, provided the original authors and source are cited.

Journal logoJOURNAL OF
APPLIED
CRYSTALLOGRAPHY
ISSN: 1600-5767
Follow J. Appl. Cryst.
Sign up for e-alerts
Follow J. Appl. Cryst. on Twitter
Follow us on facebook
Sign up for RSS feeds