Transmission matrices used in epidemiologic modelling (2024)

As a library, NLM provides access to scientific literature. Inclusion in an NLM database does not imply endorsem*nt of, or agreement with, the contents by NLM or the National Institutes of Health.
Learn more: PMC Disclaimer | PMC Copyright Notice

Transmission matrices used in epidemiologic modelling (1)

Infectious Disease Modelling

Infect Dis Model. 2024 Mar; 9(1): 185–194.

Published online 2023 Dec 4. doi:10.1016/j.idm.2023.11.009

PMCID: PMC10796975

PMID: 38249428

Author information Article notes Copyright and License information PMC Disclaimer

Abstract

Mixing matrices are included in infectious disease models to reflect transmission opportunities between population strata. These matrices were originally constructed on the basis of theoretical considerations and most of the early work in this area originates from research on sexually transferred diseases in the 80s, in response to AIDS. Later work in the 90s populated these matrices on the basis of survey data gathered to capture transmission risks for respiratory diseases. We provide an overview of developments in the construction of matrices for capturing transmission opportunities in populations. Such transmission matrices are useful for epidemiologic modelling to capture within and between stratum transmission and can be informed from theoretical mixing assumptions, informed by empirical evidence gathered through investigation as well as generated on the basis of data. Links to summary measures and threshold conditions are also provided.

Keywords: Disease transmission, Mixing matrix, Social contact, Index of disassortativity

1. Introduction

Epidemiologic models capture disease transmission in a population. This disease transmission depends on the interactions between members of the population. Where susceptible hosts interact with infected carriers, the disease can transmit from the latter to the former. These interactions are a key component of modelling approaches. This means epidemiologic modellers are interested in knowing how people move and who encounters who.

The simplest assumption of transmission is hom*ogenous mixing, meaning that all members of the population mix randomly with each other with an equally likely probability. For a stratified population, if the mixing is random within strata it is known as proportional mixing, this is where the mixing between strata is proportional to the size of the population stratum (Jacquez etal., 1988; Sattenspiel and Simon, 1988). One method of capturing and including interactions between population strata (mixing) is transmission matrices. hom*ophily in transmission matrices occurs when the diagonal elements of the transmission matrix are greater than the off-diagonal elements, which is known as assortative mixing (is also known as like-with-like mixing). When this is not the case, the mixing is considered to be disassortative (like-with-unlike) mixing, which is also sometimes referred to as “negative assortative” mixing (). Proportionate mixing falls between the two (; ). Separable mixing is when mixing is independent between strata (; Goeyvaerts etal., 2010).

Understanding human interaction is of interest for multiple scientific disciplines; anthropology, epidemiology, sociology, urban planning, geography, and ecology and so there are many options for populating transmission matrices. Here we would like to provide an overview of transmission matrices, hopefully showing the links between methodological development arising from various disciplines. We highlight the main papers making up the current state of knowledge on transmission matrices as well as approaches on the topic to provide a foundation for other researchers.

In the following we consider matrices of rates of interaction between strata i and j with entries 0≤pij ≤ 1, but if counts of interactions m are provided, rates can be calculated by normalisation: pij=mij/∑mij.

2. Theoretical transmission matrices

For n population strata, an n × n transmission matrix is needed. This means n2 transmission rates could be expected but certain simplifications (such as expectations of symmetry) allow for fewer rates needed to be determined.

2.1. Interaction

Many of the early considerations for mixing hence transmission matrices came from models for sexually transferred infections (STIs) and later developments arose from models for respiratory infections. A matrix used for STI modelling explains “who has sex with whom” (WHSWW) (Aral etal., 1999; ). In their early gonorrhoea modelling approach, Hethcote, Yorke, and Nold (1982) used a WHSWW matrix with certain heterogeneous mixing assumptions.

Such a matrix can be constructed in myriad ways. One line of thought for the construction of theoretical transmission matrices arose from the force of infection (see Estimated transmission matrices). The following options have been considered (; ; van Effelterre etal., 2009) but are not exhaustive. Examples are given for n=5.

Including the introduction of a multiplier α which could be used in situations of low case counts (Gay etal., 1995; Klepac etal., 2009) yields additional options

2.2. Mobility

For theoretic movement of people across space (i.e. migration matrices), options considered include stratifying the population and considering the interactions between those strata. A simplification for when only part of the population mixes is the 2n × 2n matrix of transmission between strata (Sattenspiel, 1988)

where only Ms mixes and can be given by random mixing (a matrix where all rates are equal), purely assortative (a diagonal matrix) or

Here the top left denotes a smaller group of size r × r(r<n) that mixes with one rate (p1=1/r) with random mixing (p2=1/n) occurring elsewhere in the population.

An alternative starting point for spatial models is the distance matrix (Bapat, 2014) (known as neighbourhood order (Meyer and Held, 2014)).

3. Empirical transmission matrices

The use of the theoretical matrices introduced in the previous section requires an active choice on the part of the modeller and the choice introduces certain assumptions into the modelling. For this reason, data-driven approaches to inform transmission matrices have seen increased use in the literature.

3.1. Interaction

Contact matrices are a form of sociomatrix introduced in Schenzle (1984) and later developments led to pilot studies for populating “who acquires infection from whom” (WAIFW) matrices () based in survey data. The landmark social contact survey is POLYMOD (Mossong etal., 2008) which is often considered the catalyst and blueprint for later contact diary surveys for respiratory infections. An overview of later surveys is found in the work by Hoang etal. (2019) and Mousa etal. (2021).

Recent developments include the distinction between symptomatic and asymptomatic infections (Santermans etal., 2017), urban and rural settings (Del Fava etal., 2021; Leung etal., 2023), biological sex (van de Kassteele etal., 2017), the specific contacts of elderly (often defined epidemiologically as those above 65 years of age) (Oh etal., 2021) which introduces elderly facilities as an additional contact setting, contacts of school children (Conlan etal., 2011; Stehlé etal., 2011), contacts of internally-displaced people (van Zandvoort etal., 2022) which introduces the setting of a contact in other houses than ones own, contacts in refugee camps (Walker etal., 2022) which adds contacts in places such as pumps, shelters, and food distribution centres, as well as contacts specific to airborne transmission (McCreesh etal., 2022), contacts between healthcare workers (Machens etal., 2013), and by HIV status (Thindwa etal., 2022).

Developments such as considering airborne diseases broadens the earlier focus on respiratory transmission which might not be applicable to all infections as some do not require situations of intimate or immediate contact to spread. The urban-rural divide (Del Fava etal., 2021; Leung etal., 2023) may depend on geographical setting, as other research groups have found little difference in urban and rural contacts (Read etal., 1785). Empirical surveys conducted during synthetic outbreaks have also been evaluated (Klepac etal., 2018). Additionally time-use survey data can be used in place of contact diary surveys to construct contact matrices (; Zagheni etal., 2008) though the performance may depend on properties of the pathogen (Hoang etal., 2022). As with WAIFW, the empirical WHSWW can also be populated from survey data (Haraldsdottir etal., 1992; ; ; ; ). Additionally, “who injects with whom” matrices can be useful transmission matrices for blood-borne infections (Smith etal., 2018) as these focus on the population at risk of acquiring this kind of infection (people who inject drugs). Godin etal. () considered a theoretical contact matrix for an incarcerated population and populated it with survey data. WHSWW matrices have been used to examine the impact of condom use ().

3.2. Mobility

Within urban planning, origin-destination matrices are used to capture and quantify the flow of people from a place of origin to a destination. An example is a matrix with entries (Hufnagel etal., 2004)

pij=tijjtijNiitij

(9)

where tij denotes the number of passengers with an origin in transport hub i (with Ni people living in its catchment area) and destination transport hub j. The gravity model (Balcan etal., 2009; Xia etal., 2004) considers movement given by

mijNiNjf(dij)

(10)

where Ni is the population in location i and Nj is the population in location j. Options for the transformation of the distance dij in the denominator include f(x)=x (identity), f(x)=exp(x) (exponential), and f(x)=xδ (power law). The function f is called an impedance function as it represents the reluctance to make the journey whereby we see that the gravity model makes a trade-off between the amount of mass (hence attraction) in population and the distance required.

The power law function has also been applied to adjacencies (spatial interaction) in spatial distance (Meyer and Held, 2014) as well as empirical commuting volume data (origin-destination) from air-, road-, and sea-based travel (). The gravity model is known as a flux model. Another option of modelling flux is the radiation model (Kraemer etal., 2019; Simini etal., 2012)

mijNiNj(Ni+Rij)(Ni+Nj+Rij)

(11)

which includes the population in the radius between i and j denoted Rij. Bjørnstad etal. compared many of these options for pre-vaccination measles surveillance data and found that the gravity and radiation models might not be the best option.

4. Estimated transmission matrices

Balance constraints ensure symmetry in certain empirical or estimated transmission matrices; at population strata level the contacts should be symmetric (Hyman and Li, 1997) also known as requirements of reciprocity. Williams and Anderson (1994) suggest various methods of how to achieve a balanced transmission matrix, including the introduction of an iterative approach adjusting values of the transmission matrix until the balance is below a threshold to ensure balancing is achieved.

4.1. Interaction

For diseases which confer life-long immunity, serological investigation can be used to determine the force of infection. In this setting, the theoretical transmission matrices (see Theoretical transmission matrices) as well as later developments (see Empirical transmission matrices) have been utilised. Serology is concerned with the quantification of anti-bodies in the blood. If the anti-bodies are present above a pre-specified threshold, the serological test is said to be positive and the person tested is said to be sero-positive, meaning they have had a prior infection. The curve of proportion of results which are seropositive (hence immune) should increase with age and can be determined from a serology survey. For an age-structured approach, the force of infection can be estimated from serology data. The force of infection is the rate at which individuals susceptible to the disease become infected (it is the flow from the S compartment to the I compartment in a traditional SIR compartmental model).

The social contact hypothesis considers disease transmission rates to be proportional to contacts () and so the serological data (force of infection) can be used with either a theoretical or empirical transmission matrix to estimate the underlying transmission of the disease (Goeyvaerts etal., 2010; van Effelterre etal., 2009; Zagheni etal., 2008). These approaches assume time hom*ogeneity and endemic equilibrium. Fumanelli, Ajelli, Manfredi, Vespignani, and Merler (2012) showed how to augment and construct empirical transmission matrices on the basis of demographic data, providing an approach that has been used to construct matrices where an empirical survey has not been conducted (; Mistry etal., 2021; Prem etal., 2017). The use of online data gathering as an alternative to a contact diary survey has also been considered and evaluated ().

4.2. Mobility

Sattenspiel (Sattenspiel, 1996) suggests a mobility matrix with entries mij informed by rates of travel. Mobile phone data can be used to determine spatial transmission (Tizzoni etal., 2014).

5. Summary measures for transmission matrices

The Q-statistic compares the diagonal elements of a transmission matrix pii (Aral etal., 1999) or the eigenvalues wi () with its dimension (here we consider matrices of dimension n × n). It is calculated as (Aral etal., 1999)

Q=ipii1n1

(12)

or (Gupta etal., 1989)

Q=iwi1n1

(13)

If Q>0, the transmission matrix exhibits assortative mixing, if Q=0 the transmission matrix exhibits random mixing, and if Q<0 the transmission matrix exhibits disassortative mixing. The trace of the transmission matrix ipii itself provides information on the size of the within-stratum (peer) mixing. The lower limit is Q=1n1 (). Some researchers have used the limits of Q (1n1 and 1) as a visual aid to plot where their matrix falls on a scale from disassortative to assortative mixing (Garnett etal., 1996).

A similar summary measure was proposed by Newman (2003)

r=ipiiipi+p+i1ipi+p+i

(14)

where pi+=jpij is the sum of the ith row in the matrix and p+i=ipij is the sum of the ith column in the matrix. This measure takes the value r=1 for assortative mixing and r=0 for random mixing.

An alternative measure of assortative mixing is (Keeling and Rohani, 2008)

q=w1w2

(15)

which considers the ratio of the largest and second largest eigenvalues (w1 and w2, respectively).

Adjacency matrices induced from networks (Iacobucci, 1994; Kretzschmar and Morris, 1996) may take the same Q for different networks. An example is given in Fig.1 where Q has a value of −1/8 (no assortative mixing) for all three adjacency matrices since their diagonal entries all take values 0, though the mixing between population strata is rather different.

Transmission matrices used in epidemiologic modelling (10)

Examples of networks from Kretzschmar and Morris (1996) (left) and their associated adjacency matrices (right). All take the value of Q at the lower limit.

For this reason, the perspective of not-assortative mixing has been considered; while the Q, r, and q statistics consider assortative mixing, the I2 statistic (; Wallinga etal., 2019) examines instead the disassortative mixing. It is given by

I2=E(ij)22σ2

(16)

where i and j are the values of the groups, σ2 is the variance associated with the overall distribution of the groups, and E denotes the expectation with respect to the density of pairs i and j.

A summary measure such as those outlined above may provide a useful manner of summarising many transmission matrices at once, for instance in approaches where they take on a temporal dimension and vary over time. The summary measure values give an overview of how the transmission matrices may change and may provide a simpler option for a graphical representation of the properties of the transmission matrices. An alternative is to plot the value of specific matrix entries to reflect potential changes to a transmission matrix (Meyer and Held, 2016).

5.1. Interaction

Transmission matrices may be updated to reflect changes in demography (Arregui etal., 2018) or sampled many times during an outbreak setting [examples from the COVID-19 pandemic of the 2020s include: Jarvis etal., 2020; ; Liu etal., 2021].

5.2. Mobility

Transmission matrices be updated by augmenting a transmission weight with an external data source which varies over time such as data tracking changes in mobility [examples from the COVID-19 pandemic of the 2020s include: Alleman etal., 2021; ; ].

6. Access

The usefulness of empirical or estimated transmission matrices is greatest when they are made available to other researchers. In order to maximise the usefulness of the empirical matrices, they should be provided as open data. We provide an overview of the openness of options mentioned in this manuscript in Table 1. Programming language-specific initiatives to enable easy access to empirical transmission matrices have been found for R (e.g. the “socialmixr” package (Willem etal., 2020)) but not for other languages used in epidemiologic modelling.

Table 1

Availability of certain empirical or estimated matrices based on original manuscripts (as of November 2023).

SourceAccessibilityAvailability for re-use
(article)(matrix)
Haraldsdottir etal. (1992)free accesspartly available; values provided in figure
Del Valle, Hyman, Hethcote, and Eubank (2007)subscriptionpartly avaialble; values provided in table
Garnett and Anderson (1993)free accesspartly available; values provided in figure
Schenzle (1984)subscriptionnot available
Edmunds etal. (1997)subscriptionnot available
Mossong etal. (2008)open accessnot available
Santermans etal. (2017)subscriptionnot available
Del Fava etal. (2021)open accessopenly available; article provides link to OSF repository (osf.io/ebwrh)
Leung etal. (2023)open accessopenly available; article provides link to Zenodo repository (zenodo.org/record/7642648)
van de Kassteele etal. (2017)open accessopenly available; provided as supporting information
Oh etal. (2021)subscriptionlimited availability; only available upon request
Conlan etal. (2011)open accessnot available; article provides link to website which does not contain data set
Stehlé etal. (2011)open accessopenly available; values provided in table
van Zandvoort etal. (2022)open accessopenly available; article provides link to GitHub repository (github.com/kevinvzandvoort/espicc_somaliland_digaale_contact_survey_2019)
Walker etal., 2022open accessopenly available; article provides link to GitHub repository (github.com/IDAS-Durham/june_mixed_method_CM_results)
McCreesh etal. (2022))open accessopenly available; article provides link to GitHub repository (github.com/NickyMcC/CasualAgeMixing)
Machens etal. (2013)open accessopenly available; values provided in table
Thindwa etal. (2022)open accessopenly available; article provides link to GitHub repository (github.com/deusthindwa/social.contact.rates.estimation.hiv.malawi) which does not contain the matrix but values provided in figure
Klepac etal. (2018)open accessnot available
Zagheni etal. (2008)free accessnot available
van Leeuwen etal. (2022)subscriptionnot available
Datta etal. (2018)open accesspartly available; part of data set provided as supporting information
Easterly etal. (2018)subscriptionopenly available; article provides link to Zenodo repository (zenodo.org/record/1322780)
Granath etal. (1991)free accesspartly available; values provided in table
Fumanelli etal. (2012)open accessopenly available; provided as supporting information
Prem etal. (2017)open accessopenly available; provided as supporting information
Arregui etal. (2018)open accessnot available
Mistry etal. (2021)open accessopenly available; article provides link to GitHub repository (github.com/mobs-lab/mixing-patterns)
Tizzoni etal. (2014)open accessopenly available; provided as supporting information
Jarvis etal. (2020)open accessopenly available; article provides link to GitHub repository (github.com/jarvisc1/comix_covid-19-first_wave)
Feehan and Mahmud (2021)open accessopenly available; article provides link to Harvard Dataverse repository (dataverse.harvard.edu/dataset.xhtml?persistentId=doi:10.7910/DVN/M74AJ4)
Smith etal. (2018)free accesspartly available; values provided in figure
Brooks-Pollock and Danon (2017)free accesspartly available; values provided in table

Having harmonised terms and abbreviations (vocabulary) such as the international classifications of diseases (Harrison etal., 2021) provides an advantage to other users of data sets using the same vocabulary. This advantage comes from consistent identifiers and labels across data sets and so allows for the merging of data sets in a simpler manner than when labels and names in data sets are not unified. Similarly, encountering newer data becomes easier as variables are provided in an established and understandable manner. In short, controlled vocabularies improve the interoperability of data sets.

Mossong etal. (2008) established a naming convention for empirical transmission matrices based on social contact data but this convention is not yet explicitly described in the scientific literature or included in the FAIRsharing repository (fairsharing.org) allowing users to find it easily. In addition, a definite controlled vocabulary would help provide information in situations where meta data can be shared but the survey itself cannot be shared or cannot be accessed such as when data is provided in a proprietary format.

7. Links to epidemiologic thresholds

If the population being considered is fully susceptible to the disease (immunologically naïve), transmission matrices can be used to determine the basic reproduction number. This threshold is a summary measure of the “baseline” ability of the disease to transmit between an infected carrier and a susceptible host. Nold (1980) defined the spectral radius condition for transmission matrices and Diekmann etal. (1990, 2010) linked the spectral radius to the next generation matrix. Both can be used to obtain an estimate of the basic reproduction number. Caution must be taken of the hom*ogenous mixing fallacy which occurs when an estimate of the basic reproduction number is obtained under an assumption of hom*ogenous mixing when the true mixing is heterogeneous. This incorrect assumption leads to an estimate of the basic reproduction number which is too low which in turn leads to an estimate of the herd immunity threshold (the proportion of the population that needs to be vaccinated to keep the incidence of disease under the epidemic threshold) which is too low (Hethcote, 1996). Farrington, Kanaan, and Gay (2001) examined the effects of including individual heterogeneity and Inaba and Nishiura (2008) extended the concept to obtain the effective next generation which leads to an estimate of the effective reproduction number and removes the requirement of a fully susceptible population. Brooks-Pollock and Danon (2017) linked the next generation matrix to the attributable fraction. Vynnycky and White (2010) examined the relationship between the Q statistic and the basic reproduction number and found it possible to determine the value of Q for which the basic reproduction number is at the epidemic threshold.

8. Discussion

Regarding further reading, earlier reviews provide a nice overview of various aspects not considered in great detail here. We refer the interested reader to Keeling and Eames (2005) for a networks-based perspective on transmission (networks can be represented in matrix form, see Haraldsdottir etal. (1992) and Del Valle etal. (2007) for example). The work by Chowell, Sattenspiel, Bansal, and Viboud (2016) examines the interplay between transmission options considered in this work and early epidemic growth. For more on contact matrices, the systematic reviews by Hoang etal. (2019) and Mousa etal. (2021) may provide fruitful insights. An overview of transmission matrices used with a specific class of surveillance-based regression models known as endemic-epidemic models is found in Bekker-Nielsen Dunbar (2023). For more information on serology and mixing, the book by Hens etal. (2012) is an excellent reference. A data repository with multiple social contact surveys in the format established by Mossong etal. (2008) is found in Willem etal. (2020), all of which are openly available and provided for re-use. An extensive review of mobility models which includes options for populating a transmission matrix is given in Barbosa etal. (2018). More information on the next generation matrix can be found in Heffernan etal. (2005) and Hurford etal. (2010).

CRediT authorship contribution statement

M. Bekker-Nielsen Dunbar: Conceptualization, Writing – original draft, Writing – review & editing.

Declaration of competing interest

The authors declare the following financial interests/personal relationships which may be considered as potential competing interests: Research support (salary): OsloMet Oslo Metropolitan University, Employment: OsloMet Oslo Metropolitan University.

Notes

Handling Editor: Dr. Raluca Eftimie

References

  • Alleman T.W., Vergeynst J., De Visscher L., Rollier M., Torfs E., Nopens I., et al. Assessing the effects of non-pharmaceutical interventions on SARS-CoV-2 transmission in Belgium by means of an extended SEIQRD model and public mobility data. Epidemics. 2021;37 doi:10.1016/j.epidem.2021.100505. [PMC free article] [PubMed] [CrossRef] [Google Scholar]
  • Anderson R.M., May R.M. Age-related changes in the rate of disease transmission: Implications for the design of vaccination programmes. Journal of Hygiene. 1985;94(3):365–436. doi:10.1017/s002217240006160x. [PMC free article] [PubMed] [CrossRef] [Google Scholar]
  • Aral S.O., Hughes J.P., Stoner B., Whittington W., Handsfield H.H., Anderson R.M., et al. Sexual mixing patterns in the spread of gonococcal and chlamydial infections. American Journal of Public Health. 1999;89(6):825–833. doi:10.2105/AJPH.89.6.825. [PMC free article] [PubMed] [CrossRef] [Google Scholar]
  • Arregui S., Aleta A., Sanz J., Moreno Y. Projecting social contact matrices to different demographic structures. PLoS Computational Biology. 2018;14(12):1–18. doi:10.1371/journal.pcbi.1006638. [PMC free article] [PubMed] [CrossRef] [Google Scholar]
  • Ashby B., Gupta S. Sexually transmitted infections in polygamous mating systems. Philosophical Transactions of the Royal Society B: Biological Sciences. 2013;368(1613) doi:10.1098/rstb.2012.0048. [PMC free article] [PubMed] [CrossRef] [Google Scholar]
  • Balcan D., Colizza V., Gonçalves B., Hu H., Ramasco J.J., Vespignani A. Multiscale mobility networks and the spatial spreading of infectious diseases. Proceedings of the National Academy of Sciences. 2009;106(51):21484–21489. doi:10.1073/pnas.0906910106. [PMC free article] [PubMed] [CrossRef] [Google Scholar]
  • Bapat R.P. Springer; 2014. Distance matrix of a tree in “Graphs and Matrices” pp. 115–131. [CrossRef] [Google Scholar]
  • Barbosa H., Barthelemy M., Ghoshal G., James C.R., Lenormand M., Louail T., et al. Human mobility: Models and applications. Physics Reports. 2018;734:1–74. doi:10.1016/j.physrep.2018.01.001. [CrossRef] [Google Scholar]
  • Bekker-Nielsen Dunbar M. 2023. Time-varying transmission weights in endemic-epidemic models. [CrossRef] [Google Scholar]
  • O. N. Bjørnstad, B. T. Grenfell, C. Viboud, A. A. King, Comparison of alternative models of human movement and the spread of disease [pre-print], bioRxiv doi:10.1101/2019.12.19.882175.
  • Blower S.M., Hartel D., Dowlatabadi H., Anderson R.M., May R.M. Drugs, sex and HIV: A mathematical model for New York city. Philosophical Transactions of the Royal Society of London - Series B: Biological Sciences. 1991;331(1260):171–187. doi:10.1098/rstb.1991.0006. [PubMed] [CrossRef] [Google Scholar]
  • Blythe S.P., Castillo-Chavez C., Palmer J.S., Cheng M. Toward a unified theory of sexual mixing and pair formation. Mathematical Biosciences. 1991;107(2):379–405. doi:10.1016/0025-5564(91)90015-B. [PubMed] [CrossRef] [Google Scholar]
  • Brooks-Pollock E., Danon L. Defining the population attributable fraction for infectious diseases. International Journal of Epidemiology. 2017;46(3):976–982. doi:10.1093/ije/dyx055. [PMC free article] [PubMed] [CrossRef] [Google Scholar]
  • Chowell G., Sattenspiel L., Bansal S., Viboud C. Mathematical models to characterize early epidemic growth: A review. Physics of Life Reviews. 2016;18:66–97. doi:10.1016/j.plrev.2016.07.005. [PMC free article] [PubMed] [CrossRef] [Google Scholar]
  • Conlan A.J.K., Eames K.T.D., Gage J.A., von Kirchbach J.C., Ross J.V., Saenz R.A., et al. Measuring social networks in British primary schools through scientific engagement. Proceedings of the Royal Society B. 2011;278:1467–1475. doi:10.1098/rspb.2010.1807. [PMC free article] [PubMed] [CrossRef] [Google Scholar]
  • Datta S., Mercer C.H., Keeling M.J. Capturing sexual contact patterns in modelling the spread of sexually transmitted infections: Evidence using Natsal-3. PLoS One. 2018;13(11):1–18. doi:10.1371/journal.pone.0206501. [PMC free article] [PubMed] [CrossRef] [Google Scholar]
  • Del Fava E., Adema I., Kiti M.C., Poletti P., Merler S., Nokes D.J., et al. Individual's daily behaviour and intergenerational mixing in different social contexts of Kenya. Scientific Reports. 2021;11(1) doi:10.1038/s41598-021-00799-1. [PMC free article] [PubMed] [CrossRef] [Google Scholar]
  • Del Valle S.Y., Hyman J.M., Hethcote H.W., Eubank S.G. Mixing patterns between age groups in social networks. Social Networks. 2007;29(4):539–554. doi:10.1016/j.socnet.2007.04.005. [CrossRef] [Google Scholar]
  • Diekmann O., Heesterbeek J.A.P., Metz J.A.J. On the definition and the computation of the basic reproduction ratio R0 in models for infectious diseases in heterogeneous populations. Journal of Mathematical Biology. 1990;28(4):365–382. doi:10.1007/BF00178324. [PubMed] [CrossRef] [Google Scholar]
  • Diekmann O., Heesterbeek J.A.P., Roberts M.G. The construction of next-generation matrices for compartmental epidemic models. Journal of The Royal Society Interface. 2010:7873–7885. doi:10.1098/rsif.2009.0386. [PMC free article] [PubMed] [CrossRef] [Google Scholar]
  • Easterly C.W., Alarid-Escudero F., Enns E.A., Kulasingam S. Revisiting assumptions about age-based mixing representations in mathematical models of sexually transmitted infections. Vaccine. 2018;36(37):5572–5579. doi:10.1016/j.vaccine.2018.07.058. [PMC free article] [PubMed] [CrossRef] [Google Scholar]
  • Edmunds W.J., O’callaghan C.J., Nokes D.J. Who mixes with whom? A method to determine the contact patterns of adults that may lead to the spread of airborne infections. Proceedings of the Royal Society of London - Series B: Biological Sciences. 1997;264(1384):949–957. doi:10.1098/rspb.1997.0131. [PMC free article] [PubMed] [CrossRef] [Google Scholar]
  • van Effelterre T., Shkedy Z., Aerts M., Molenberghs G., van Damme P., Beutels P. Contact patterns and their implied basic reproductive numbers: An illustration for varicella-zoster virus. Epidemiology and Infection. 2009;137:48–57. doi:10.1017/S0950268808000563. [PubMed] [CrossRef] [Google Scholar]
  • Ensoy-Musoro C., Nguyen M.H., Hens N., Molenberghs G., Faes C. Spatio-temporal model to investigate COVID-19 spread accounting for the mobility amongst municipalities. Spatial and Spatio-temporal Epidemiology. 2023 doi:10.1016/j.sste.2023.100568. [PMC free article] [PubMed] [CrossRef] [Google Scholar]
  • Farrington C.P., Whitaker H.J., Wallinga J., Manfredi P. Measures of disassortativeness and their application to directly transmitted infections. Biometrical Journal. 2009;51(3):387–407. doi:10.1002/bimj.200800160. [PubMed] [CrossRef] [Google Scholar]
  • Farrington M.N., Kanaan C.P., Gay N.J. Estimation of the basic reproduction number for infectious diseases from age-stratified serological survey data. Journal of the Royal Statistical Society: Series C (Applied Statistics) 2001;50:251–292. doi:10.1111/1467-9876.00233. [CrossRef] [Google Scholar]
  • Feehan D.M., Mahmud A.S. Quantifying population contact patterns in the United States during the COVID-19 pandemic. Nature Communications. 2021;12(893) doi:10.1038/s41467-021-20990-2. [PMC free article] [PubMed] [CrossRef] [Google Scholar]
  • Fumanelli L., Ajelli M., Manfredi P., Vespignani A., Merler S. Inferring the structure of social contacts from demographic data in the analysis of infectious diseases spread. PLoS Computational Biology. 2012;8(9):1–10. doi:10.1371/journal.pcbi.1002673. [PMC free article] [PubMed] [CrossRef] [Google Scholar]
  • Garnett G.P., Anderson R.M. Contact tracing and the estimation of sexual mixing patterns: The epidemiology of gonococcal infections. Sexually Transmitted Diseases. 1993;20(4) [PubMed] [Google Scholar]
  • Garnett G.P., Hughes J.P., Anderson R.M., Stoner P., Aral B.S.O., Whittington W.L., et al. Sexual mixing patterns of patients attending sexually transmitted diseases clinics. Sexually Transmitted Diseases. 1996;23(3):248–257. doi:10.1097/00007435-199605000-00015. [PubMed] [CrossRef] [Google Scholar]
  • Gay N.J., Hesketh L.M., Morgan-Capner P., Miller E. Interpretation of serological surveillance data for measles using mathematical models: Implications for vaccine strategy. Epidemiology and Infection. 1995;115(1):139–156. doi:10.1017/S0950268800058209. [PMC free article] [PubMed] [CrossRef] [Google Scholar]
  • Geilhufe M., Held L., Skrøvseth S.O., Simonsen G.S., Godtliebsen F. Power law approximations of movement network data for modeling infectious disease spread. Biometrical Journal. 2014;56(3):363–382. doi:10.1002/bimj.201200262. [PubMed] [CrossRef] [Google Scholar]
  • Godin A., Kronfli N., Cox J., Alary M., Maheu-Giroux M. The role of prison-based interventions for hepatitis C virus (HCV) micro-elimination among people who inject drugs in Montréal, Canada. International Journal of Drug Policy. 2021;88 doi:10.1016/j.drugpo.2020.102738. [PubMed] [CrossRef] [Google Scholar]
  • Goeyvaerts N., Hens N., Ogunjimi B., Aerts M., Shkedy Z., Van Damme P., et al. Estimating infectious disease parameters from data on social contacts and serological status. Journal of the Royal Statistical Society: Series C (Applied Statistics) 2010;59(2):255–277. doi:10.1111/j.1467-9876.2009.00693.x. [CrossRef] [Google Scholar]
  • Granath F., Giesecke J., Scalia-Tomba G., Ramstedt K., Forssman L. Estimation of a preference matrix for women's choice of male sexual partner according to rate of partner change, using partner notification data. Mathematical Biosciences. 1991;107(2):341–348. doi:10.1016/0025-5564(91)90013-9. [PubMed] [CrossRef] [Google Scholar]
  • Grimée M., Bekker-Nielsen Dunbar M., Hofmann F., Held L. Modelling the effect of a border closure between Switzerland and Italy on the spatiotemporal spread of COVID-19 in Switzerland. Spatial Statistics. 2021 doi:10.1016/j.spasta.2021.100552. [PMC free article] [PubMed] [CrossRef] [Google Scholar]
  • Gupta S., Anderson R.M., May R.M. Networks of sexual contacts: Implications for the pattern of spread of HIV. AIDS. 1989;3 doi:10.1097/00002030-198912000-00005. [PubMed] [CrossRef] [Google Scholar]
  • Haraldsdottir S., Gupta S., Anderson R.M. Preliminary studies of sexual networks in a male hom*osexual community in Iceland. JAIDS Journal of Acquired Immune Deficiency Syndromes. 1992;5(4) [PubMed] [Google Scholar]
  • Harrison J.E., Weber S., Jakob R., Chute C.G. ICD-11: An international classification of diseases for the twenty-first century. BMC Medical Informatics and Decision Making. 2021;21(6):206. doi:10.1186/s12911-021-01534-6. [PMC free article] [PubMed] [CrossRef] [Google Scholar]
  • Heffernan J.M., Smith R.J., Wahl L.M. Perspectives on the basic reproductive ratio. Journal of The Royal Society Interface. 2005;22(2):281–293. doi:10.1098/rsif.2005.0042. [PMC free article] [PubMed] [CrossRef] [Google Scholar]
  • Hens N., Shkedy Z., Aerts M., Faes C., Van Damme P., Beutels P. Springer; 2012. Modeling infectious disease parameters based on serological and social contact data. [CrossRef] [Google Scholar]
  • Hethcote H.W. Cambridge University Press; 1996. Modeling heterogeneous mixing in infectious disease dynamics. [CrossRef] [Google Scholar]
  • Hethcote H.W., Yorke J.A., Nold A. Gonorrhea modeling: A comparison of control methods. Mathematical Biosciences. 1982;58(1):93–109. doi:10.1016/0025-5564(82)90053-0. [CrossRef] [Google Scholar]
  • Hoang T., Coletti P., Melegaro A., Wallinga J., Grijalva C.G., Edmunds J.W., et al. Asystematic review of social contact surveys to inform transmission models of close-contact infections. Epidemiology. 2019;30(5):723–736. doi:10.1097/EDE.0000000000001047. [PMC free article] [PubMed] [CrossRef] [Google Scholar]
  • Hoang T.V., Willem L., Coletti P., Van Kerckhove K., Minnen J., Beutels P., et al. Exploring human mixing patterns based on time use and social contact data and their implications for infectious disease transmission models. BMC Infectious Diseases. 2022;22(1):954. doi:10.1186/s12879-022-07917-y. [PMC free article] [PubMed] [CrossRef] [Google Scholar]
  • Hufnagel L., Brockmann D., Geisel T. Forecast and control of epidemics in a globalized world. Proceedings of the National Academy of Sciences. 2004;101(42):15124–15129. doi:10.1073/pnas.0308344101. [PMC free article] [PubMed] [CrossRef] [Google Scholar]
  • Hurford A., Cownden D., Day T. Next-generation tools for evolutionary invasion analyses. Journal of The Royal Society Interface. 2010;7:561–571. doi:10.1098/rsif.2009.0448. [PMC free article] [PubMed] [CrossRef] [Google Scholar]
  • Hyman J.M., Li J. Disease transmission models with biased partnership selection. Applied Numerical Mathematics. 1997;24(2):379–392. doi:10.1016/S0168-9274(97)00034-2. [CrossRef] [Google Scholar]
  • Iacobucci D. Graphs and Matrices in “Social Network Analysis Methods and Applications” 1994:92–166. doi:10.1017/CBO9780511815478.005. [CrossRef] [Google Scholar]
  • Inaba H., Nishiura H. The basic reproduction number of an infectious disease in a stable population: The impact of population growth rate on the eradication threshold. Mathematical Modelling of Natural Phenomena. 2008;3(7):194–228. doi:10.1051/mmnp:2008050. [CrossRef] [Google Scholar]
  • Jacquez J.A., Simon C.P., Koopman J., Sattenspiel L., Perry T. Modeling and analyzing HIV transmission: The effect of contact patterns. Mathematical Biosciences. 1988;92(2):119–199. doi:10.1016/0025-5564(88)90031-4. [CrossRef] [Google Scholar]
  • Jarvis C.I., Van Zandvoort K., Gimma A., Prem K., CMMID COVID-19 working group. Klepac P., Rubin G.J., et al. Quantifying the impact of physical distance measures on the transmission of COVID-19 in the UK. BMC Medicine. 2020;18(1):124. doi:10.1186/s12916-020-01597-8. [PMC free article] [PubMed] [CrossRef] [Google Scholar]
  • van de Kassteele J., van Eijkeren J., Wallinga J. Efficient estimation of age-specific social contact rates between men and women. Annals of Applied Statistics. 2017;11(1):320–339. doi:10.1214/16-AOAS1006. [CrossRef] [Google Scholar]
  • Keeling M.J., Eames K.T.D. Networks and epidemic models. Journal of The Royal Society Interface. 2005;2(4):295–307. doi:10.1098/rsif.2005.0051. [PMC free article] [PubMed] [CrossRef] [Google Scholar]
  • Keeling M.J., Rohani P. 2008. Host heterogeneities. chap. 3, ISBN 9780691116174. [CrossRef] [Google Scholar]
  • Klepac P., Kissler S., Gog J. Contagion! The BBC four pandemic–the model behind the documentary. Epidemics. 2018;24:49–59. doi:10.1016/j.epidem.2018.03.003. [PubMed] [CrossRef] [Google Scholar]
  • Klepac P., Pomeroy L.W., Bjørnstad O.N., Kuiken T., Osterhaus A.D.M.E., Rijks J.M. Stage-structured transmission of phocine distemper virus in the Dutch 2002 outbreak. Proceedings of the Royal Society B: Biological Sciences. 2009;276(1666):2469–2476. doi:10.1098/rspb.2009.0175. [PMC free article] [PubMed] [CrossRef] [Google Scholar]
  • Koltai J., Vásárhelyi O., Röst G., Karsai M. Reconstructing social mixing patterns via weighted contact matrices from online and representative surveys. Scientific Reports. 2022;12(4690) doi:10.1038/s41598-022-07488-7. [PMC free article] [PubMed] [CrossRef] [Google Scholar]
  • Kraemer M.U.G., Golding N., Bisanzio D., Bhatt S., Pigott D.M., Ray S.E., et al. Utilizing general human movement models to predict the spread of emerging infectious diseases in resource poor settings. Scientific Reports. 2019;9(1):5151. doi:10.1038/s41598-019-41192-3. [PMC free article] [PubMed] [CrossRef] [Google Scholar]
  • Kretzschmar M., Morris M. Measures of concurrency in networks and the spread of infectious disease. Mathematical Biosciences. 1996;133:165–195. doi:10.1016/0025-5564(95)00093-3. [PubMed] [CrossRef] [Google Scholar]
  • Leung W.T.M., Meeyai A., Holt H.R., Khieu B., Chhay T., Seng S., et al. Social contact patterns relevant for infectious disease transmission in Cambodia. Scientific Reports. 2023;13(1):5542. doi:10.1038/s41598-023-31485-z. [PMC free article] [PubMed] [CrossRef] [Google Scholar]
  • Liu C.Y., Berlin J., Kiti M.C., Del Fava E., Grow A., Zagheni E., et al. Rapid review of social contact patterns during the COVID-19 pandemic. Epidemiology. 2021;32(6):781–791. doi:10.1097/EDE.0000000000001412. [PMC free article] [PubMed] [CrossRef] [Google Scholar]
  • Machens A., Gesualdo F., Rizzo C., Tozzi A.E., Barrat A., Cattuto C. An infectious disease model on empirical networks of human contact: Bridging the gap between dynamic network data and contact matrices. BMC Infectious Diseases. 2013;13(1):185. doi:10.1186/1471-2334-13-185. [PMC free article] [PubMed] [CrossRef] [Google Scholar]
  • McCreesh N., Mohlamonyane M., Edwards A., Olivier S., Dikgale K., Dayi N., et al. Improving estimates of social contact patterns for airborne transmission of respiratory pathogens. Emerging Infectious Diseases. 2022;28(10):2016–2026. doi:10.3201/eid2810.212567. [PMC free article] [PubMed] [CrossRef] [Google Scholar]
  • Meyer S., Held L. Power-law models for infectious disease spread. Annals of Applied Statistics. 2014;8:1612–1639. doi:10.1214/14-AOAS743. [CrossRef] [Google Scholar]
  • Meyer S., Held L. Incorporating social contact data in spatio-temporal models for infectious disease spread. Biostatistics. 2016;18(2):338–351. doi:10.1093/biostatistics/kxw051. [PMC free article] [PubMed] [CrossRef] [Google Scholar]
  • Mistry D., Litvinova M., Pastore y Piontti A., Chinazzi M., Fumanelli L., Gomes M.F.C., et al. Inferring high-resolution human mixing patterns for disease modeling. Nature Communications. 2021;12(1):323. doi:10.1038/s41467-020-20544-y. [PMC free article] [PubMed] [CrossRef] [Google Scholar]
  • Mossong J., Hens N., Jit M., Beutels P., Auranen K., Mikolajczyk R., et al. Social contacts and mixing patterns relevant to the spread of infectious diseases. PLoS Medicine. 2008;5(3):1. doi:10.1371/journal.pmed.0050074. [PMC free article] [PubMed] [CrossRef] [Google Scholar]
  • Mousa A., Winskill P., Watson O.J., Ratmann O., Monod M., Ajelli M., et al. Social contact patterns and implications for infectious disease transmission – a systematic review and meta-analysis of contact surveys. Elife. 2021;10 doi:10.7554/eLife.70294. [PMC free article] [PubMed] [CrossRef] [Google Scholar]
  • Newman M.E.J. Mixing patterns in networks. Physical Review A. 2003;67 doi:10.1103/PhysRevE.67.026126. [PubMed] [CrossRef] [Google Scholar]
  • Nold A. Heterogeneity in disease-transmission modeling. Mathematical Biosciences. 1980;52(3):227–240. doi:10.1016/0025-5564(80)90069-3. [CrossRef] [Google Scholar]
  • Oh H.S., Jeong S.Y., Yang Y. Apilot study investigating the social contact patterns of Korean elderly. Public Health Nursing. 2021;38(5):926–930. doi:10.1111/phn.12884. [PubMed] [CrossRef] [Google Scholar]
  • Prem K., Cook A.R., Jit M. Projecting social contact matrices in 152 countries using contact surveys and demographic data. PLoS Computational Biology. 2017;13(9):1–21. doi:10.1371/journal.pcbi.1005697. [PMC free article] [PubMed] [CrossRef] [Google Scholar]
  • Read J.M., Lessler J., Riley S., Wang S., Tan L.J., Kwok K.O., et al. Social mixing patterns in rural and urban areas of southern China. Proceedings of the Royal Society B. 1785;281 doi:10.1098/rspb.2014.0268. [PMC free article] [PubMed] [CrossRef] [Google Scholar]
  • Santermans E., Van Kerckhove K., Azmon A., John Edmunds W., Beutels P., Faes C., et al. Structural differences in mixing behavior informing the role of asymptomatic infection and testing symptom heritability. Mathematical Biosciences. 2017;285:43–54. doi:10.1016/j.mbs.2016.12.004. [PubMed] [CrossRef] [Google Scholar]
  • Sattenspiel L. Spread and maintenance of a disease in a structured population. American Journal of Physical Anthropology. 1988;77(4):497–504. doi:10.1002/ajpa.1330770410. [PubMed] [CrossRef] [Google Scholar]
  • Sattenspiel L. 1996. Spatial heterogeneity and the spread of infectious diseases. [CrossRef] [Google Scholar]
  • Sattenspiel L., Simon C.P. The spread and persistence of infectious diseases in structured populations. Mathematical Biosciences. 1988;90(1):341–366. doi:10.1016/0025-5564(88)90074-0. [CrossRef] [Google Scholar]
  • Schenzle D. An age-structured model of pre- and post-vaccination measles transmission. Mathematical Medicine and Biology: A Journal of the IMA. 1984;1(2):169–191. doi:10.1093/imammb/1.2.169. [PubMed] [CrossRef] [Google Scholar]
  • Simini F., González M.C., Maritan A., Barabási A.-L. Auniversal model for mobility and migration patterns. Nature. 2012;484(7392):96–100. doi:10.1038/nature10856. [PubMed] [CrossRef] [Google Scholar]
  • Smith M.K., Graham M., Latkin C.A., Go V.L. Using contact patterns to inform HIV interventions in persons who inject drugs in northern vietnam. JAIDS Journal of Acquired Immune Deficiency Syndromes. 2018;78(1):1–8. doi:10.1097/QAI.0000000000001632. [PMC free article] [PubMed] [CrossRef] [Google Scholar]
  • Stehlé J., Voirin N., Barrat A., Cattuto C., Isella L., Pinton J.-F., et al. High-Resolution measurements of face-to-face contact patterns in a primary school. PLoS One. 2011;6(8):1–13. doi:10.1371/journal.pone.0023176. [PMC free article] [PubMed] [CrossRef] [Google Scholar]
  • Thindwa D., Jambo K.C., Ojal J., MacPherson P., Dennis Phiri M., Pinsent A., et al. Social mixing patterns relevant to infectious diseases spread by close contact in urban Blantyre, Malawi. Epidemics. 2022;40 doi:10.1016/j.epidem.2022.100590. [PMC free article] [PubMed] [CrossRef] [Google Scholar]
  • Tizzoni M., Bajardi P., Decuyper A., Kon Kam King G., Schneider C.M., Blondel V., et al. On the use of human mobility proxies for modeling epidemics. PLoS Computational Biology. 2014;10(7):1–15. doi:10.1371/journal.pcbi.1003716. [PMC free article] [PubMed] [CrossRef] [Google Scholar]
  • Vynnycky E., White R. 2010. Sexually transmitted infections. chap. 8, ISBN 9780198565765. [Google Scholar]
  • Walker J., Aylett-Bullock J., Shi D., Maina A.G.K., Evers E.S., Harlass S., Krauss F. Amixed-method approach to determining contact matrices in the cox’s bazar refugee settlement [pre-print] medRxiv. 2022 doi:10.1101/2022.11.22.22282641. [PMC free article] [PubMed] [CrossRef] [Google Scholar]
  • Wallinga J., Teunis P., Kretzschmar M. Using data on social contacts to estimate age-specific transmission parameters for respiratory-spread infectious agents. American Journal of Epidemiology. 2006;164(10) doi:10.1093/aje/kwj317. [PubMed] [CrossRef] [Google Scholar]
  • Wallinga J., van de Kassteele J., Hens N. Handbook of infectious disease data analysis. Chapman and Hall/CRC; 2019. Contact patterns for contagious diseases; pp. 93–110. [CrossRef] [Google Scholar]
  • Willem L., Hoang V.T., Funk S., Coletti P., Beutels P., Hens N. SOCRATES: An online tool leveraging a social contact data sharing initiative to assess mitigation strategies for COVID-19. BMC Research Notes. 2020;13(1):293. doi:10.1186/s13104-020-05136-9. [PMC free article] [PubMed] [CrossRef] [Google Scholar]
  • Williams J.R., Anderson R.M. Journal of the Royal Statistical Society: Series A. 1994;157(1):69–87. doi:10.2307/2983506. [PubMed] [CrossRef] [Google Scholar]
  • Xia Y., Bjørnstad O.N., Grenfell B.T. Measles metapopulation dynamics: A gravity model for epidemiological coupling and dynamics. The American Naturalist. 2004;164(2):267–281. doi:10.1086/422341. [PubMed] [CrossRef] [Google Scholar]
  • Yamamoto N., Ejima K., Nishiura H. Modelling the impact of correlations between condom use and sexual contact pattern on the dynamics of sexually transmitted infections. Theoretical Biology and Medical Modelling. 2018;15(6) doi:10.1186/s12976-018-0078-9. [PMC free article] [PubMed] [CrossRef] [Google Scholar]
  • Zagheni E., Billari F.C., Manfredi P., Melegaro A., Mossong J., Edmunds W.J. Using time-use data to parameterize models for the spread of close-contact infectious diseases. American Journal of Epidemiology. 2008;168(9):1082–1090. doi:10.1093/aje/kwn220. [PubMed] [CrossRef] [Google Scholar]
  • van Leeuwen E., PHE Joint Modelling Group, Sandmann F. Augmenting contact matrices with time-use data for fine-grained intervention modelling of disease dynamics: A modelling analysis. Statistical Methods in Medical Research. 2022;31(9):1704–1715. doi:10.1177/09622802211037078. [PubMed] [CrossRef] [Google Scholar]
  • van Zandvoort K., Bobe M.O., Hassan A.I., Abdi M.I., Ahmed M.S., Soleman S.M., et al. Social contacts and other risk factors for respiratory infections among internally displaced people in Somaliland. Epidemics. 2022;41 doi:10.1016/j.epidem.2022.100625. [PubMed] [CrossRef] [Google Scholar]

Articles from Infectious Disease Modelling are provided here courtesy of KeAi Publishing

Transmission matrices used in epidemiologic modelling (2024)
Top Articles
Commercial Vehicle Insurance: Taxi & Passenger OLA/UBER Cab Insurance Online
EliteMarketMovers : FOREX - BTCUSD FORECAST TODAY
Splunk Stats Count By Hour
Pga Scores Cbs
T Mobile Rival Crossword Clue
Senior Tax Analyst Vs Master Tax Advisor
Craigslist Free Stuff Appleton Wisconsin
Mawal Gameroom Download
41 annonces BMW Z3 occasion - ParuVendu.fr
Rubfinder
Lost Pizza Nutrition
Dusk
Craigslist Alabama Montgomery
Evil Dead Rise Showtimes Near Regal Columbiana Grande
What is Cyber Big Game Hunting? - CrowdStrike
The Banshees Of Inisherin Showtimes Near Regal Thornton Place
Dr Manish Patel Mooresville Nc
Mile Split Fl
Nashville Predators Wiki
Canvas Nthurston
Roll Out Gutter Extensions Lowe's
Unterwegs im autonomen Freightliner Cascadia: Finger weg, jetzt fahre ich!
Jeff Now Phone Number
Craigslist Maryland Trucks - By Owner
Ou Class Nav
Horn Rank
SOGo Groupware - Rechenzentrum Universität Osnabrück
Is Light Raid Hard
Giantbodybuilder.com
Combies Overlijden no. 02, Stempels: 2 teksten + 1 tag/label & Stansen: 3 tags/labels.
1964 Impala For Sale Craigslist
Trust/Family Bank Contingency Plan
O'reilly Auto Parts Ozark Distribution Center Stockton Photos
Adecco Check Stubs
The Ride | Rotten Tomatoes
Greater Keene Men's Softball
Weather Underground Cedar Rapids
Mississippi weather man flees studio during tornado - video
John M. Oakey & Son Funeral Home And Crematory Obituaries
How To Get To Ultra Space Pixelmon
Cvs Coit And Alpha
Meet Robert Oppenheimer, the destroyer of worlds
Fine Taladorian Cheese Platter
Online TikTok Voice Generator | Accurate & Realistic
Craigslist Com Brooklyn
Billings City Landfill Hours
Costco Tire Promo Code Michelin 2022
Lagrone Funeral Chapel & Crematory Obituaries
32 Easy Recipes That Start with Frozen Berries
Latest Posts
Article information

Author: Tyson Zemlak

Last Updated:

Views: 6660

Rating: 4.2 / 5 (63 voted)

Reviews: 94% of readers found this page helpful

Author information

Name: Tyson Zemlak

Birthday: 1992-03-17

Address: Apt. 662 96191 Quigley Dam, Kubview, MA 42013

Phone: +441678032891

Job: Community-Services Orchestrator

Hobby: Coffee roasting, Calligraphy, Metalworking, Fashion, Vehicle restoration, Shopping, Photography

Introduction: My name is Tyson Zemlak, I am a excited, light, sparkling, super, open, fair, magnificent person who loves writing and wants to share my knowledge and understanding with you.