1 Introduction

Capture–recapture (CRC) analysis is widely used to estimate the size of hidden human populations, who may stay under the radar because of societal stigma or lack of exhaustive recording. Estimating their size can help governments make informed funding/policy decisions and be useful for sociological or medical research. Examples include estimating the number of people who inject drugs (PWID) (e.g., [16, 23, 47]), female sex-workers (FSW) (e.g., [23, 42, 46]), men who have sex with men (MSM) (e.g., [23, 46]), the homeless (e.g., [19, 20]) and cases during the recent Covid-19 pandemic (e.g., [13, 50]).

CRC estimates population size from the overlap of sample units between two or more capture occasions (repeated samples of the target population). The presence of a sample unit at a capture occasion is called a ‘capture’, so if present at three capture occasions we would say it is ‘captured’ three times. The often-stigmatised nature of hidden human populations makes it hard to obtain capture occasions via direct sampling, so various indirect approaches have been devised. For example, a register from an external organisation (such as a subscriber list) is often used as a capture occasion (e.g., [19, 47]).

Exploration continues into ways to source CRC data of human populations. The current paper considers a novel pairing of CRC data derived from self-reported social networks of participants and a CRC analytical technique called ‘zero-truncated count distribution modelling’. In the current paper’s approach, a sample group of participants from the target population each submit a list of individuals they know from that population. Each of these self-reported social networks is treated as a distinct capture occasion. Because participants cannot be named in their own list, we also consider if being a participant can be treated as an extra capture occasion.

Treating each participant’s self-reported social network as a distinct capture occasion means many capture occasions can be obtained from a single sample of participants. Pairing this with the zero-truncated modelling approach makes use of this strength, as it is particularly suited for analysing overlap between many capture occasions.

1.1 Advantages of using three or more capture occasions

An advantage of using three or more capture occasions is that otherwise CRC is heavily reliant on the assumed independence between capture occasions. Positive dependence between capture occasions is known to lead to underestimates of population size and vice versa, which can be severe when only using two capture occasions because estimation is limited to the Lincoln–Petersen estimator or equivalent [41]. Sensitivity to the independence assumption increases sensitivity to other CRC assumptions that play into it. For example, false matches of individuals between capture occasions can cause negative dependence whereas unequal catchability of individuals can cause positive dependence because some individuals are more likely to be captured multiple times.

Obtaining three or more capture occasions enables more nuanced methods of estimating population size with less sensitivity to the independence assumption. Examples include: log linear modelling, which can consider dependencies between three or more capture occasions (e.g., [3, 19, 20, 47, 53]); Bayesian latent class modelling (e.g., [23, 42, 44]); the ‘ratio plot’ diagnostic tool in zero-truncated modelling [5, 9]; and continuous time CRC modelling that, when using many capture occasions over time, can allow for a delayed onset of behaviours relating to being repeatedly captured [25].

There is hence great interest in exploring ways to source three or more capture occasions. Examples include: using a single external register of repeated entries, such as a hospital admissions register (e.g., [8, 35, 54]); selecting a sample of participants from the target population as a third capture occasion in combination with two external registers (e.g., [47, 53]); and a 3-capture-occasion adaptation of the Unique-object Multiplier method (e.g., [23, 42, 44]).

1.2 Utilising social networks in capture–recapture

Several of the above approaches involve selecting a (theoretically) representative sample from the target population as a capture occasion. In the absence of a sample frame, these approaches often utilise the social ties between members of the target population to perform respondent-driven sampling (RDS), which is thought to approach a somewhat representative sample [34].

While it is hence not uncommon for participants’ social networks to be used in CRC to expand the reach of sampling, their use as capture occasions is less explored. However, interest in this area is growing. One approach by Dombrowski et al. [22] derives two capture occasions, where the first capture is being a participant and the second is being named in at least one other participant’s self-reported social network. In a more recent example in Buchanan et al. [16], the first capture was being a participant and the second was being named in at least one other participant’s self-reported social network who did not appear in one’s own. Recent advances in population size estimation using Privatised Network Sampling have extended the Dombrowski et al. [22] approach with more nuanced population-size estimators [26, 37]. Two of these from Fellows [26], the ‘cross-alter’ and ‘cross-network’ estimators, considered the overlap of non-participants and participants between participants’ self-reported social networks, which was found to produce better estimates due to increasing the number of individuals in the data [26]. The latter was a promising sign that participants’ self-reported social networks could function as a viable source of captures of participants and non-participants. Besides CRC, other methods that use social-network data from a target population to estimate its size include the Snowball method [20, 27], the Network Scale-up method (e.g., [24]), Successive Sampling [32, 33] and CRC Successive Sampling [38].

While methods like Successive Sampling ask participants how many individuals they know from the target population, the current paper’s approach instead asks them to list individuals. This increases data sensitivity, particularly as CRC is often used on stigmatised populations. However, a number of studies have demonstrated ways for participants to submit anonymised or pseudo-anonymised lists of social ties, showing it can be practicable [16, 22, 26, 37].

We initially present a case study, using real-world data, that attempts to estimate a known target population size. Methodology is outlined in Sect. 2, initial inspection of data is in Sect. 3.1, model fitting is in Sect. 3.2 and estimation of population size is in Sect. 3.3. A follow-up simulation study is described in Sect. 4 and further discussion is in Sect. 5.

2 Methods

2.1 Zero-truncated count distribution modelling

‘Zero-truncated count distribution modelling’ has long been used to estimate population size from CRC data. See Böhning et al. [11] for an introduction on its application to human populations. The method uses aggregate-level data of how many sample units are captured exactly 1 time (\(f_1\)), exactly 2 times (\(f_2\)), exactly 3 times (\(f_3\)), etc. across several capture occasions. This is summarised into a frequency distribution of discrete counts (Table 1). The number of sample units captured at least once is referred to as the observed population (n). The number of uncaptured sample units is \(f_0\). Hence, the total size of the target population is \(N=f_0+n\).

Table 1 Underlying structure of CRC data when summarised into a frequency distribution

To estimate the size (N) of the target population, the method assumes that the observed data (\(f_1\) to \(f_m\)) follow the same shape/profile as \(f_0\) to \(f_m\). Hence, it assumes that the shape of \(f_1\) to \(f_m\) can be projected further to the left to estimate \(f_0\). To approximate the shape of \(f_1\) to \(f_m\), a zero-truncated (ZT) probability distribution is used (e.g., ZT Poisson). Based on estimated parameters from that ZT distribution, \(f_0\) (and subsequently N) is estimated as though \(f_0\) to \(f_m\) follow the untruncated version of that ZT distribution.

Several ZT distributions can be considered to see which is most consistent with the data. For example, in Fig. 1, the data is closer to ZT distribution B than ZT distribution A, suggesting estimates of N should be based on B rather than A. While a variety of distributions can theoretically be considered, the current literature tends to focus on a small number of distributions stemming from the ZT Poisson and ZT geometric, described further in Sect. 2.2.

The approach tends to involve choosing one or two ZT distributions as starting points and then using model-fitting diagnostic tests, like the chi-square goodness of fit (GoF) test, to check for inconsistency with the observed data. A conceptual drawback is that, by testing for inconsistency, ZT distributions can be deemed consistent with the data via acceptance of the null hypothesis. For this reason, the use of the GoF test here only tends to provide partial evidence of consistency between a ZT distribution and the observed data rather than a more conclusive finding. However, the ‘ratio plot’ diagnostic tool (described in Sect. 2.3) can help inform the results from these tests by providing a more granular inspection of how consistent the ZT distributions are to the data.

Fig. 1
figure 1

Graphical display demonstrating zero-truncated modelling approach

As more generally with CRC analysis, this rests on some key assumptions: firstly, that capture occasions are independent; secondly, that all members of the target population (sample units) are equally catchable (i.e., have a homogeneous probability of being captured) at any capture occasion; thirdly, that individuals are accurately matched between capture occasions; and fourthly, that the target population is closed/static across all capture occasions.

2.2 Zero-truncated distributions considered

As \(f_1\) to \(f_m\) are non-negative integers, the ZT Poisson tends to be a starting point for modelling them. However, with ZT Poisson the mean and variance are assumed equal, which is often not true because heterogeneity (unequal capture probabilities among sample units) can cause over-or-underdispersion of variance. Hence, other commonly used distributions include the ZT geometric and various mixing distributions of the ZT Poisson, such as the Poisson-gamma and the Conway–Maxwell Poisson (CMP). In the current paper, the ZT Poisson and ZT geometric were used as starting points, as their popularity has led to the development of population-size estimators with robustness to heterogeneity. Also considered were the zero-truncated one-inflated (ZTOI) Poisson and ZTOI geometric, as they can factor in an overly large \(f_1\).

The untruncated Poisson has a probability mass function (pmf) of: \(p_x=\exp {(-\theta )}\theta ^x/x!\) for \(x=0,1,\dots ,m\) where \(\hat{\theta }_{MLE}=\bar{x}\). The ZT Poisson has pmf: \(p_x=\theta ^x/((\exp {(\theta ) }-1) x!)\) for \(x=1,2,\dots ,m\). Although \(\hat{\theta }_{MLE}\) cannot be given in closed form for the ZT Poisson, it can be derived from the untruncated Poisson using the E-M algorithm [11, 21]. This oscillates between two steps, using an arbitrary initial value (e.g., 0.5) for \(\hat{\theta }\). In Step 1, \(\hat{f_0}=n\times \exp {(-\hat{\theta })}/(1-\exp {(-\hat{\theta })})\) and, in Step 2, \(\hat{\theta }=S/(n+\hat{f_0})\) where \(S=\sum _{x=1}^m xf_x\) and \(n=\sum _{x=1}^m f_x\). The ZTOI Poisson has pmf:

$$\begin{aligned} p_{x}^{1+}= \begin{Bmatrix} w+(1-w)\frac{\lambda }{\exp {(\lambda )}-1} \dots \text { if } x=1 \\ (1-w)\frac{\lambda ^{x}}{(\exp {(\lambda )}-1)x!} \dots \text { if } x>1 \end{Bmatrix} \end{aligned}$$

for \(x=1,2,\dots ,m\) where w is a weight parameter. For the ZTOI Poisson, \(\hat{\lambda }_{MLE}\) and \(\hat{w}_{MLE}\) cannot be given in closed form and were hence iteratively calculated using an E-M algorithm approach from Godwin and Böhning [28]. This cycled through the following steps. In step 1, we assigned arbitrary initial values (e.g., 0.5) for \(\hat{N}\) and \(\hat{\delta _1}\). (The \(\hat{\delta _1}\) is the number of unobservable inflated 1s.) In step 2, we estimated \(\hat{w}\) via \(\hat{w}=\hat{\delta _1}/n\). In step 3, we estimated \(\hat{\lambda }\) via \(\hat{\lambda }=(\sum _{i=1}^{\hat{N}}{x_i} - \hat{\delta _1}) / (\hat{N}(1-\hat{w}))\) where \(\sum _{i=1}^{\hat{N}}x_i \equiv \sum _{x=1}^m xf_x\). In step 4, we estimated \(\hat{\delta _1}\) via \(\hat{\delta _1}=f_1\times \hat{w}(1-\exp {(-\hat{\lambda })}) / \big (\hat{w}(1-\exp {(-\hat{\lambda })}) + (1-\hat{w})\hat{\lambda }\exp {(-\hat{\lambda })}\big )\). In step 5, we repeated steps 2-4 until \(\hat{\delta _1}\) converged. Then, in step 6, we re-estimated \(\hat{N}\) via \(\hat{N}=n/(1-\exp {(-\hat{\lambda })})\). We repeated steps 2-6 until \(\hat{N}\) converged, at which point \(\hat{\lambda }_{MLE}\) and \(\hat{w}_{MLE}\) would also have converged.

The untruncated geometric has pmf: \(p_x=(1-\theta )^x \theta \) for \(x=0,1,\dots ,m\) where \(\hat{\theta }_{MLE}=1/(\bar{x}+1)\). The ZT geometric has pmf: \(p_x=(1-\theta )^{(x-1)}\theta \) for \(x=1,2,\dots ,m\) where \(\hat{\theta }_{MLE}=1/\bar{x}=1/(S/n)\). The ZTOI geometric has pmf:

$$\begin{aligned} p_{x}^{1+}= \begin{Bmatrix} w(1-\theta )^x\theta /(1 - w\theta ) \dots \text { if } x>1 \\ \big ((1-w)+w(1-\theta )^x\theta \big )/(1-w\theta ) \dots \text { if } x=1 \end{Bmatrix} \end{aligned}$$

for \(x=1,2,\dots ,m\) where w is a weight parameter; \(0\le w\le 1\). For the ZTOI geometric, \(\hat{\theta }_{MLE}\) and \(\hat{w}_{MLE}\) cannot be given in closed form and were hence calculated via the nested E-M algorithm approach from Kaskasamkul and Böhning [36]. This oscillates between the following two steps, using arbitrary initial values (e.g., 0.5) for \(\hat{\theta }\) and \(\hat{w}\). In Step 1, \(\hat{f}_0=n \times \hat{w} \times \hat{\theta }/(1-\hat{w} \times \hat{\theta })\) and \(\hat{N}=n+\hat{f}_0\). In Step 2, \(\hat{w}=1-(f_1/\hat{N})(1-\hat{w})/\big ((1-\hat{w})+\hat{w}(1-\hat{\theta })\hat{\theta }\big )\) and

$$\begin{aligned} \hat{\theta }= \frac{\hat{N}-f_1 (1-\hat{w})/((1-\hat{w})+\hat{w}(1-\hat{\theta })\hat{\theta })}{\hat{N}+\sum _{i=1}^{\hat{N}}x_i-2f_1 (1-\hat{w})/\big ((1-\hat{w})+\hat{w}(1-\hat{\theta })\hat{\theta }\big )} \end{aligned}$$

where \(\sum _{i=1}^{\hat{N}}x_i \equiv \sum _{x=1}^m xf_x\).

2.3 The ratio plot

The ‘ratio plot’ diagnostic tool [9] was used to help inform model-fitting diagnostics. This is a graph in which horizontality of data-points indicates the level of consistency between a given ZT distribution and observed data.

The tool is based on the power series density \(p_x(\theta )=a_x\theta ^x/\sum _{x=0}^{\infty }\{a_x\theta ^x\}\) where \(\sum _{x=0}^{\infty }\{a_x\theta ^x\}\) is a normalising constant that converts \(p_x(\theta )\) into proportions that sum to 1. If \(a_x\) is set to \(a_x=1/x!\), the power series density becomes the Poisson density: \(p_x(\theta )=(\theta ^{x}/x!)/\sum _{x=0}^{\infty }\{\theta ^{x}/x!\}\). If \(a_x\) is set to \(a_x=1\), the power series density becomes the geometric density: \(p_x(\theta ^{\star })=\theta ^{\star x}/\sum _{x=0}^{\infty }\{\theta ^{\star x}\}\) where \(\theta ^{\star }=1-\theta \). Although outside the scope of the current paper, it can also model a binomial density by setting \(a_x=\big ({\begin{matrix} T\\ x \end{matrix}}\big )\) for \(x=0,\dots T\) where T are positive integers and where \(a_x=0\) when \(x>T\) [5].

A property of the power series density is that the ratio of \(p_{x+1}(\theta )a_x\) to \(p_x(\theta )a_{x+1}\) is always equal to \(\theta \), which is a constant. That is, \(r_x=p_{x+1}(\theta )a_x/(p_x(\theta )a_{x+1})=\theta \). This means \(r_x\) should also be a constant. Hence, to check for consistency between a set of observed data (\(f_1\) to \(f_m\)) and a specific power series density (e.g., Poisson), the observed data can be combined with \(a_x\) to produce \(r_x\) and the level of constancy in \(r_x\) can be inspected.

Although \(p_x(\theta )\) is unknown, \(f_x/N\) can be used instead as a non-parametric estimate; the quantity of N is unknown but cancels out in the ratio. Hence, \(r_x\) can be estimated as \(\hat{r}_x=(a_x/a_{x+1})\times (f_{x+1}/f_x)\) where \(f_{x+1}/f_x\) is the ratio of each adjacent pair of counts in the observed data and \(a_x/a_{x+1}\) is the inverse of their respective coefficients from the power series distribution. In the Poisson case, \(a_x/a_{x+1}=x+1\) and hence \(\hat{r}_x=(x+1)f_{x+1}/f_x\). In the geometric case, \(a_x/a_{x+1}=1\) and hence \(\hat{r}_x=f_{x+1}/f_x\).

If \(\hat{r}_x\) is a constant then a horizontal series of data-points should occur when plotting \(\log {(\hat{r}_x)}\) against x. From this, the consistency between a given ZT distribution and the observed data can be visually appraised by inspecting the gradient of a linear regression line plotted between x and \(\log {(\hat{r}_x)}\). A more horizontal line provides some evidence that the ZT distribution under consideration is consistent with the observed data [9]. Care needs to be taken, however, as heterogeneity (unequal catchability among sample units) can also contribute to causing a slope in the linear regression line.

Rocchetti et al. [49] advise using weighted linear regression to reduce the impact of heteroskedasticity in the ratio plot, using weights (\(W_i\)) derived from diagonal components of \((cov(Y))^{-1}\):

$$\begin{aligned} W=\begin{bmatrix} \frac{1}{f_1}+\frac{1}{f_2} &{}\quad \dots &{}\quad \dots \\ \dots &{}\quad \frac{1}{f_i}+\frac{1}{f_{i+1}} &{}\quad \dots \\ \dots &{}\quad \dots &{}\quad \frac{1}{f_{m-1}}+\frac{1}{f_m} \end{bmatrix}^{-1} = \begin{bmatrix} f_1+f_2 &{}\quad \dots &{}\quad \dots \\ \dots &{}\quad f_i+f_{i+1} &{}\quad \dots \\ \dots &{}\quad \dots &{}\quad f_{m-1}+f_m \end{bmatrix} \end{aligned}$$

These weights (\(W_i=f_i+f_{i+1}\)) are the same for both the ZT Poisson and ZT geometric. For example, to calculate weighted regression in the Poisson case, we take \(\hat{\beta }=(X^{T} WX)^{-1} X^{T} WY\) where

$$\begin{aligned} Y=\begin{pmatrix} \log {(2f_2/f_1)} \\ \log {(3f_3/f_2)} \\ \vdots \\ \log {(mf_m/f_{m-1})} \end{pmatrix}, X=\begin{pmatrix} 1 &{}\quad &{}\quad &{}\quad 1 \\ 1 &{}\quad &{}\quad &{}\quad 2 \\ \vdots &{}\quad &{}\quad &{}\quad \vdots \\ 1 &{}\quad &{}\quad &{}\quad m-1 \end{pmatrix}, W=\begin{pmatrix} f_1+f_2 \\ f_2+f_3 \\ \vdots \\ f_{m-1}+f_m \end{pmatrix} \end{aligned}$$

A useful property of the power series density (and hence the ratio plot) is that it is the same when untruncated or zero-truncated. This brings an added utility to ratio plots, as the intercept of the weighted linear regression line can extrapolate the ratio plot to where \(x=0\). From this can be derived the Weighted Least Squares (WLS) estimator of population size [5].

2.4 The ratio plot under the null

Because the ratio plot (\(\hat{r}_x\)) is the same for both the untruncated and zero-truncated versions of a given distribution, a 95% confidence interval can be displayed pertaining to the null hypothesis that \(f_1\) to \(f_m\) are consistent with the untruncated version of the ZT distribution being tested. This ‘ratio plot under the null’ is demonstrated in Fig. 2 using dummy data. See the accompanying supplementary material (Online Resources 3–4) for examples in R code.

Fig. 2
figure 2

Graphs demonstrating concept of ratio plot under the null. Left panel: demonstration of high consistency. Central panel: demonstration of worse consistency. Right panel: demonstration of overall consistency apart from one ratio

While the null hypothesis holds in both the left and central panels of Fig. 2, evidence of consistency is stronger in the left panel because the weighted regression line is nearly horizontal and all ratios are well within the null hypothesis region. We might hence choose the ZT distribution that was being tested in the left panel over that of the centre panel. In the right panel, the weighted regression line is nearly horizontal but the ratio of \(f_2\) to \(f_3\) is outside the null hypothesis region. The latter gives some evidence of overall consistency apart from \(f_2\) and/or \(f_3\), prompting consideration of other ZT distributions.

Null hypothesis regions for the ratio plots were calculated using methodology from Böhning and Punyapornwithaya [7]. In both the Poisson and geometric cases, these were calculated as \(\log {\hat{r}_x}\pm 1.96\times \sqrt{var{(\log {\hat{r}_x})}}\).

The null hypothesis in the geometric case is that \(f_1\) to \(f_m\) are consistent with the untruncated geometric distribution (i.e., consistent with the power series density when \(a_x=1\)). Setting \(a_x=1\) makes the power series density become \(p_x(\theta ^{\star })\) from the geometric density, where \(\theta ^{\star }=1-\theta \). Because we are using the log scale in our ratio plot, the null hypothesis region is calculated around \(\log {(1-\hat{\theta })}\). Using an approximated variance term for \(var{(\log {\hat{r}_x})}\), the null hypothesis region in the geometric case is hence

$$\begin{aligned} \log {(1-\hat{\theta })}\pm 1.96\times \sqrt{ \frac{1}{n(1-\hat{\theta })^{x+1}\hat{\theta }} + \frac{1}{n(1-\hat{\theta })^{x}\hat{\theta }} } \end{aligned}$$

where \(\hat{\theta }\) is \(\hat{\theta }_{MLE}\) from the ZT geometric. This is given in closed form as: \(\hat{\theta }_{MLE}=1/\bar{x}=1/(S/n)\) where \(S=\sum _{x=1}^m xf_x\) and \(n=\sum _{x=1}^m f_x\).

The null hypothesis in the Poisson case is that \(f_1\) to \(f_m\) are consistent with the untruncated Poisson (i.e., consistent with the power series density when \(a_x=1/x!\)). Because we are using the log scale in our ratio plot, the null hypothesis region is calculated around \(\log {\hat{\theta }}\). Using an approximated variance term for \(var{(\log {\hat{r}_x})}\), the null hypothesis region in the Poisson case is hence

$$\begin{aligned} \log {\hat{\theta }} \pm 1.96 \times \sqrt{\frac{1}{n\times \exp {(-\hat{\theta })} \hat{\theta }^{x+1} / (x+1)!} + \frac{1}{n\times \exp {(-\hat{\theta })} \hat{\theta }^{x} / x!} } \end{aligned}$$

where \(\hat{\theta }\) is \(\hat{\theta }_{MLE}\) from the ZT Poisson. However, as \(\hat{\theta }_{MLE}\) cannot be given in closed form from the ZT Poisson, it is iteratively calculated from the untruncated Poisson using the E-M algorithm (see Sect. 2.2).

2.5 Estimators of population size

Several estimators of population size (N) are available, specific to each ZT distribution (see e.g., [11]). Some give equal weight to the whole of \(f_1\) to \(f_m\), such as the Turing [29] and MLE estimators. Other estimators give more importance to \(f_1\) and \(f_2\), as they are typically larger than \(f_3\) to \(f_m\) and hence less impacted by fluctuation caused by heterogeneity (unequal catchability of sample units). This gives the latter some robustness to heterogeneity. A popular example is the Chao estimator [18]. A bias-corrected Chao (BC Chao) is also available for small populations [10], and both the Chao and BC Chao have been adapted for one-inflated data [12]. The WLS estimator (mentioned at the end of Sect. 2.3) is similarly robust, as it gives more weight to \(f_1\) and \(f_2\) [5]. The current paper hence prioritised using BC Chao and WLS.

The BC Chao estimator is: \(\hat{N} = n+f_1 (f_1-1)/(2f_2+2)\) for ZT Poisson, \(\hat{N} = n+f_1 (f_1-1)/(f_2+1)\) for ZT geometric, \(\hat{N} = n + 2 \times (f_2^3-3f_2^2+2f_2) / (9 \times (f_3+1) \times (f_3+2))\) for ZTOI Poisson and \(\hat{N} = n+(f_2^3-3f_2^2+2f_2) / ((f_3+1) \times (f_3+2))\) for ZTOI geometric. The WLS estimator is \(\hat{N}=n+f_1 \times \exp {(-\hat{\beta }_0)}\) where \(\hat{\beta }_0\) is the intercept of the weighted linear regression line of the ratio plot. This meant the WLS estimator was only available for the ZT Poisson and ZT geometric in the current paper. See Online Resource 1 for worked examples.

2.6 Confidence intervals for population-size estimators

In the case study, 95% confidence intervals for estimators of N were calculated via the imputed bootstrap approach. Methodology was sourced from Anan et al. [1] which had earlier roots in Buckland and Garthwaite [17], Norris and Pollock [43] and Zwane and Van der Heijden [56]. In this approach, the point-estimate of \(\hat{N}\) and \(\hat{f}_0\) are used to estimate probabilities of \(f_0\) to \(f_m\) as \(\hat{p}_i = \begin{Bmatrix} \frac{\hat{f}_{0}}{\hat{N}},&\quad \frac{f_{1}}{\hat{N}},&\quad \frac{f_{2}}{\hat{N}},&\quad \frac{f_{3}}{\hat{N}},&\quad \dots ,&\quad \frac{f_{m}}{\hat{N}} \end{Bmatrix}\). These are entered as the ‘prob’ parameter of the ‘rmultinom’ function in R statistical software [48], with ‘size’ set to \(\hat{N}\) and the ‘n’ parameter set to the desired number of bootstrap samples. Hence, each bootstrap sample is a multinomially distributed random vector in which \(\hat{N}\) is split across \(f_0\) to \(f_m\). The same point-estimate of \(\hat{N}\) is then calculated from each bootstrap sample, producing a bell-curve around the original point-estimate. The 97.5th and 2.5th percentiles become the 95% confidence interval.

Anan et al. [1] advise a bootstrap approach as it avoids needing to assume estimators of N are normally distributed. There would be more reliance on this assumption if instead basing confidence intervals on variance estimators, as these are known to be impacted by skewness of N in CRC [18]. Nevertheless, a slight bias can enter the estimated probabilities (\(\hat{p}_i\)) of \(x_i\) when drawing bootstrap samples if the point-estimate of \(\hat{N}\) is an over-or under-estimate. While Anan et al. [1] suggest only needing \(\hat{N}\) number of bootstrap samples, the use of R software meant thousands could be drawn.

In the follow-up simulation study, confidence intervals were instead the 2.5th and 97.5th percentiles of \(\hat{N}\) across many simulated target populations.

2.7 Case study target population and sampling approach

The target population in the case study was a cohort of 182 students on a university course. As this was an experimental setting, a participation invite was sent across the target population and a sample group of participants was formed by voluntary participation. Invites were sent at the same time, helping ensure the population was closed. Participants were asked to submit their self-reported social network (list of social ties) independently. The non-sensitive nature of the population, and its small size, meant full names were able to function as unique identifiers.

2.8 Case study derivation of capture–recapture data

In the case study, participants’ self-reported social networks were treated as distinct capture occasions (referred to as ‘nomination capture occasions’). For each unique individual (participant or non-participant) named in at least one self-reported social network, we counted how many they were named in (i.e., how many nomination capture occasions they were captured at). Out of K self-reported social networks, non-participants could be named up to K times. However, participants could only be named up to \(K-1\) times because they could not be named in their own social network. As this violated the assumed equal catchability of sample units at all capture occasions, we considered if being a participant could be an extra capture to make the maximum number of captures equal (referred to as the ‘participation capture occasion’).

The structure of how captures were recorded is shown in Table 2. In this demonstration table, persons A, B and D are participants whereas persons C and E are non-participants. Hence, the columns ‘Nominated by C’ and ‘Nominated by E’ contain only 0s. The two right-hand columns show the total number of times each person was captured when either including or excluding the participation capture occasion. Data from either of the right-hand columns could then be summarised into a frequency distribution (\(f_1\) to \(f_m\)) of how many individuals were captured exactly 1 time, exactly 2 times, exactly 3 times, etc. when either including or excluding the participation capture occasion.

Table 2 General structure of how captures were recorded, using dummy data

3 Case study results

3.1 Case study results from data-collection

16 participants took part in the study, each submitting a self-reported social network. As described above, captures were summarised into a frequency distribution (\(f_1\) to \(f_m\)) of the number of individuals captured x number of times. Two versions of the dataset were derived that either included or excluded the participation capture occasion, shown in Tables 3 and 4 respectively.

Table 3 Number of individuals captured x number of times at nomination and participation capture occasions, by whether participant
Table 4 Number of individuals captured x number of times at nomination capture occasions, by whether participant

In Table 3, 61 individuals were captured at least once, of which 29 were captured exactly 1 time, 9 were captured exactly 2 times, 11 exactly 3 times, etc. The number of uncaptured individuals (\(f_0\)) would ordinarily be unknown but in this case was known to be 121 in Table 3 and 122 in Table 4. The inclusion of the participation capture occasion increased the number of times each participant was captured by 1, which meant the ‘Participants’ row was shifted right in Table 3 compared to in Table 4.

Collapsing the data into two binary capture occasions (Table 5) showed there was strong positive dependence between participation and nomination. 15 out of 16 participants were nominated at least once. Positive dependence between capture occasions is known to produce underestimates of population size [14]. Hence, while an argument for including the participation capture occasion was that it would make the maximum number of captures equal between participants and non-participants, a counter-argument was that excluding it could be a way to effectively deflate captures of participants by 1 to help counter-balance the positive dependence that had been found.

Table 5 Nomination and participation captures collapsed into two capture occasions

As sample units are assumed to have equal probabilities of capture, \(f_1\) to \(f_m\) should theoretically have a similar shape/profile across all sub-sections of the observed population. However, the shape/profile was found to substantially differ for participants and non-participants, even when only considering nomination capture occasions (Fig. 3). This suggested that estimates of population size (N) could potentially be more accurate if \(f_1\) to \(f_m\) only included captures of non-participants, as this would have a more homogeneous shape/profile and be easier to fit with a ZT distribution. The number of participants would need to be added on to any such estimates of N afterwards because the exclusion of participants from the CRC data would mean treating participants as outside the target population during the calculation of estimators.

Fig. 3
figure 3

Number of individuals captured x number of times. Left panel: from nomination and participation capture occasions. Right panel: from nomination capture occasions

To summarise, initial inspection of the case study data suggested three possible ways of deriving CRC data (\(f_1\) to \(f_m\)) with which to estimate N. These were: (a) as captures from nomination and participation capture occasions (total row from Table 3); (b) as captures from nomination capture occasions (total row from Table 4); or (c) as captures of just non-participants (from either Table 3 or 4). Model-fitting and population-size estimation was performed on all three versions to see how they compared.

3.2 Case study results from model-fitting

Visual appraisal of observed vs expected values (Fig. 4) suggested that, for all three versions of the observed data (\(f_1\) to \(f_m\)) under consideration, the data was more consistent with the ZT geometric than the ZT Poisson. MLE parameter values (\(\hat{\theta }_{MLE}\)) were used for expected values as per Sect. 2.2.

Fig. 4
figure 4

Observed data (\(f_1\) to \(f_m\)) vs expected values from ZT Poisson and ZT geometric. Left panel: from nomination and participation capture occasions. Centre panel: from nomination capture occasions. Right panel: from captures of just non-participants

Ratio plots under the null are shown in Fig. 5. For all three versions of \(f_1\) to \(f_m\), the weighted regression line was more horizontal in the geometric case. This again gave some evidence that the ZT geometric was more consistent with \(f_1\) to \(f_m\). In the Poisson case, the left-most ratio (that of \(f_1\) to \(f_2\)) fell outside the null hypothesis region, indicating that \(f_1\) and/or \(f_2\) were inconsistent with the untruncated Poisson. There was thus a particular risk of bias entering population-size estimators from the ZT Poisson that prioritise \(f_1\) and/or \(f_2\).

Fig. 5
figure 5

Ratio plots under the null in Poisson and geometric cases. Left panels contain only 4 ratios because \(f_5=0\), which meant \(f_5\) to \(f_8\) were collapsed into \(f_5\) so there was no gap in the ratio plots. As the x-axis starts at 1, the intercept of the weighted regression line is not displayed. Intercepts were: − 0.586 (top left panel); − 0.5102 (top centre); − 1.7566 (top right); − 0.9957 (bottom left); − 0.9747 (bottom centre); and − 2.1106 (bottom right)

To test if the observed data were inconsistent with the ZT Poisson or ZT geometric, the chi-square goodness-of-fit statistic, \(\chi ^2=\sum _{x=1}^{m-1} \{(\log {\hat{r_x}}-\log {\bar{r_x}})^2/\widehat{var}(\log {\hat{r_x}})\}\), was used on the ratio-plot data where \(\widehat{var}(\log {\hat{r_x}})=1/f_{x+1} + 1/f_x\) [5, 9, 49]. The mean ratio (\(\bar{r}_x\)) was used as the expected \(\hat{r}_x\) value which, as described in Sect. 2.3, is expected to be a constant. The mean ratio is \(\bar{r}_x=\sum _{x=1}^{m-1}\{(x+1)f_{x+1}\}/\sum _{x=1}^{m-1}f_x\) in the Poisson case and \(\bar{r}_x=\sum _{x=1}^{m-1}f_{x+1}/\sum _{x=1}^{m-1}f_x\) in the geometric case. A significant p-value (\(<0.05\)) meant rejecting the null hypothesis that \(f_1\) to \(f_m\) were consistent with the ZT distribution being tested. Results are shown in Table 6. There was not enough evidence to reject the null hypothesis in the geometric case, which was partial evidence of consistency with ZT geometric. While a non-significant p-value occurred in the Poisson case when only using nomination capture occasions (\(p=0.13\)), this was outweighed by the earlier finding in the Poisson ratio-plots that the ratio of \(f_1\) to \(f_2\) was outside the null hypothesis region because of the particular importance of fitting \(f_1\) and \(f_2\).

Table 6 Chi-square goodness of fit test results

For all three versions of \(f_1\) to \(f_m\), a sizeable jump between \(f_2\) and \(f_1\) suggested possible consistency with a one-inflated distribution. We therefore considered if the ZTOI geometric or ZTOI Poisson offered a closer fit than their ZT counterparts. In each case, the likelihood ratio test (LRT) was used with \(\alpha =0.05\). This test is calculated as \(LRT=-2\times (l_0(0,\tilde{\theta }) - l_A(\hat{w},\hat{\theta }))\) where \(l_0\) and \(l_A\) are the log likelihoods under the null and alternative hypotheses respectively. In the null hypothesis case, \(\tilde{\theta }\) was the \(\hat{\theta }_{MLE}\) from the ZT distribution whereas, in the alternative hypothesis case, \(\hat{\theta }\) and \(\hat{w}\) were the \(\hat{\theta }_{MLE}\) and \(\hat{w}_{MLE}\) from the ZTOI distribution. All MLEs were calculated as per Sect. 2.2.

In the geometric case, log likelihoods were calculated as per methodology in Kaskasamkul and Böhning [36] with \(l_A(\hat{w},\hat{\theta })=f_1\log \{((1-\hat{w})+\hat{w}(1-\hat{\theta })\hat{\theta })/(1-\hat{w}\hat{\theta })\} + \sum _{x=2}^{m}f_x \log \{\hat{w}(1-\hat{\theta })^{x}\hat{\theta }/(1-\hat{w}\hat{\theta })\}\). In the null hypothesis case, \(w=1\) so hence \(l_0(0,\tilde{\theta })=\sum _{x=1}^{m}f_x\log \{(1-\tilde{\theta })^{x}\tilde{\theta }/(1-\tilde{\theta })\}\).

In the Poisson case, log likelihoods were calculated as per methodology in Godwin and Böhning [28], with \(l_A(\hat{w},\hat{\lambda })=f_1\log \{\hat{w}+(1-\hat{w})\hat{\lambda }/(\exp {(\hat{\lambda })}-1)\} + \sum _{x=2}^{m}f_x \log \{(1-\hat{w})\hat{\lambda }^{x}/((\exp {(\hat{\lambda })}-1)x!)\}\). In the null hypothesis case, \(w=0\) so hence \(l_0(0,\tilde{\lambda })=\sum _{x=1}^{m}f_x \log \{\tilde{\lambda }^{x}/((\exp {(\tilde{\lambda })}-1)x!)\}\).

See also Böhning and van der Heijden [6] for a simplified LRT approach that uses the ‘zero-one truncated’ likelihood instead of the ZTOI likelihood, as this can act as an equivalent and is more straightforward to calculate.

As the LRT approximates a two-tailed \(\chi ^2\) test with 1 degree of freedom, a critical value of 2.706 was used (\(\alpha =0.05\)). Results are shown in Table 7. For all three versions of \(f_1\) to \(f_m\), the LRT in the geometric case was less than 2.706, indicating there was not enough evidence to say the ZTOI geometric was a closer fit than ZT geometric. However, in the Poisson case the LRT was larger than 2.706, indicating the ZTOI Poisson was a closer fit than ZT Poisson.

Table 7 LRT of ZTOI vs ZT geometric and of ZTOI vs ZT Poisson

In summary, for all three versions of \(f_1\) to \(f_m\), model-fitting diagnostics showed partial evidence of consistency with the ZT geometric as well as the ZTOI Poisson. Hence, estimates of N in Sect. 3.3 were based on both.

3.3 Case study results from population size estimation

Estimates of population size (N) are shown in Table 8. When the CRC data only contained captures of non-participants (right column), the number of participants (16) was added on to estimators of N because excluding participants from the data meant they were treated as outside the target population during the calculation of estimators.

Table 8 Estimated size of the case study population (\(\hat{N}\) with 95% confidence intervals)

N was consistently underestimated when participants were included in the CRC data (left and centre columns of Table 8). This was as expected due to the positive dependence found between participation and nomination. When instead excluding participants (right column of Table 8), underestimation was less severe in the ZT geometric’s BC Chao estimator and there was in fact overestimation in the ZT geometric’s WLS estimator. However, excluding the participants reduced the size of the observed population (n) (i.e., the number of individuals captured at least once in the data) from 61 to 45, leading to wider confidence-intervals for both the ZT geometric’s estimators. A guide to the minimum required size of n relative to N (a.k.a., the minimum capture proportion) is given in Xi et al. [55] which advises that n be at least 58% of N when \(N=200\). This was used as a rough guide of the required n when \(N=182\). This was not satisfied by any of the three versions of the CRC data, as the capture proportion was only 34% (61/182) when the data included nomination and participation capture occasions, 33% (60/182) when just including nomination capture occasions and 27% (\(45/(182-16)\)) when just including captures of non-participants. In the latter case, the capture proportion was calculated as \(n/(N-\)number of participants) because excluding participants from the CRC data meant treating participants as outside the target population when calculating estimators of N. The ZTOI Poisson’s BC Chao estimator produced a particularly low underestimate of N with all three versions of the CRC data, suggesting it had heavy reliance on a sufficient capture proportion.

4 Follow-up simulation study

To help inform the case study, the approach was next performed on thousands of simulated target populations where N was 182 or 500. In each case, we simulated a target population as a complete social network, drew a sample of participants from it and derived a CRC dataset (\(f_1\) to \(f_m\)) from the overlap of participants’ alters (social ties). Each participant’s network of alters (which could include other participants and/or non-participants) was treated as equivalent to a nomination capture occasion from the case study. For example, being an alter of three participants meant being captured three times via nomination. Like in the case study, being a participant was also considered as a possible extra capture occasion to make the maximum captures equal.

A further motivation was to consider the impact of using respondent-driven sampling (RDS) to select participants. In real-world settings, the often-elusive nature of target populations may often necessitate using a multi-wave sampling method like RDS to select participants. However, RDS poses a difficulty for the current paper’s approach because all non-seed participants are recruited via being an alter of another participant (described in Sect. 4.2), hence creating positive dependence between participation and nomination. We therefore derived CRC datasets from participants selected via either RDS or simple random sampling (SRS) to see how estimates of N would compare.

4.1 Simulation study target populations

Each simulated target population was a complete undirected social network, of either 182 or 500 actors, simulated by an exponential random graph model (ERGM). In this method, a network of actors (182 or 500 in this case) is simulated via a probability function that cycles through random pairs of actors (with replacement), each time generating a binary outcome of whether they are tied (1 = tie; 0 = no tie) (see e.g., [40, 52]). The probability function can include several parameterised terms that each make 1 or 0 more likely. While similar to logistic regression, a key difference is that the probability function often uses terms that cause partial dependence between binary outcomes.

The current study’s ERGM used four terms: ‘edges’, ‘k-star’, ‘k-triangle’ and ‘k-2path’. This has been advised as an effective baseline for modelling social networks [39, 52], particularly for modelling heterogeneity of degree size and transitivity (a clustering effect wherein ties are more likely between actors who share mutual ties). Further description is given in Online Resource 1.

Parameter values for edges, k-star, k-triangle and k-2path were set to − 4, 0.2, 1 and − 0.2 respectively, each with effect size (\(\lambda \)) = 2. These values were sourced from Pattison et al. [45], who found they were in line with findings from empirical datasets. These were thus treated as somewhat representative of typical social networks. Simulation was via R statistical software [48], using the ‘ergm’ package [31] from the ‘statnet’ suite. In the ergm package, the k-triangle and k-2path terms are substituted with the ‘geometrically weighted edgewise shared partners’ (GWESP) and ‘geometrically weighted dyadwise shared partners’ (GWDSP) terms respectively, which are equivalent terms but use \(\lambda = \log {(2)} = 0.693\) [52]. This combination of terms produced an average degree size of 3.13 in networks of 182 actors and 4.88 in networks of 500 actors.

As the ERGM’s probability function proceeds, it forms a Markov Chain Monte Carlo (MCMC) process wherein, after a suitable burn-in period, social network characteristics should reach convergence and settle into a regular pattern. Snapshots of complete social networks can then be taken, which is referred to as ‘sampling the graph’ [40]. We used sampled graphs as simulated target populations of either 182 or 500 actors. A wide gap between sampled graphs protected against autocorrelation. Like Pattison et al. [45], we sampled every 100,000th graph with a burn-in of 1,000,000. For example, to generate 1000 simulated target populations of 182 actors, the ERGM was specified as:

figure a

4.2 Simulation study sampling approach

When selecting participants via RDS, this began by randomly selecting five seed participants (without replacement) from the target population. From each participant’s alters, two further participants were randomly selected where possible from among those who were not already sampled. From this new wave of participants was selected a further wave of participants in the same way, and so on across several waves until the desired number of total participants was reached. This almost always resulted in four waves and a partial fifth wave.

When instead selecting participants via SRS, a random number generator was used to select the desired number of participants (without replacement) from the target population in just one sampling wave.

4.3 Simulation study derivation of capture–recapture data

As mentioned near the beginning of Sect. 4, a particular challenge when using RDS was that all non-seed participants were recruited via being an alter of another participant and were hence captured at least once via nomination. Moreover, the average shape/profile of \(f_1\) to \(f_m\) showed that participants had a higher probability of being captured multiple times than non-participants (Fig. 6). Therefore, like in the case study, three ways of deriving \(f_1\) to \(f_m\) were considered: (a) as captures from nomination and participation capture occasions; (b) as captures from nomination capture occasions; or (c) as captures of just non-participants. Option ‘b’ would include participants in the CRC data but effectively deflate their captures by not counting participation as an extra capture, partially offsetting positive dependence. Option ‘c’ would remove more positive dependence by removing all captures of participants, though at the cost of reducing the size of n. Meanwhile, when using SRS to select participants, only option ‘a’ was used because participation was independent of being nominated by (an alter of) others.

Fig. 6
figure 6

Average number of individuals captured x times from CRC datasets where participant selection was via RDS, the number of participants \(=60\) and \(N=182\). Left panel: nomination and participation capture occasions. Right panel: nomination capture occasions

4.4 Simulation study model-fitting diagnostics

As with the case study, model-fitting diagnostics were used to check how consistent each CRC dataset was with the ZT/ZTOI Poisson/geometric although only if \(\sum _{x=3}^{m}f_x>0\). Any gaps partway through \(f_1\) to \(f_m\) were removed by collapsing data from further along the tail. For a dataset to be deemed consistent with the ZT Poisson, it needed a non-significant \(\chi ^2\) GoF p-value for ZT Poisson, a more horizontal ratio plot in the Poisson case than the geometric, and a non-significant LRT for ZTOI Poisson. Consistency with ZT geometric was equivalently assessed. For a dataset to be consistent with ZTOI Poisson or ZTOI geometric, the LRT needed to be > 2.706. Like the case study, some datasets were consistent with the ZT version of one distribution and the ZTOI version of the other (e.g., consistent with ZT geometric and ZTOI Poisson).

4.5 Simulation study results

Table 9 shows mean estimates of N from CRC datasets where participant selection was via SRS and N was 182. The 2.5th and 97.5th percentiles of \(\hat{N}\) were used as 95% confidence intervals. As participant selection via SRS was independent of being nominated by (an alter of) others, \(f_1\) to \(f_m\) were derived from nomination and participation capture occasions in all three columns.

The left column of Table 9 was somewhat comparable to the case study, as each dataset was derived using 16 participants and the number of individuals (participants and non-participants) captured at least once (n) was, on average, 56. This was similar to the case study wherein n had been 61. Accuracy of estimators was low in this column, indicating that, even if there had been independence between nomination and participation, the case study would have needed more than 16 participants to produce reliable estimates of N.

Table 9 Estimated size of simulated populations (mean \(\hat{N}\)) where participant selection was via SRS and CRC datasets were derived from nomination and participation capture occasions in all columns

Using 40 participants (centre column of Table 9) increased \(\bar{n}\) to 112 (62% of N). This was now roughly in line with the advised minimum capture proportion in Xi et al. [55], which was that n be at least 58% of N when \(N=200\). Estimates of N in this column were somewhat more accurate overall, but especially when CRC datasets were consistent with ZT Poisson. This also occurred in the right column of Table 9 wherein N was 500 and each dataset was derived using 50 participants, producing an \(\bar{n}\) of 227 (45% of 500) that was in line with the advice in Xi et al. [55] that n be at least 44% of N when \(N=500\).

Tables 10 and 11 show estimates of N when instead using RDS to select participants, with N being 182 and 500 respectively. When CRC datasets only contained captures of non-participants (right column of both tables) the number of participants was added on to estimators of N because excluding participants from the data meant they were treated as outside the target population during the calculation of estimators.

In all columns of Tables 10 and 11 apart from the right column of Table 10, the number of participants was such that \(\bar{n}\) met the advised minimum capture proportion. Exclusion of participants from CRC datasets in the right column of both tables meant the capture proportion was \(\bar{n}/(N-\)number of participants) and, because n only included non-participants, it was more difficult to meet the advised minimum capture proportion. When N was 500 (right column of Table 11), 140 participants were used because this produced, on average, 219 non-participants and hence a capture proportion of \(219/(500-140)=61\%\). This satisfied the advised minimum proportion (58%) pertaining to \(N=200\), which was used as the threshold because \(500-140\) was between 200 and 500. However, when N was 182, the average number of non-participants did not reach the advised minimum capture proportion no matter how many participants were used. Instead, for demonstration purposes, the right column of Table 10 used 60 participants, producing a capture proportion of only \(51/(182-60)=42\%\). This was lower than the advised minimum (69%) pertaining to \(N=100\), which was used as the threshold because \(182-60\) was between 100 and 200.

Table 10 Estimated size of simulated populations (mean \(\hat{N}\)) where participant selection was via RDS and \(N=182\)
Table 11 Estimated size of simulated populations (mean \(\hat{N}\)) where participant selection was via RDS and \(N=500\)

Returning to Table 9, there were indications in its centre and right columns that, when nomination and participation were independent and n was sufficiently large, the current paper’s approach would more often produce a Poisson shape across \(f_0\) to \(f_m\). Approximately 70% of CRC datasets in the centre and right columns were consistent with ZT Poisson and produced accurate average estimates of N from ZT Poisson estimators. There was also greater accuracy from the ZT Poisson’s BC Chao estimator than that of the ZT geometric when only \(f_1\) and \(f_2\) were larger than 0. Meanwhile, approximately 28% of datasets in these columns were consistent with ZT geometric and overestimated N using ZT geometric estimators. Further work is needed to explore if this would occur outside the current simulation setting and, if so, how to obtain better estimates from datasets consistent with ZT geometric. For example, the ZT Conway–Maxwell Poisson (CMP) distribution can function as a mixture of ZT Poisson and ZT geometric and has its own WLS estimator of N [2, 51].

This suggested a similar trend towards ZT Poisson might occur when using RDS to select participants and using captures of just non-participants. This was indeed found in the right column of Table 11, wherein 7479 out of 10,000 datasets were consistent with ZT Poisson and produced only slight underestimates of N. However, this only occurred when the advised minimum capture proportion was satisfied, as the same trend was not present in the right column of Table 10 wherein the advised minimum proportion had not been met.

Further exploration of different sized simulated target populations found that, when using captures of just non-participants, n could, on average, only meet the advised minimum capture proportion when \(N>=300\) (described in Online Resource 1). Hence, for target populations where \(N<300\), it instead seemed optimal to use captures of participants and non-participants from just nomination capture occasions, which would partially offset positive dependence between participation and nomination while sustaining a large enough n to feasibly meet the minimum capture proportion. The centre column of Table 10 showed that, at least in the current simulation setting, this could lead to relatively accurate estimates of N from the approximately 70% of datasets consistent with ZT geometric in that column. This was because the more general tendency for ZT geometric estimators to overestimate N (seen earlier in the centre and right columns of Table 9) was somewhat counter-balanced by underestimation caused by positive dependence in the data.

Using RDS to select participants meant the Successive Sampling population size estimator (SS-PSE) of N could also be performed as a comparison [32, 33]. The SS-PSE is Bayesian and requires a prior estimate of N. For this, we used high, low or accurate priors based on those used in a simulation study in Handcock et al. [32]. Each prior was a beta distribution with a median of either \(2\times N\) in the high case, \((N+\)number of participants)/2 in the low case or N in the accurate case. The median of the posterior distribution of \(\hat{N}\) was taken as a point-estimate from each simulated target population. Calculation was via the ‘sspse’ R package [30]. See Online Resource 7 for an example. While the ‘visibility’ parameter of ‘sspse’ can improve estimates of N by imputing an adjusted degree size that factors in self-reporting bias, this was set to ‘false’ because participants’ social ties were not self-reported in this study.

SS-PSE estimates of N are shown in Table 12. The 2.5th and 97.5th percentiles of the point-estimates from across simulated target populations were used as 95% confidence intervals. We found that, when the median of the prior was accurate (i.e., N), the median of the posterior distribution tended to underestimate N. This may have been partly due to either transitivity in the ERGM model that generated the target populations, the relatively small mean degree size (3.13 when \(N=182\) and 4.88 when \(N=500\)) and/or using only five seed participants. This seeming sensitivity of the SS-PSE to one or more of these factors contrasted with the high level of accuracy we had seen from ZT Poisson estimators in the centre and right columns of Table 9 and the right column of Table 11, which appeared to be comparatively shielded by being based on only non-participant data.

Table 12 Successive Sampling population size estimate (SS-PSE) from 1000 simulated populations (mean \(\hat{N}\) with 95% confidence intervals)

5 Discussion

Results from the case study and simulation study gave a mixed picture of the current paper’s approach. Selecting participants via RDS in the simulation study and via voluntary participation in the case study both brought substantial positive dependence between participation and nomination, leading to deviation from the assumed independence of capture occasions and the assumed equal catchability of participants and non-participants.

However, the accuracy of ZT Poisson estimators in the right column of Table 11 was a promising early finding that, when using RDS to select participants, many CRC datasets produced quite accurate estimates of N when only including captures of non-participants and having a sufficiently large n after excluding participants. The accuracy of ZT geometric estimators in the centre column of Table 10 was also a promising finding that, when needing to include participants in the data to keep n sufficiently large, approximately 70% of CRC datasets produced relatively accurate estimates of N by including captures of participants and non-participants from just nomination capture occasions. Nevertheless, further work is needed to obtain better estimates of N from the remaining 30% of CRC datasets in both of those columns.

More work is also needed to explore other RDS settings and methodological extensions. For example, for target populations with more clustering, a system of design weights could potentially be used to offset clustering in each participant’s list of social ties. The current paper’s approach also needs a fuller comparison to the SS-PSE [32, 33] and/or Privatised Network Sampling (PNS) [26, 37], as they can factor in the multi-wave structure of RDS data.

Simulating target populations via ERGMs opens the door to exploring more varied target populations that could better reflect real-world settings. For example, while the average degree size in the current simulation study (3.13 when \(N=182\) and 4.88 when \(N=500\)) may have been in line with general human populations, degree size may be smaller among elusive populations and make one-inflation more prevalent. As another example, an ERGM simulating a population of drug users could incorporate covariate information, like a metric of drug use, that could be a major additional source of heterogeneity in individuals’ degree size and hence their capture-rate.

An advantage of sourcing captures from participants’ self-reported social networks is that many non-participants can be captured indirectly. Frank and Snijders [27] highlight a similar advantage in the Snowball Method, noting that it allows time and resources to be devoted to fewer but longer interviews that may encourage participation from stigmatised populations.

However, self-reported lists of social ties have a known susceptibility to memory-bias [4, 15]. This could be exacerbated if participants are asked to list individuals using pseudo-anonymised identifiers derived from several demographic variables, which may be necessary with stigmatised populations [16, 22, 26, 37]. For example, in Buchanan et al. [16], participants were asked to list individuals by combining abbreviations of initials, gender, age and district. Some populations might not be familiar enough to know such details about each other or remember them accurately, potentially leading to false matches. One possible solution is the Telefunken approach of Dombrowski et al. [22], in which participants are asked to include the last four digits of peoples’ phone numbers (if known) when compiling unique identifiers.

As the target population in the case study were sitting the same university course, they could feasibly appear in each other’s social networks. This was necessary to satisfy the equal-catchability assumption. For a larger or more fragmented target population, a stratified approach could potentially be used. For example, the ZT Poisson’s BC Chao estimator could be performed on M subsections of a target population via \(\hat{N}=\sum _{i=1}^m\{n_i+ f_{1i}(f_{1i}-1)/(2f_{2i}+2)\}\). Similarly, stratification could be used to factor in a variable (e.g., ethnicity) if it was suspected to affect the capture-rate of sample units.

6 Conclusion

There has been growing interest in deriving CRC data from self-reported social networks, and the current paper adds to this by considering a methodology for applying zero-truncated modelling to this type of data. This included an early exploration of how the approach could be applied when selecting participants via RDS, which was an important practical consideration for real-world settings. The approach still needs to be more fully compared to others, particularly as RDS brings a multi-wave structure to data that methods like Successive Sampling and Privatised Network Sampling can factor in. Further work is also needed to explore more varied target populations and key limitations such as task-fatigue, memory-bias, stratification of large populations and the necessary level of social ties/cohesion in the target population.