Skip to main content
Advertisement
Browse Subject Areas
?

Click through the PLOS taxonomy to find articles in your field.

For more information about PLOS Subject Areas, click here.

  • Loading metrics

Estimating Finite Rate of Population Increase for Sharks Based on Vital Parameters

  • Kwang-Ming Liu ,

    kmliu@mail.ntou.edu.tw

    Affiliations Institute of Marine Affairs and Resource Management, National Taiwan Ocean University, Keelung 202, Taiwan, George Chen Shark Research Center, National Taiwan Ocean University, 2 Pei-Ning Road, Keelung 202, Taiwan, Center of Excellence for the Oceans, National Taiwan Ocean University, 2 Pei-Ning Road, Keelung 202, Taiwan

  • Chien-Pang Chin,

    Affiliations Institute of Oceanography, Nation Taiwan University, Taipei 106, Taiwan, Fisheries Research Institute, Council of Agriculture, 199, Heyi Road, Keelung 202, Taiwan

  • Chun-Hui Chen,

    Affiliation Institute of Marine Affairs and Resource Management, National Taiwan Ocean University, Keelung 202, Taiwan

  • Jui-Han Chang

    Affiliation School of Marine Sciences, University of Maine, Orono, Maine 04469, United States of America

Abstract

The vital parameter data for 62 stocks, covering 38 species, collected from the literature, including parameters of age, growth, and reproduction, were log-transformed and analyzed using multivariate analyses. Three groups were identified and empirical equations were developed for each to describe the relationships between the predicted finite rates of population increase (λ’) and the vital parameters, maximum age (Tmax), age at maturity (Tm), annual fecundity (f/Rc)), size at birth (Lb), size at maturity (Lm), and asymptotic length (L). Group (1) included species with slow growth rates (0.034 yr-1 < k < 0.103 yr-1) and extended longevity (26 yr < Tmax < 81 yr), e.g., shortfin mako Isurus oxyrinchus, dusky shark Carcharhinus obscurus, etc.; Group (2) included species with fast growth rates (0.103 yr-1 < k < 0.358 yr-1) and short longevity (9 yr < Tmax < 26 yr), e.g., starspotted smoothhound Mustelus manazo, gray smoothhound M. californicus, etc.; Group (3) included late maturing species (Lm/L ≧ 0.75) with moderate longevity (Tmax < 29 yr), e.g., pelagic thresher Alopias pelagicus, sevengill shark Notorynchus cepedianus. The empirical equation for all data pooled was also developed. The λ’ values estimated by these empirical equations showed good agreement with those calculated using conventional demographic analysis. The predictability was further validated by an independent data set of three species. The empirical equations developed in this study not only reduce the uncertainties in estimation but also account for the difference in life history among groups. This method therefore provides an efficient and effective approach to the implementation of precautionary shark management measures.

Introduction

Sharks are the top predators in the ocean and play an important role in the marine ecosystem [1, 2]. Recent estimates indicated that shark populations have declined significantly in many regions of the world [3, 4, 5, 6, 7]. Worldwide trade in shark fin has increased dramatically. In 1980, the figure was less than 2000 MT, but by 2000 this had risen to 11602 MT [8], indicating a significant increase in shark exploitation during that period, but the shark landings deceased thereafter [9]. As a result, shark conservation and management have become issues of great concern in recent years. Many countries and international management and conservation organizations have taken their own steps with respect to sharks. For example, the USA, Australia, and the Maldives have regulations controlling the total allowable catch (TAC) and have also limited fishing grounds. According to the International Union for the Conservation of Nature and Natural Resources (IUCN) red list criteria, 32% of open ocean sharks are now considered threatened [9]. The convention on International Trade in Endangered Species of Wild Fauna and Flora (CITES) has placed the whale shark, Rhincondon typus, basking shark, Cetorhinus maximus, great white shark, Carcharodon carcharias, scallooped hammerhead, Sphyrna lewini, smooth hammerhead, S. zyganea, great hammerhead, S. mokarran, oceanic whitetip, Carcharhinus longimanus, porbeagle shark, Lamna nasus and manta rays, Manta spp. on its Appendix II list [10]. All these various measures serve to accentuate the urgency of shark management. Consequently, the regional fisheries management organizations have taken various management measures for sharks, i.e. prohibition of bigeye thresher, Alopias superciliosus, silky, Carcharhinus falciformis, oceanic whitetip, and Sphyrnidae except for Spyrna tiburo retaining on board in the Atlantic Ocean [11], prohibition of oceanic whitetip and thresher sharks, Alopias spp. retaining on board in the Indian Ocean [12], and prohibition of oceanic whitetip and silky shark retaining on board in the Pacific Ocean [13].

At least 498 species (8 orders) of sharks exist worldwide [14]. Many different life history traits have been found among these species. Maximum size ranges from 22 cm total length (TL) for the spined pigmy shark, Squaliolus laticaudus [15] to 1800 cm TL for the whale shark, Rhincodon typus [16]. Growth rates range from k = 0.034 yr-1 for the pike dogfish, Squalus acanthias [17] to k = 0.358 yr-1 for the spadenose shark, Scoliodon laticaudus [18]. In terms of reproductive strategy, three general categories have been identified: oviparity, viviparity, and aplacental viviparity. However, the litter size varies remarkably among species even for those falling within the same reproductive type. For example, for viviparous sharks, litter size ranges from six for the basking shark [19] to 82 for the blue shark, Prionace glauca [20]. For aplaental viviparous sharks, litter size ranges from two for the bigeye and pelagic thresher shark [21, 22] to more than 300 for the whale shark [23] (S1 Table). It is clear that, compared to teleosts, sharks have a far more complex and varied set of life history traits particularly the reproductive traits.

Due to the fact that sharks have a much lower commercial value than tunas and other teleost fish, catch, effort, and bycatch data for shark species are not readily available. Consequently, conventional stock assessment methods, such as surplus production and stock-recruitment models, have seldom been applied to examine shark population dynamics despite of recent works on blue sharks [24, 25]. However, because sharks have similar life histories to mammals, demographic models which have been applied to mammals have been found to better describe the dynamics of shark populations [26].

To date, the assessment of shark stock status using demographic analysis has, for the most part, been based on the hypothesis of a unit stock [26, 27, 28, 29, 30, 31, 32, 33, 34, 35, 36, 37]. However, this approach needs detailed information on vital parameters such as natural mortality, age at maturity, litter size, reproductive cycle, and longevity. It is difficult to apply this approach to species with limited available life history information [38], and only few demographic models consider density-dependent effects. There is an urgent need to manage and conserve shark stocks, and empirical equations based on vital parameters, which could be used to estimate the finite rate of population increase for particular categories of shark would make this task easier and more efficient based on a precautionary approach.

A number of authors have applied multivariate analyses, including principal component analysis (PCA), cluster analysis (CA), and regression analysis, to fish resource management [38, 39, 40, 41, 42]. Using PCA, Winemiller and Rose [39] identified four categories of species, namely periodical, opportunist, equilibrium, and intermediate. They suggested using size limits and maintaining adult abundance to manage and protect teleost fish larger than 100 cm and large sharks. King and McFarlane [38] also used PCA to identify three groups based on the vital parameters of growth rate, litter size, asymptotic length, and size at birth. They concluded that stock assessment is required every 1–2 years for species with a short life span, fast growth and small litter size. Cortés [41] applied CA to shark data and identified three groups based on litter size, longevity, asymptotic length, size at birth, and growth rate. Jennings et al. [40] used regression analysis to estimate vital parameters and predict fish abundance. Frisk et al. [43] described the effects of size at maturity and age at maturity on maximum observed length for elasmobranches. However, none of these studies has provided an empirical equation to estimate the finite rate of population increase.

The objectives of this study were 1) to use multivariate analysis to categorize sharks into groups based on their vital parameters, 2) to develop an empirical equation to estimate the finite rate of population increase for each group, and 3) to propose appropriate management measures for each group. It is hoped that these empirical equations can be applied to other shark species with limited life history information so as to achieve the goal of precautionary management.

Materials and Methods

In our search of the existing literature, we collected and analyzed vital parameter data from 83 studies. Only stocks with complete data (both age and growth and reproduction) were analyzed. In total, data of vital parameters were collected for 38 species of shark (62 stocks), comprising five orders and 10 families, as follows: one species in Hemiscylliidae of Orectolobiformes; seven species in Alopiidae, Cetorhinidae, and Lamnidae of Lamniformes; 27 species in Triakidae, Carcharhinidae, and Sphyrnidae of Carcharhiniformes; two species in Squalidae of Squaliformes; and one species in Hexanchidae of Hexanchiformes [44] (S1 and S2 Tables).

As conventional demographic analysis assumes that males are not the limiting factor regulating population growth, this study used data only from females. Where sex-specific parameters were not available, sexes-combined parameters were used. In total, 12 vital parameters were selected. These included five age and growth parameters: asymptotic length (L), growth coefficient (k), age at zero length (t0), maximum age (Tmax), and maximum observed length (Lmax); and seven reproduction parameters: age at maturity (Tm), reproductive strategy (R), size at maturity (Lm), size at birth (Lb), litter size (f), gestation period (Gp), and reproductive cycle (Rc). Different studies define vital parameters in slightly different ways. To account for this inconsistency, we used the following definitions:

  1. Size at maturity (Lm): size at 50% maturity, or mean size of mature specimens, or the mean of the maximum and minimum size at maturity if only the range of size at maturity was given.
  2. Size at birth (Lb): the smallest free swimmer, or the mean of the largest full term embryo and the smallest free swimmer.
  3. Maximum age (longevity) (Tmax): the maximum ages were assumed as follows: the blacknose shark, Carcharhinus acronotus in northern California waters, the northwestern Atlantic and Mexican waters, 26, 20, and 17 yrs respectively; the blue shark in the Northwest Pacific, 18 yrs; basking shark, 49 yrs; and shortfin mako, 41 yrs [19, 45, 46, 47, 48, 49]. The Tmax of other species was estimated from Taylor’s [50] equation as follow:
  4. Fecundity (f): the mean litter size of pregnant females, or the mean of the maximum and minimum of litter sizes.
  5. Maximum observed size (Lmax): the maximum size of observed sharks.
  6. Age at maturity (Tm): the age at 50% maturity, or the mean age of mature specimens, or the mean of the maximum and minimum age at maturity if only the range of age at maturity was given.
  7. Reproduction cycle (Rc): including gestation and resting periods, if only gestation data were available, Rc was estimated using data from similar species.

Input parameters

Large variations in Lb, Lm, and L were found between different species (S1 and S2 Tables) and this may affect the results of analysis. To eliminate the size-effect in our baseline analysis (scenario 1), we used 7 vital input parameters, namely Lb, the ratio between size at birth and asymptotic length (Lb/L), the ratio between size at maturity and asymptotic length (Lm/L), Tmax, Tm, k, and annual fecundity (f/Rc). As input parameters may affect the results of multivariate analysis, we simulated two other scenarios for comparison using different input parameters. Five vital parameters, namely Lm, Lb, f, k, and Tmax proposed by Cortés [41] were used in scenario 2, and six parameters (those in scenario 2 pluses one additional parameter, L) proposed by King and Mcfarlane [38] were used in scenario 3.

Demographic analysis

The conventional demographic analysis requires an input of natural mortality (M). Thus, Hoenig’s equations [51] were used to estimate the mean M for each stock dependent on the longevity as follows: ln(M) = ln(Z) = 0.941 − 0.873 * ln(Tmax), for L > 100 cm; ln(M) = ln(Z) = 1.46 − 1.01 * ln(Tmax) for L < 100 cm [26], where Z is total mortality. Natural mortality approaches Z when the fish stock is unfished or at light exploitation levels. We followed Krebs’s [52] formula to calculate demographic parameters, assuming a sex ratio of 1:1. Since , the initial intrinsic rate of population growth, r, can be calculated by iteration; net reproductive value per generation , where mx is fecundity at age x, lx is the survival rate until age x; generation length in years, ; the intrinsic rate of natural increase r = ln(R0)/G; and the finite rate of population increase, λ = er. The 95% confidence interval of λ were obtained from 1000 iterations using bootstrap method by randomly selecting M from the following four methods: (1) Hoenig’s equation [51], (2) M = 1.65/tmat [53], (3) M = 1.6 * k [53], (4) M = −ln(0.01)/tmax [51, 54].

Multivariate analysis

Due to inconsistencies in measurement units, our PCA used correlation matrices, R, rather than variance-covariance matrices. All parameters were log-transformed and normalized and the eiganvectors and eiganvalues were estimated. A non-parametric multiple dimensional scaling (NMDS) was used to draw the biplot. Vital parameters were reduced to several independent principal components and the scores of principal components were then analyzed using the cluster analysis.

The cluster analysis with Ward's method was used to estimate the scores of the first to third principal components and to draw the tree plot. Species with similar parameter values were grouped together and named according to their shared life history traits. After grouping, the general linear model (GLM) was used to develop an empirical equation for each group describing the relationship between the finite rate of population increase and vital parameters. A GLM was also used to describe the finite rate of population increase for all 62 shark stocks. The Akaike information criterion (AIC) and Bayesian information criterion (BIC) were both used for model selection [55]. A variance inflation factor (VIF) [56] was used to examine the multicollinearity of vital parameters in our multiple regression analysis: .

Multicollinearity exists among vital parameters when VIFj ≥ 10, and the parameter can be removed from the regression model.

Robustness of estimation

We used Jack-knife resampling simulations to estimate the robustness of our empirical equations. For each simulation, we randomly eliminated 1–3 samples from each group and repeated GLM estimations 1000 times. We also estimated the means and standard errors of intercept and coefficient of regression of each of these simulations. To validate the results of our empirical equations, an independent data set including three species which had not been used in developing the equations was substituted into the empirical equations.

Results

Vital parameters

Age and growth, reproduction, and litter size.

For age and growth parameters, the maximum value of L was 970 cm TL for the basking shark [19], the minimum was 71.5 cm TL for the spadenose shark, Scoliodon laticaudus [18] and the median was 265.4 cm TL. The maximum k value was 0.369 yr-1 for the whiskery shark, Furgaleus macki [57], the minimum was 0.034 yr-1 for the piked dogfish [17] and the median was 0.107 yr-1. The minimum Lmax was 69 cm TL for the spadenose shark [18], the maximum was 970 cm TL for the basking shark [19] (S1 Table) and the median was 242 cm TL.

For reproductive parameters, the maximum Lb was 174 cm TL for the pelagic thresher [22], the minimum was 14 cm TL for the spadenose shark [18, 58], and the median was 61 cm TL. Size at maturity, Lm, ranged from a minimum of 34 cm TL for the spadenose shark, to a maximum of 500 cm TL for the basking shark [19] with a median of 185 cm TL. The age at maturity ranged from 2 yrs for the spadenose shark to 30 yrs for the sandbar shark [59]. The gestation period ranged from 5 months for the bonnethead shark Sphyrna tiburo [60] to 31 months for the basking shark [17], with a median of 12 months.

Thirty of the 62 stocks (48.4%) have a 2-yr reproductive cycle, e.g. the bull shark in northern Mexican waters [61] and the spinner shark [62]; 15 stocks (24.2%) have a 1-yr cycle, e.g. the porbeagle shark, [63], and the Carcharhinid sharks [64]; 17 stocks (27.4%) have a 3-yr cycle, e.g. the school shark, Galeorhinus galeus in Brazilian waters [65], and the shortfin mako, Isurus oxyrinchus in the northwestern Pacific [16] (S2 Table).

Litter size varies remarkably among species even for the same reproductive trait. For example, for viviparous sharks, the smallest litter size was six for the basking shark [19], while the largest was 82 for the blue shark [20]. A similar situation was found for aplacental viviparous sharks, with litter size ranging from two for the bigeye and pelagic thresher sharks [21, 22] to 55 for the tiger shark, Galeocerdo cuvier [66] (S2 Table).

The ratios of Lb/L, Lm/L, and Lb/Lm.

The Lb/L ratios of the 38 species (62 stocks) ranged from 0.12 to 0.47 with a median value of 0.21. Fifty-two of 62 stocks (83.9%) fell in the range 0.12–0.28 (mean = 0.23), while the remaining 10 stocks (16.1%) were in the range 0.30–0.47. The highest value was 0.47 for the spottail shark Carcharhinus sorrah in northern Australia and the blacknose shark C. acronotus in northwest Atlantic, while the lowest value was for the pike dogfish Squalus acanthias in the southeastern Black Sea (S1 Table).

The Lm/L ratios ranged from 0.45 for the thresher shark Alopias vulpinus in Californian waters to 0.94 for the whiskery shark Furgaleus macki in southwest Australian waters. The median value was 0.68. Thirty-four of 62 stocks (54.8%) fell in the range 0.6–0.8, seventeen (27.4%) had values between 0.45 and 0.59, and eleven stocks (17.7%) were in the range 0.81–0.94.

The Lb/Lm ratios ranged from 0.07 for the sevengill shark, Notorynchus cepedianus to 0.86 for the blacknose shark, with a median value of 0.33. Fifty-three of 62 stocks (85.5%) fell in the range 0.2–0.5, two stocks (3.2%) had values between 0.17 and 0.19, and seven stocks (11.3%) were in the range 0.52–0.86 (S1 Table).

Maximum age and natural mortality.

The maximum age ranged from 7 years for the spottail shark, Carcharhinus sarrah to 81 years for the pike dogfish; for 45 of 62 stocks (72.6%) the range was 7–37 years, while 17 (27.4%) fell in the range 38–81 years (S1 Table). Natural mortality rates estimated from Hoenig’s [51] equation range from 0.055 yr-1 for the pike dogfish to 0.474 yr-1for the spottail shark (S1 Table).

Litter size per year.

The litter size per year ranged from 1.67 for the longnose spurdog, Squalus blainville, in Italian waters to 41 for the blue shark in California, with a median value of 8.5. Fifty of 62 stocks (80.6%) ranged from 1.67–8.06; ten stocks (16.1%) fell in the range 11–19; and two stocks (3.2%) were in the range 39.5–41 (S2 Table).

Finite rate of population increase

The finite rate of population increase estimated from conventional demographic analysis ranged from 0.929±0.064 for the grey reef shark, Carcharhinus amblyrhynchos to 1.470 ± 0.114 for the spadenose shark, Scoliodon laticaudus. Thirty-nine of 62 stocks (62.9%) fell in the range 1.0084 ± 0.060–1.1453 ± 0.0722, 19 stocks (30.6%) had values greater than 1.1542 ± 0.0901, and four stocks (6.5%) had values lower than 1 (S3 Table).

Multivariate analyses

Non-parametric multiple dimensional scaling.

The bivariate plot of the two dimensional NMDS was showed in Fig 1:

  1. In dimension 1, the positive scores represent fast growing species with large k, such as the brown smooth–hound and oceanic whitetip sharks; the negative scores represent late maturing species with large Tmax, and extended longevity, such as the dusky and sandbar sharks.
  2. In dimension 2, the positive scores represent species with high Lm/L ratio and high annual fecundity, such as the tiger shark and smooth hammerhead; the negative scores represent species with large size at birth and large Lb/L ratio, such as the three threshers.
thumbnail
Fig 1. The biplot of two dimensional NMDS.

Black labels are species, blue arrows are life history traits.

https://doi.org/10.1371/journal.pone.0143008.g001

Cluster analysis and empirical equations of the finite rate of population increase.

In scenario 1, three groups were identified based on the cluster analysis (Fig 2):

Group 1: Slow growing species (0.034 yr-1 < k < 0.111 yr-1) with high maximum age (26 yr < Tmax < 81 yr). A total of 17 stocks fell into this group, most being large sharks such as the shortfin mako, and dusky shark. The maximum age ranged from 26 yrs for the tiger shark to 81 yrs for the piked dogfish, with the majority of stocks (10) being in the range 28–51 yrs. Longevity, age at maturity and fecundity per year were significant parameters in this group. The empirical equation for estimating the finite rate of population increase is: λ′ = 1.064 + 0.076 * ln(Tmax) − 0.128 * ln(Tm) + 0.035 * ln(f / Rc) (n = 17, r2 = 0.97, sd = 0.0070) (Table 1).

thumbnail
Fig 2. Dendrogram from a cluster analysis of seven vital parameter of 62 stocks from 38 species of sharks.

The grouping shows similarities in life history traits among species and stocks from scenario 1.

https://doi.org/10.1371/journal.pone.0143008.g002

thumbnail
Table 1. Vital parameters of the species in group 1 from scenario 1.

https://doi.org/10.1371/journal.pone.0143008.t001

Group 2: Fast growing species (0.103 < k < 0.358 yr-1), with small Tmax (9 < Tmax < 26). A total of 16 stocks fell into this group. The value of k for 11 of 16 stocks (68.75%) fell in the range 0.103–0.18 yr-1, with the remaining five stocks (61.25%) ranging from 0.21 to 0.358 yr-1. The largest k value was for the spadenose shark (k = 0.358 yr-1), while the smallest was for the oceanic whitetip shark (k = 0.103 yr-1). Tmax ranged from 9 yrs for the spadenose shark to 27 yrs for the oceanic whitetip shark. Most species in this group were small size, such as the spotless smoothhound, starspotted smoothhound, and whitespotted bamboo shark. The significant parameters for this group were found to be the ratio between size at maturity and asymptotic length, longevity, age at maturity, growth rate and fecundity per year. The empirical equation for estimating the finite rate of population increase is: (n = 16, r2 = 0.95, sd = 0.0359) (Table 2).

thumbnail
Table 2. Vital parameters of the species in group 2 from scenario 1.

https://doi.org/10.1371/journal.pone.0143008.t002

Group 3: Late-maturing species (Lm/L ≥ 0.67) with moderate Tmax (Tmax ≤ 29 yr). Lm/L ranged from 0.67 for the silky shark to 0.94 for the Australian whiskery shark, with 13 of 29 stocks (72.2%) in the range 0.75–0.85. A second characteristic of this group was larger values of f/Rc and Lb/L. The species with low f/Rc have high Lb/L such as the pelagic thresher (f/Rc = 1, Lb/L = 0.45) and blacknose shark (f/Rc = 1.25, Lb/L = 0.47). Conversely, those with high f/Rc have low Lb/L such as the blue shark (f/Rc = 20.5, Lb/L = 0.16), sevengill shark (f/Rc = 19.75, Lb/L = 0.15), and tiger shark (f/Rc = 13.75, Lb/L = 0.15). The empirical equation for estimating the finite rate of population increase is: λ′ = 1.377 − 0.057 * ln(Lb) + 0.169 * ln(Lb/L) − 0.261 * ln(Lm/L) + 0.160 * ln(Tmax) − 0.340 * ln(Tm) + 0.152 * ln(f/Rc) (n = 29, r2 = 0.93, sd = 0.0297) (Table 3). Since VIF < 10, this indicates an absence of multicollinearity for the three equations.

thumbnail
Table 3. Vital parameters of the species in group 3 from scenario 1.

https://doi.org/10.1371/journal.pone.0143008.t003

To sum up, Group 1 species reach asymptotic length at an older age (Tmax > 25 yr); than Group 2 species (Tmax < 13 yr). Only Groups 1 and 2 have overlapping values of Lm/L and Tmax between the groups.

The empirical equation for 62 stocks combined is λ′ = 1.116 − 0.029 * ln(Lb) + 0.108 * ln(Lb/Lm) − 0.141 * ln(Lm/L) + 0.154 * ln(Tmax) − 0.242 * ln(Tm) + 0.119 * ln(f/Rc) (n = 62, r2 = 0.77, sd = 0.0654).

The results of Jack-knife simulations indicated the robustness of the empirical equations for Groups 1–3, as well as the combined equation (Table 4). Using 1000 simulations, the coefficients of variation for each parameter mean of Groups 1–3 and the combined-group equation were 4.28%–21.48%, 3.17%–13.55%, 4.11%–97.27%, and 3.45%–12.28%, respectively. Moreover, the 95% confidence intervals of the parameter means also indicate that the parameters of each of the four equations are statistically significant and robust (Table 4).

thumbnail
Table 4. The partial regression coefficients and their coefficient of variation of the empirical equations for Groups 1 to 3 and combined-group.

https://doi.org/10.1371/journal.pone.0143008.t004

In scenario 2, two groups were identified by cluster analysis. No significant relationship was found between vital parameters and λ´ for cluster 1 (p = 0.123; n = 13), while only fecundity was correlated to λ´ for cluster 2: λ' = 1.1136 + 0.0038f, (p = 0.037; n = 49). Two groups were also identified for cluster 3 but no significant relationship was found between vital parameters and λ´ for either cluster in this scenario.

Validation of empirical equations.

The independent data set for validation the results of our empirical equations included the vital parameters for Groups 1–3 species, leopard shark Triakis semifasciata, grey nurse shark Carcharias taurus and gummy shark Mustelus antarcticus. The predicted values of λ’ for each group (1.154, 1.002, and 1.268) showed good agreement with those derived from conventional demographic analysis (1.199, 0.977, and 1.239) (Fig 3, Tables 13). High correlations between predicted λ’ and λ for Groups 1–3 and combined equation (r2 = 0.97, 0.95, 0.93 and 0.77, respectively) were found.

thumbnail
Fig 3. Box plot of vital parameters for Groups 1, 2, and 3.

(Figs 3A, 3B, and 3C, respectively). Lb: size at birth, Lm: size at maturity, Linf: asymptotic length, Lb/ Linf: ratio of Lb and Linf, Tmax: maximum age, K: growth coefficient, f/RC: annual fecundity.

https://doi.org/10.1371/journal.pone.0143008.g003

Discussion

In this study, we used the vital parameters of 62 shark stocks to develop empirical equations to estimate population increase rates. Although these data have been filtered by existed knowledge, neither quantitative analysis nor rigorous criterion was used in choosing these data set. Therefore, the inconsistence of data quality may occurred in this study. Thorson et al. [67] mentioned that meta-analyses employing hierarchical models could account for experimental design differences, covariates and non-random assignment of study sites to treatment and control groups, and would likely increase precision for effect-size estimates. Hierarchical models should be included in the analysis in the future.

Factors affect estimate ofλ

Several factors may affect the estimate of λ.

Body length.

To deal with inconsistencies in length measurement found in the literature, this study converted all lengths to TL, other than those which were already designated as total length. This standardization improved the quality of our results.

Growth coefficient.

Inconsistencies were also found in the literature with respect to age determination, even for the same species. For example, Branstetter [68] reported annual band pair formation for the scalopped hammerhead in the northwestern Mexican waters, while Chen et al. [69], Anislado-Tolentino et al. [70] and Kotas et al. [71] reported a biennial formation for the same species in the northeastern Taiwanese waters, southern coast of Mexico, and Brazil waters, respectively. Similarly, for shortfin mako, Pratt and Casey [72] reported biannual band pair deposition while Cailliet et al. [73] suggested an annual deposition. Neer et al. [74] have indicated that the accuracy of age determination significantly affects stock assessment. To ensure as wide a range of data as possible, and allow for the above-mentioned inconsistencies, this study collected and analyzed growth parameters for the same species in different waters.

Under Branstetter's [75] categorization, k values of 0.05–0.10 yr-1 indicate slow growth, 0.10–0.20 yr-1 indicate moderate growth, and 0.20–0.50 yr-1 indicate rapid growth. In our study, examples of slow growth species included the dusky shark (k = 0.043 yr-1; [76]), shortfin mako (k = 0.05 yr-1, [49]), and porbeagle shark (k = 0.061 yr-1; [77]); examples of moderate growth species included the blacknose shark (k = 0.114 yr-1; [78]), spinner shark (k = 0.151 yr-1; [62]), and blue shark (k = 0.1614 yr-1; [48]); and examples of rapid growth species included the whitespotted bamboo shark (k = 0.224 yr-1; [79]), and grey reef shark (k = 0.294 yr-1; [41]). As the growth parameters used in this study covered a wide range of growth rates, our results derived from this study can be applied to the species with different growth rates.

Reproduction cycle.

Wourms [80] identified three basic types of reproductive cycle: (1) reproduction occurring throughout the year; (2) a partially defined annual cycle with one or two peaks during the year; and (3) a well-defined annual or biennial cycle. The pelagic thresher shark [22] is an example of a first-category type, while the epaulette shark, Hemiscyllium ocellatum [81], falls into the second category. Examples of a third-category type include the shortfin mako, with a 3-year reproduction cycle (2 years of gestation and 1 year of resting) [16], and the spinner shark, with a 2-year reproduction cycle (1 year of gestation and 1 year of resting) [62]. The result of shark stock assessment based on demographic analysis is affected by both the gestation and resting periods [49]. The estimates in this study take both gestation and resting periods into account and therefore, we believe, provide more accurate and realistic results.

Litter size.

The litter size may be underestimated when it was estimated based on the carcasses at the fish market. Embryos may be lost during the capture process for viviparous or aplacental viviparous sharks which result in the underestimation of litter size. Branstetter [82] and Bonfil [83] documented that female silky sharks may have aborted pups from uterus during capture if litter sizes less than 5 pups. To reduce the uncertainty, future study should focus on collecting more reliable litter size information from on board observation.

The ratio of Lb/L.

Branstetter [75] documented a trade-off between litter size and size at birth. Species with small litter size compensate by having a larger Lb and higher Lb/L. Joung [84] stated that the ratio of Lb/L ranged from 0.15–0.35 for most elasmobranches. With few exceptions, the species in this study were in this range. Usually, a negative relation between f/Rc and Lb/L was evident. For example, the blue shark (Lb/L = 0.14;f/Rc = 14.5); blacknose shark (Lb/L = 0.38;f/Rc = 4.5), pelagic thresher shark (Lb/L = 0.45;f/Rc = 2) and spottail shark (Lb/L = 0.47; f/Rc = 3). Unlike the studies by Cortés [41] and King and Mcfarlane [38], Lb/L instead of Lb/Lmax was used as an input parameter in this study. Given the high correlation between Lmax and L for the 62 stocks in this study (r = 0.937), our approach should be considered an acceptable alternative.

The ratio of Lm/L.

Since there is considerable variation in size at maturity among species (34–336.6 cm TL, S2 Table), analysis based on the input parameter Lm might produce bias. This study therefore used the ratio of Lm/L instead. Compagno [58] stated that for sharks, Lm/L was 0.6–0.8. Pratt and Casey [85] also concluded that the Lm/L of most elasmobranches is above 0.5. The 38 species analyzed in this study included all the maturing types defined by Joung [84]. That is, early maturing species (Lm/L < 0.6), such as the thresher shark (Lm/L = 0.45), and blacknose shark (Lm/L = 0.46) moderate maturing species (0.6 < Lm/L < 0.8), such as the dusky shark (Lm/L = 0.68), whitespotted bamboo shark (Lm/L = 0.7), and silky shark (Lm/L = 0.77); and late maturing species (Lm/L > 0.8), such as the whiskery shark (Lm/L = 0.94), tiger shark (Lm/L = 0.9) and sevengill shark (Lm/L = 0.84).

Maximum age.

Skomal and Natanson [86] pointed out that using the maximum observed age to represent the maximum age might result in underestimates. Equations developed by Taylor [50], Fabens [87] and Pauly [88] are commonly used to estimate maximum age. In this study, the values estimated from the latter two equations were much higher than the maximum observed age, and therefore Taylor’s [50] equation was used. Froese and Binohlan [89] suggested that for most sharks the total length was in the range 100–300 cm with a ratio Lmax/L of 0.97–0.987. Chen and Yuan [37] claimed that the maximum age estimated from Taylor’s [50] equation is more reasonable than those derived from other equations. In this study, apart from the blacknose, basking, blue, and shortfin mako, for which maximum ages were adopted from the literature, the Tmax was estimated using Taylor [50], and we believe the results to be reasonable.

Estimate of natural mortality

The natural mortality of marine animals is difficult to estimate. Ohsumi et al. [90] proposed a linear relationship between L and longevity to estimate the natural mortality of the minke whale. Pauly’s [91] empirical equation between M and L, k, and habitat mean water temperature has been widely used to estimate M for teleost fishes. Several attempts have been made by other authors e.g., Peterson and Wroblewski [92], Chen and Watanabe [93] and Jensen [53], but these studies have also focused on teleosts. As little is known of the life history parameters of sharks, Hoenig’s [51] relationship between longevity and total mortality has been adopted by many authors [26, 27, 28, 29, 30, 32, 94]. Hoenig [51] put forward three empirical equations, of which Cortés [26] suggested using the equation for marine mammals to represent sharks larger than 100 cm and the equation for teleosts to represent sharks smaller than 100 cm. Our study follows this suggestion. Chen and Yuan [37], on the other hand, used Hoenig’s [51] equation for teleosts to estimate M for sharks greater than 100 cm. We believe this might have led to an overestimation. Recently, Then et al. [95] suggested that a new tmax-based estimator is better than other empirical equations in natural mortality estimation. Although this method has not been tried in this study, since the empirical equations [51] we used is also tmax-based equation and is the most frequently used method for elasmobranchs, we believe our estimates are robust.

Estimation of λ

Cortés [34] and Chen and Yuan [37] have applied demographic analysis to sharks using vital parameter data. The estimates of λ in this study using conventional demographic methods were comparable to those of Cortés [34]. The λ value of sharks derived by Chen and Yuan [37] may be an underestimate as they calculated natural mortality using Pauly’s [91] method, which is not suitable (an overestimation) for sharks.

Cortés [33] estimated intrinsic population growth rate through stochastic demographic analysis by applying Monte Carlo simulations based on Tm, Tmax, fecundity, and M. In this study, both gestation and resting periods were included in the calculation of λ and stochastic effects have also been considered in estimating the confidence interval of λ. Therefore, we believe this produces a reasonable estimate.

Input parameters.

The reproductive cycle was not used as an input parameter in scenarios 2 and 3, but was included in scenario 1. In addition, variations in size among species were reduced by using the ratios Lb/L and Lm/L rather than Lb and Lm. We believe the output of scenario 1 is more reasonable than those derived from scenarios 2 and 3.

Cluster analysis

The groups defined in scenario 1 have distinct life history characteristics. With a few exceptions, the 62 stocks can be correctly categorized based on their life history parameters. In contrast, distinct life history characteristics did not appear in scenarios 2 and 3. Therefore, we believe that the results obtained in scenario 1 are likely to be more realistic than those in scenarios 2 and 3.

Validation and application of empirical equation

The high correlation between predicted λ’ and λ for Groups 1–3 and combined equation and the randomly distributed residuals, suggest that the empirical equations developed in this study can predict λ precisely than other models, and also need fewer vital parameters in Groups 1 and 2. It therefore provides an effective and efficient approach to shark management. The predicted values of λ’ for each group of the independent data set showed good agreement with those derived from conventional demographic analysis suggesting that the empirical equations can be applied to predict λ for other shark species. In other words, the empirical equations derived in this study reduce the uncertainties, and increase the accuracy, of population increase estimates, even without the inclusion of a natural mortality variable.

Bayesian production model has been used in shark stock assessment [96, 97, 98, 99]. One of the key input prior for this model is the intrinsic population growth rate r (r = ln(λ)). Our empirical equations, which can accurately estimate the λ can enhance the ability of stock assessment.

Uncertainty of vital parameter

The reproduction cycle is one of the most ambiguous vital parameters. This information is available for only 21 of 62 stocks in the literature. For the remaining stocks estimates were made using data on gestation periods and by referring to the reproduction cycle of similar species. However, discrepancies may exist due to variations in geography and these may result in inaccurate estimates of annual litter size. Most age at maturity and maximum age values have been estimated from the VBGE, but many uncertainties have been found, including sample size, specimen size range, band reading etc. These uncertainties may lead to inaccurate estimates of λ in empirical equations.

Management measures

Based on life history characteristics, conventional studies have categorized fish strategies into r and K types. Fish with r strategy are small size, early-maturing, and have a short life span. Those with K strategy are large size, late-maturing, and have an extended life span. These strategies correspond to the management measures of teleost and chondrichthyan fishes. Walker [100] suggested that a K rather than r strategy should be adopted for shark management and marine mammals. Also, management measures should vary according to the catch and stock status of different species and areas.

Recommendations for management

In this study, management recommendation was given only for scenario 1, as this was considered more realistic than the other two scenarios. Group 1 stocks are mostly large, slow-growing species with small litter size. Given that these populations recover slowly even when they experience slight overfishing, a protection of adults or TAC management measure has been suggested e.g., school shark, Galeorhinus galeus [101], and shortfin mako [49]. Group 2 stocks are mostly small, fast-growing species with large litter size. Regular stock assessment with management of the fishing area and fishing season closure has been suggested [39]. Group 3 stocks are mostly late-maturing species which recover slowly. A reduce of catch or TAC management has been suggested e.g., thresher shark, Alopias vulpinus [4] and pelagic thresher [102, 103].

Conclusions

Conventional stock assessment analysis requires fishing effort or other biological information. In this study, we provide a new approach to the accurate estimation of the finite rate of population increase. The empirical equations developed herein not only provide accurate predictions of λ but also reduce estimate bias resulting from parameter uncertainties. We believe that this is an effective and efficient approach to the implementation of precautionary shark management measures. However, we recognize that these equations could be improved further. Our study considered only 38 of 498 shark species existing worldwide, [15]. Therefore, our estimates may not take into account all the various life history traits of different shark species. Moreover, potentially influential environmental factors such as water temperature, water depth, and salinity were not considered in this study. To improve the accuracy and usefulness of these empirical equations, we suggest that future studies be directed toward these areas.

Supporting Information

S1 Table. Age and growth parameters for the 62 stocks (38 species) of sharks used in this study.

https://doi.org/10.1371/journal.pone.0143008.s001

(DOCX)

S2 Table. Reproductive parameters for the 62 stocks (38 species) of sharks used in this study.

https://doi.org/10.1371/journal.pone.0143008.s002

(DOCX)

S3 Table. Finite population increase rate derived from demographic analysis for 62 stocks (38 species) of sharks.

https://doi.org/10.1371/journal.pone.0143008.s003

(DOCX)

Author Contributions

Conceived and designed the experiments: KML. Performed the experiments: CHC CPC. Analyzed the data: CHC CPC JHC KML. Contributed reagents/materials/analysis tools: KML CHC CPC JHC. Wrote the paper: KML.

References

  1. 1. Stevens JD, Bonfil R, Dulvy NK, Walker PA (2000) The effects of fishing on sharks, rays, and chimeras (chondrichthyans), and the implications for marine ecosystems. ICES J Mar Sci 57: 476–494.
  2. 2. Schindler DE, Essington TE, Kitchell JF, Boggs C, Hilborn R (2002) Sharks and tunas: fisheries impacts on predators with contrasting life histories. Ecol Appl 12: 735–748.
  3. 3. Baum JK, Myers RA, Kehler DG, Worm B, Harley SJ, Doherty PA (2003) Collapse and conservation of shark populations in the northwest Atlantic. Science 299: 389–392. pmid:12532016
  4. 4. Myers RA, Worm B (2003) Rapid worldwide depletion of predatory fish communities. Nature 423(6937): 280–283. pmid:12748640
  5. 5. Burgess GH, Beerkircher LR, Cailliet GM, Carlson JK, Cortés E, Goldman KJ, et al. (2005) Is the collapse of shark populations in the Northwest Atlantic Ocean and Gulf of Mexico real? Fisheries 30(10): 19–26.
  6. 6. Camhi MD, Lauck E, Pikitch EK, Babcock EA (2009) A global overview of commercial fisheries for open ocean sharks. In: Camhi MD, Pikitch EK, Babcock EA, editors. Sharks of the open ocean: biology, fisheries and conservation. Blackwell, Oxford, UK, pp 166–192. https://doi.org/10.1002/9781444302516
  7. 7. Dulvy NK, Baum JK, Clarke S, Compagno LJV, Cortés E, Domingo A, et al. (2008) You can swim but you can't hide: the global status and conservation of oceanic pelagic sharks and rays. Aquatic Conserv: Mar Freshw Ecosyst 18(5): 459–482.
  8. 8. Clarke SC, McAllister MK, Milner-Gulland EJ, Kirkwood GP, Michielsens CGJ, Agnew DJ, et al. (2006) Global estimates of shark catches using trade records from commercial markets. Ecol Lett 9(10): 1115–1126. pmid:16972875
  9. 9. Davidson LNK, Krawchuk MA, Dulvy NK (2015) Why have global shark and ray landings declined: improved management of overfishing? Fish Fish
  10. 10. The CITES Appendices (2013) www.cites.org/eng/app/index.php.
  11. 11. ICCAT Compliance Committee (2011) Report of the Inter-sessional Meeting of the Compliance Committee, Barcelona, Spain, February 21–25, 2011.
  12. 12. IOTC WPEB (2010). Report of the Sixth Session of the IOTC Working Party on Ecosystems and Bycatch. Victoria, Seychelles, October 27–30, 2010.
  13. 13. WCPFC TCC, Summary Report (2012) Technical and Compliance Committee Eighth Regular Session. Busan, Korea, August 7–15, 2012.
  14. 14. Compagno L, Dando M, Fowler S (2005) A field guide to sharks of the world. Princeton University Press, 496pp.
  15. 15. Froese R, Pauly D Editors. 2015. FishBase. World Wide Web electronic publication. www.fishbase.org, version (02/2015).
  16. 16. Joung SJ, Hsu HH (2005) Reproduction and embryonic development of the shortfin mako, Isurus oxyrinchus Rafinesque, 1810, in the northwestern Pacific Zool Stud 44(4): 487–496.
  17. 17. Ketchen KS (1975) Age and growth of dogfish Squalus acanthias in British Columbia waters. J Fish Res Board Can 32(1): 43–59.
  18. 18. Devadoss P (1998) Growth and population parameters of the spade nose shark, Scoliodon laticaudus from Calicut coast. Indian J Fish 45(1): 29–34.
  19. 19. Pauly D (2002) Growth and mortality of the basking shark Cetorhinus maximus and their implications for management of whale sharks Rhincodon typus. Paper presented at the Elasmobranch Biodiversity, Conservation and Management. Proceedings of the International Seminar and Workshop, Sabah, Malaysia, July 1997.
  20. 20. Cailliet GM, Bedford DW (1983) The biology of three pelagic sharks from California waters, and their emerging fisheries: a review. Reports of Calif Coop Ocean Fish Invest 24: 57–69.
  21. 21. Chen CT, Liu KM, Chang YC (1997) Reproductive biology of the bigeye thresher shark, Alopias superciliosus, (Lowe, 1839)(Chondrichthyes: Alopiidae), in the northwestern Pacific. Ichthyol Res 44 (3): 227–235.
  22. 22. Liu KM, Chen CT, Liao LH, Joung SJ (1999) Age, growth, and reproduction of the pelagic thresher shark, Alopias pelagicus, in the northwestern Pacific. Copeia 1999: 68–74.
  23. 23. Joung SJ, Chen CT, Clark E, Uchida S, Huang WYP (1996) The whale shark, Rhincodon typus, is a livebearer: 300 embryos found in one 'megamamma' supreme. Environ Biol Fishes 46(3): 219–223.
  24. 24. ICCAT Shark Working Group Report 2012 Shortfin mako stock assessment and ecological risk assessment meeting.
  25. 25. ISC Shark Working Group Report (2014) Stock assessment and future projections of blue shark in the North Pacific.
  26. 26. Cortés E (1998) Demographic analysis as an aid in shark stock assessment and management. Fish Res 39: 199–208.
  27. 27. Hoff TB (1990) Conservation and management of the western North Atlantic shark resource based on the life history strategy limitations of sandbar sharks. Ph. D Dissertation, University of Delaware, Newark, DE, 282pp.
  28. 28. Cailliet GM (1992) Demography of the central California population of the leopard shark (Triakis semifasciata). Aust J Mar Freshwater Res 43: 183–193.
  29. 29. Cortés E (1995) Demographic analysis of the Atlantic sharpnose shark, Rhizoprionodon terraenovae, in the Gulf of Mexico. Fish Bull 93: 57–66.
  30. 30. Sminkey TR, Musick JA (1995) Age and growth of the sandbar shark, Carcharhinus plumbeus, before and after population depletion. Copeia 1995: 871–883.
  31. 31. Cortés E, Parsons GR (1996). Comparative demography of two populations of the bonnethead shark (Sphyrna tiburo). Can J Fish Aquat Sci 53: 709–718.
  32. 32. Au DW, Smith SE (1997) A demographic method with population density compensation for estimating productivity and yield per recruit of the leopard shark (Triakis semifasciata). Can J Fish Aquat Sci 54: 415–420.
  33. 33. Liu KM, Chen CT (1999) Demographic analysis of the scalloped hammerhead, Sphyrna lewini, in the northwestern Pacific. Fish Sci 65(2): 218–223.
  34. 34. Cortés E (2002) Incorporating uncertainty into demographic modeling: application to shark populations and their conservation. Conserv Biol 16 (4): 1048–1062.
  35. 35. Mollet HF, Cailliet GM (2002) Comparative population demography of elasmobranches using life history tables, Leslie matrices and stage-based matrix models. Mar Freshwater Res 53: 503–516.
  36. 36. Takeuchi Y, Senba Y, Nakano H (2005). Demographic analysis on Atlantic blue and shortfin mako sharks. Collec Vol Sci Pap ICCAT 58 (3): 1157–1165.
  37. 37. Chen PM, Yuan WW (2006) Demographic analysis based on the growth parameter of shark. Fish Res 78: 374–379.
  38. 38. King JR, Mcfarlane GA (2003) Marine fish life history strategies: applications to fishery management. Fish Manag Ecol 10: 249–264.
  39. 39. Winemiller KO, Rose KA (1992) Patterns of life-history diversification in North American fishes: implications for population regulation. Can J Fish Aquat Sci 49 (10): 2196–2218.
  40. 40. Jennings S, Greenstreet SPR, Reynolds JD (1999) Structural change in an exploited fish community: a consequence of differential fishing effects on species with contrasting life histories. J Anim Ecol 68: 617–627.
  41. 41. Cortés E (2000) Life-history patterns and correlations in sharks. Rev Fish Sci (8): 299–344.
  42. 42. Cope JM (2006) Exploring intraspecific life history patterns in sharks. Fish Bull 104: 311–320.
  43. 43. Frisk MG, Miller TJ, Fogarty MJ (2001) Estimation and analysis of biological parameters in elasmobranch fishes: a comparative life history study. Can J Fish Aquat Sci 58: 969–981.
  44. 44. Nelson J (1984) Fishes of the World. Wiley, New York.
  45. 45. Driggers WB, Carlson JK, Cullum B, Dean JM, Oakley D, Ulrich G (2004) Age and growth of the blacknose shark, Carcharhinus acronotus, in the western North Atlantic Ocean with comments on regional variation in growth rates. Environ Biol Fishes 71: 171–178
  46. 46. Schwartz FJ (1984) Occurrence, abundance, and biology of the blacknose shark, Carcharhinus acronotus, in North Carolina. Northeast Gulf Sci 7: 29–47.
  47. 47. Carlson JK, Cortés E, Johnson AG (1999) Age and growth of the blacknose shark, Carcharhinus acronotus, in the Eastern Gulf of Mexico. Copeia 1999: 684–691.
  48. 48. Huang JC (2006) Age and growth of the blue shark, Prionace glauca in the Northwest Pacific. M Sc. Thesis, National Taiwan Ocean University, Keelung, Taiwan.
  49. 49. Chang JH, Liu KM (2009) Stock assessment of the shortfin mako shark, Isurus oxyrinchus, in the Northwest Pacific Ocean using per-recruit and virtual population analyses. Fish Res 98: 92–103.
  50. 50. Taylor CC (1958) Cod growth and temperature. Journal du Conseil International pour L’exploration de la Mer 23: 366–370.
  51. 51. Hoenig JM (1983) Empirical use of longevity data to estimate mortality rates. Fish Bull 82: 898–903.
  52. 52. Krebs CJ (1985) Ecology: the experimental analysis of distribution and abundance, 3rded. Harper and Row, New York, NY, 800pp.
  53. 53. Jensen AL (1996) Beverton and Holt life history invariants result from optimal trade-off of reproduction and survival. Can J Fish Aquat Sci 53(4): 820–822.
  54. 54. Campana S, Joyce, W, Marks, L, Harley, S (2001) Analytical assessment of the porbeagle shark (Lamna nasus) population in the northwest Atlantic, with estimates of long-term sustainable yield. Canadian Stock Assessment Research Document 2001/067, Ottawa, Ontario.
  55. 55. Quinn TJ II, Deriso RB (1999) Quantitative fish dynamics. Oxford University Press, New York, NY, 542pp.
  56. 56. Wooldridge JM (2009) Introductory econometrics: A modern approach. 5th edition, South-Western, Mason, OH, 881pp.
  57. 57. Simpfendorfer CA, Chidlow J, McAuley R, Unsworth P (2000) Age and growth of the whiskery shark, Furgaleus macki, from southwestern Australia. Environ Biol Fishes 58: 335–343.
  58. 58. Compagno LJV (1984) Sharks of the world. An annotated and illustrated catalogue of shark species known to data. FAO Species Catalogue. Vol. 4, Parts 1 and 2. FAO Fish. Synopsis 125. FAO, Rome, Italy.
  59. 59. Casey JG, Natanson LJ (1992) Revised estimates of age and growth of the sandbar shark (Carcharhinus plumbeus) from the western North Atlantic. Can J Fish Aquat Sci 49: 1474–1477.
  60. 60. Carlson JK, Baremore IE (2003) Changes in biological parameters of Atlantic sharpnose shark Rhizoprionodon terraenovae in the Gulf of Mexico: evidence for density-dependent growth and maturity. Mar Freshwater Res 54: 227–234.
  61. 61. Branstetter S, Stiles R (1987) Age and growth estimates of the bull shark, Carcharhinus leucas, from the northern Gulf of Mexico. Environ Biol Fishes 20(3): 169–181.
  62. 62. Joung SJ, Liao YY, Liu KM, Chen CT, Leu LC (2005) Age, growth, and reproduction of the spinner shark, Carcharhinus brevipinna, in the northeastern waters of Taiwan. Zool Stud 44(1): 102–110.
  63. 63. Jensen CF, Natanson LJ, Pratt HL, Kohler NE, Campana SE (2002). The reproductive biology of the porbeagle shark (Lamna nasus) in the western North Atlantic Ocean. Fish Bull 100: 727–738.
  64. 64. Stevens JD, Wiley PD (1986) Biology of two commercially important Carcharhinid sharks from Northern Australia. Aust J Mar Freshwater Res 37: 671–688.
  65. 65. Peres MB, Vooren CM (1991) Sexual development, reproductive cycle, and fecundity of the school shark Galeorhinus galeus off Southern Brazil. Fish Bull 89: 655–667.
  66. 66. Branstetter S, Musick JA, Colvocoresses JA (1987) A comparison of the age and growth of the tiger shark, Galeocerdo cuvier, from off Virginia and from the northwestern gulf of Mexico. Fish Bull 85(2): 269–279.
  67. 67. Thorson JT, Cope JM, Kleisner KM, Samhouri JF, Shelton AO, Ward EJ (2015) Giants' shoulders 15 years later: lessons, challenges and guidelines in fisheries meta-analysis. Fish and Fisheries 16: 342–361.
  68. 68. Branstetter S (1987) Age, growth and reproductive biology of the silky shark, Carcharhinus falciformis, and the scalloped hammerhead, Sphyrna lewini, from the northwestern Gulf of Mexico. Environ Biol Fishes 19(3): 161–173.
  69. 69. Chen CT, Leu TC, Joung SJ, Lo NCH (1990) Age and growth of the scalloped hammerhead Sphyrna lewini in northeastern Taiwan waters. Pac Sci 44(2): 156–170.
  70. 70. Anislado-Tolentino V, Gallardo-Cabello M, Amezcua-Linares F, Robinson-Mendoza C (2008) Age and growth of the scalloped hammerhead shark, Sphyrna lewini (Griffith & Smith, 1834) from the Southern coast of Sinaloa, México. Hidrobiol 18(1): 31–40.
  71. 71. Kotas JE, Mastrochirico VB, Petrere MC Jr. (2011) Age and growth of the scalloped hammerhead shark, Sphyrna lewini (Griffith and Smith, 1834), from the southern Brazilian coast. Brazil J Biol 71(3): 755–761.
  72. 72. Pratt HL, Casey JG (1983) Age and growth of the shortfin mako, Isurus oxyrinchus, using four methods. 40(11): 1944–1957.
  73. 73. Cailliet GM, Martin LK, Harvey JT, Kusher D, Welden BA (1983) Preliminary studies on the age and growth of blue, Prionace glauca, common thresher, Alopias vulpinus, and shortfin mako, Isurus oxyrinchus, sharks from California waters. NOAA Technical Reports, NMFS 8: 179–188.
  74. 74. Neer JA, Thompson BA, Carlson JK (2005) Age and growth of Carcharhinus leucas in the northern Gulf of Mexico: incorporating variability in size at birth. J Fish Biol 67: 370–383.
  75. 75. Branstetter S (1990) Early life-history implications of selected Carcharhinoid and Lamnoid sharks of the northwest Atlantic. In: Pratt HL, Gruber SH, Taniuchi T, editors. Elasmobranchs as living resources: advances in the biology, ecology, systematics, and the status of the fisheries. NOAA Technical Reports, NMFS 90: 17–28.
  76. 76. Simpfendorfer CA, McAuley RB, Chidlow J, Unsworth P (2002) Validated age and growth of the dusky shark, Carcharhinus obscurus, from Western Australian waters. Mar Freshwater Res 53: 567–573.
  77. 77. Natanson LJ, Mello JJ, Campana SE (2002) Validated age and growth of the probeagle shark, Lamna nasus, in the western North Atlantic Ocean. Collec Vol Sci Pap ICCAT 54 (4): 1261–1279.
  78. 78. Santana FM, Lessa R (2004) Age determination and growth of the night shark (Carcharhinus signatus) off the northeastern Brazilian coast. Fish Bull 102:156–167.
  79. 79. Chen WK, Chen PC, Liu KM, Wang SB (2007) Age and growth estimates of the whitespotted bamboo shark, Chiloscyllium plagiosum, in the northern waters of Taiwan. Zool Stud 46(1): 92–102.
  80. 80. Wourms JP (1977). Reproduction and development in chondrichthyan fishes. Am Zool 17: 379–410.
  81. 81. West JG, Carter S (1990) Observations on the development and growth of the epaulette shark Hemiscyllium ocellatum (Bonnaterre) in captivity. J. Aquaricul Aquat Sci 4: 111–117.
  82. 82. Branstetter S (1987) Age, growth and reproductive biology of the silky shark Carcharhinus falciformis and the scalloped hammerhead, Sphyrna lewini, from the northwestern Gulf of Mexico. Environ Biol Fishes 19 (3): 161–174.
  83. 83. Bonfil R, Mena R, de Anda D (1993) Biological parameters of commercially exploited silky sharks, Carcharhinus falcifomis, from the Campeche Bank, Mexico. NOAA Technical Reports, NMFS 115: 73–86.
  84. 84. Joung SJ (1993). Biology of the sandbar shark, Carcharhinus plumbeus, in northeastern waters of Taiwan. Ph. D. Dissertation, National Taiwan Ocean University Keelung, Taiwan.
  85. 85. Pratt HLJ, Casey JG (1990) Shark reproductive strategies as a limiting factor in directed fisheries, with a review of Holden's method of estimating growth-parameters. In: Pratt HL, Gruber SH, Taniuchi T, editors. Elasmobranchs as living resources: advances in the biology, ecology, systematics, and the status of the fisheries. NOAA Technical Reports, NMFS 90: 98–129.
  86. 86. Skomal GB, Natanson LJ (2003) Age and growth of the blue shark (Prionace glauca) in the North Atlantic Ocean. Fish Bull 101: 627–639.
  87. 87. Fabens AJ (1965) Properties and fitting of the von Bertalanffy growth curve. Growth 29: 265–289. pmid:5865688
  88. 88. Pauly D (1984). Fish population dynamics in tropical waters. ICLARM Studies and Reviews 8: 325pp.
  89. 89. Froese R, Binohlan C (2000) Empirical relationships to estimate asymptotic length, length at first maturity and length at maximum yield per recruit in fishes, with a simple method to evaluate length frequency data. J Fish Biol 56: 578–773.
  90. 90. Oshitani S, Nakano H, Tanaka S (2003) Age and growth of silky shark Carcharhinus falciformis from the Pacific Ocean. Fish Sci 69: 456–464.
  91. 91. Pauly D (1980) On the interrelationships between natural mortality, growth parameters, and mean environmental temperature in 175 fish stocks. ICES J Mar Sci 39(2): 175.
  92. 92. Peterson I, Wroblewski JS (1984) Mortality rate of fishes in the pelagic ecosystem. Can J Fish Aquat Sci 41(7): 1117–1120.
  93. 93. Chen S, Watanabe S (1989) Age dependence of natural mortality coefficient in fish population dynamics. Nippon Suisan Gakkaishi 51: 205–208.
  94. 94. Hoenig JM, Gruber SH (1990). Life-history patterns in the elasmobranchs: implications for fisheries management. In: Pratt HL, Gruber SH, Taniuchi T, editors. Elasmobranchs as living resources: advances in the biology, ecology, systematics, and the status of the fisheries. NOAA Technical Reports, NMFS 90: 1–16.
  95. 95. Then AY, Hoenig JM, Hall NG, Hewitt DA (2014) Evaluating the predictive performance of empirical estimators of natural mortality rate using information on over 200 fish species. ICES J Mar Sci 72: 82–92.
  96. 96. Kleiber P, Clarke S, Bigelow K, Nakano H, McAllister M, Takeuchi Y (2009) North Pacific Blue Shark Stock Assessment. U.S. Dep. Commer., NOAA
  97. 97. Tech. Memo., NOAA-TM-NMFS-PIFSC-17, 74 p.
  98. 98. Jiao Y, Hayes C, Cortés E (2009) Hierarchical Bayesian approach for population dynamics modelling of fish complexes without species-specific data. ICES J Mar Sci 66: 367–377.
  99. 99. Babcock EA, Cortés E (2009) Updated Bayesian surplus production model applied to blue and mako shark catch, CPUE and effort data. Collec Vol Sci Pap ICCAT 64(5): 1568–1577.
  100. 100. ISC Shark Working Group (2013) Stock Assessment and Future Projections of Blue Shark in the North Pacific Ocean. 82 pp.
  101. 101. Walker TI (1998) Can shark resources be harvested sustainably? A question revisited with a review, of shark fisheries. Mar Freshw Res 49(7): 553–572.
  102. 102. Francis MP (1998) New Zealand shark fisheries: development, size and management. Mar Freshwater Res 49 (7): 579–591.
  103. 103. Tsai WP, Liu KM, Joung SJ (2010) Demographic analysis of the pelagic thresher shark, Alopias pelagicus, in the north-western Pacific using a stochastic stage-based model. Mar Freshw Res 61(9): 1056–1066.