Skip to content
Publicly Available Published by De Gruyter February 21, 2019

Wolf diet and livestock selection in central Greece

  • Maria Petridou EMAIL logo , Dionisios Youlatos , Yorgos Lazarou , Kiriakos Selinides , Charilaos Pylidis , Alexios Giannakopoulos , Vassiliki Kati and Yorgos Iliopoulos
From the journal Mammalia

Abstract

Understanding the feeding habits of wolves is essential for designing and implementing fundamental management processes across the range of the species. This is even more important within human-dominated areas, such as southern Europe, and more especially Greece. In this context, we analyzed 123 scat samples, collected between 2010 and 2012, from a mixed agricultural, forested and human-dominated area, centered on the municipality of Domokos in central continental Greece. We used standard laboratory procedures for scat analysis and calculated percentages of frequency of occurrence (FO%), average volume (AV%) and biomass index (BM%) to assess diet composition, and estimated prey selectivity. Domestic prey composed the bulk of wolf diet (FO%=73.5, AV%=84.8, BM%=97.2), wild ungulates were almost absent (FO%=0.5, AV%=0.8, BM%=1.2), whereas grass consumption was high in our area (FO%=19.5, AV%=11.0). The high dependence on livestock corroborates previous studies from Greece and other countries in southern Europe. Goat (FO%=46.0, AV%=61.2, BM%=64.9) was the main prey and was strongly selected, with sheep (FO%=11.5, AV%=9.0, BM%=11.2), pig carrion and cattle ranking behind (FO%=11.5, AV%=10.1, BM%=8.7 and FO%=4.5, AV%=4.5, BM%=12.4, respectively). No differences across seasons were detected, except from pig carrion, which increased during winter. The preference for goats is probably associated with its grazing behavior. High livestock consumption generally results in increased human-wolf conflict. Thus, substantial improvement of husbandry practices and restoration of wild ungulate populations are recommended to facilitate wolf-human coexistence in Greece.

Introduction

In southern Greece, the gray wolf (Canis lupus Linnaeus, 1758) became extinct in the late 1930s from Peloponnesus while its distribution in continental Greece shrank considerably following several decades of persecution, bounties and legal use of poison baits (Iliopoulos 2010). During the 1980s–1990s, stricter legal status reversed the wolf population decline and its range expanded, mainly in southern central Greece (Iliopoulos 2010). Expansion since 2005 involved regions in the prefectures of Boeotia and Attica (Iliopoulos et al. 2015). Recolonization of such new areas, especially in human-dominated landscapes, may lead to new conflicts, as wolves may have an impact on people and their resources, such as livestock and game (Chapron et al. 2014, Carter and Linnell 2016, Newsome et al. 2016). Therefore, to assess this impact on wild and domestic prey populations, it is crucial to understand wolf diet in these areas.

Wolf food habits have been widely studied throughout its European range. In central and northeastern Europe, wolves rely mainly on wild ungulates, such as wild boar (Sus scrofa Linnaeus, 1758), roe deer (Capreolus capreolus Linnaeus, 1758) and red deer (Cervus elaphus Linnaeus, 1758) (Kojola et al. 2004, Nowak et al. 2011, Jedrzejewski et al. 2012, Lanszki et al. 2012, Wagner et al. 2012). On the other hand, in southern Europe, wolves depend heavily on livestock, and more especially in livestock farming areas and in periods of high livestock production (Meriggi et al. 1991, Blanco et al. 1992, Vos 2000, Migli et al. 2005, Gazzola et al. 2008, Álvares 2011, Llaneza and Lopez-Bao 2015, Torres et al. 2015, Ciucci et al. 2018).

Livestock predation represents one of the main reasons of wolf-human conflicts, and it is directly related to extensive killing of wolves (Meriggi and Lovari 1996, Alvares 2011). In Greece, the annual compensation is €1,098,000±243,000 standard deviation (SD) corresponding to 2198±604 SD attacks attributed to wolves (2006–2012, Hellenic Farmers Insurance Organization – ELGA – data), and illegal human-caused mortality remains high and may locally reach up to 25% of estimated wolf numbers (Iliopoulos et al. 2015). In order to mitigate the conflict between wolves and local communities and to design concrete conservation and restoration strategies, it is critical to have concrete data on this predator’s diet (Meriggi et al. 2011, Newsome et al. 2016).

Despite the abundance of studies on wolf diet in Europe, similar research is scarce in human-dominated areas (Newsome et al. 2016). This is also the case for Greece, where wolves prey heavily on livestock (Papageorgiou et al. 1994, Migli et al. 2005, Iliopoulos et al. 2009). In order to increase our knowledge on the feeding ecology of the gray wolf, we analyzed the diet of the species in central Greece from 2010 to 2012 via scat analysis. The current study was conducted in a highly anthropogenic area, where livestock density is high and wild ungulate density and diversity are low. The present study aims to (a) investigate if the wolf diet comprises mainly livestock and (b) examine if the wolf shows any preference toward specific livestock species. We examined the hypothesis that wolves would primarily feed on livestock and on the most vulnerable livestock species according to grazing patterns rather than seasonal availability. We expected that the most vulnerable livestock species would be goats because of their tendency to graze in a scattered manner and in more inaccessible and dense areas (Iliopoulos et al. 2009, Torres et al. 2015). Such data are of critical importance for the management of wolf-human conflicts.

Materials and methods

Study area

The study area extended over approximately 612 km2 of agricultural, forested and human-dominated areas roughly centered on the municipality of Domokos and extending across the prefectures of Fthiotida and Karditsa in central continental Greece (Figure 1). The area has a typical Mediterranean climate with mild winters and warm summers and almost half of all precipitation occurring during winter (the mean annual precipitation is 554.9 mm, the mean annual temperature is 13.9°C, Worldclim Database 2.0; Fick and Hijmans 2017). Elevation ranges from 74 to 1064 m a.s.l. The landscape is dominated by irrigated agricultural land at lower elevations (40%), by evergreen shrub land of Quercus coccifera (35%) and by deciduous oak forests of Quercus spp. (25%). The mean road density is 2.93±1.15 km/km2 including paved and dirt (agricultural/forestry) roads. A large number of human settlements are scattered throughout the landscape, ranging from large villages and towns to small settlements of a few farmhouses. The human population density is 27.9 inhabitants/km2 excluding large cities.

Figure 1: The study site in central continental Greece, indicating the location of wolf scats, human settlements and the garbage dump.
Figure 1:

The study site in central continental Greece, indicating the location of wolf scats, human settlements and the garbage dump.

Extensive free ranging livestock raising (sheep, goat and cattle) is widespread in the study area and the majority of livestock grazes year round with varying seasonal intensity. We collected data on livestock availability by interviewing all livestock holdings in the area (n=266). During the winter grazing period (mid Oct–end April), 14,000 goats, 26,000 sheep and 1950 cattle graze in the area. During the summer period (May–Oct), these numbers drop to 12,300 goats, 23,500 sheep and 1750 cattle, as some herders move to higher altitude pastures. Sheep herd size averaged 148 animals (range 10–1200), goat herd size averaged 114 animals (range 10–800) and cattle herd size averaged 103 animals (range 5–400). The number of livestock guardian dogs per herd averaged 3.8 animals (range 0–15) with a total estimated number of 865 guardian dogs present in the study area. One pig farm of 100 fenced pigs was located in the center of the study area which provided a nearby open dumping site with pig remains accessible to wildlife and dogs. Goats and sheep graze only during daylight guarded by a shepherd and guardian dogs, while during the night they are kept inside folds. Sheep typically graze in forest openings, abandoned fields and cultivated land forming more compact large herds. On the other hand, goats tend to graze mostly inside dense scrubland in a more scattered manner with large herds often breaking into several smaller groups. Cattle and calves graze mainly unguarded, both day and night, without or with loose supervision.

With the exception of five game reserves, covering a surface of 61 km2, hunting is allowed throughout the region. In addition to wolves, there is a diverse mesocarnivore community consisting of the wildcat (Felis silvestris silvestris Schreber, 1777), European badger (Meles meles Linnaeus, 1758), stone marten (Martes foina Erxleben, 1777), weasel (Mustela nivalis Linnaeus, 1766) and fox (Vulpes vulpes Linnaeus, 1758). Wild boar and roe deer were found in relatively low densities in the area as revealed by the relative abundance index (RAI), i.e. frequency of animal detections (species detections per 100 nights of camera trap sampling). The RAI index was calculated for wild and domestic ungulates as follows: wild boar RAI=3.43, roe deer RAI=0.83, livestock RAI=22.67, based on 45 camera traps and 4225 trap nights with a time interval detection set at 30 min.

Based on data from genetic analyses, camera trapping and snow tracking (Iliopoulos et al. 2010, 2013), the wolf population of the area was estimated to be at a minimum of 15 animals, belonging to two wolf packs (wolf density approximately 2.4 individuals/100 km2). During the study, three wolves (two females, one male) from two packs were live-captured and equipped with GPS-GSM collars (Global Position System-Global System for Mobile Communication, Vectronic Aerospace, Berlin, Germany) in order to study wolf movements and habitat use in relation to large linear infrastructures in the area. The visitation rate of the one wolf pack at the dumping site was 14 visits per month during the summer period (n=103 days) and 12.5 visits per month during the winter period (n=118 days).

Scat collection and laboratory procedures

We collected wolf scats between January 2010 and January 2012, by examining frequently visited wolf areas and travel routes, as indicated by soil, mud and snow tracks, camera trapping and wolf GPS relocation data. Due to the high density of guardian dogs, we collected scats cautiously by evaluating scat characteristics (size, smell, shape and texture; Ciucci et al. 1996). In case of persistent uncertainty, the scat was not collected. Epithelial samples from the surface of fresh wolf scats were also taken for genetic analysis. We used the QIAmp DNA Stool Mini Kit (QIAGEN, Manchester, UK) for DNA extraction. The eluted DNA was amplified in a multiplex reaction with the Caniplex system for dog and wolf identification (FH3210, FH3241, FH2004, FH2658, FH4012, REN214L11, FH2010, FH236, C38) (van Asch et al. 2010) through polymerase chain reaction (PCR). We followed the multitube approach with three replications per sample. Replicate genotypes were pooled manually to create consensus profiles and the probability of identity (PID) exact match function in CERVUS was used to determine matches between independent profiles. The Bayesian clustering algorithm in STRUCTURE software (Pritchard et al. 2000) was used to assign profiles into species.

Collected scats were placed in plastic bags, coded and stored in a freezer at −30°C. For each scat, we recorded the date and location of collection. Laboratory procedures for diet analysis followed Bassi et al. (2012). Initially, scats were diluted into water and washed through a sieve with 0.5 mm mesh size to separate the macro-components and, then, each scat was oven-dried separately at 65°C for 24 h. Subsequently, we separated all elements belonging to different food categories, such as plant matter (stems, leaves, seeds), animal matter (hair, bones, skin, feathers) and garbage. The proportional volume of each food category was visually estimated using a reference grid.

Species identification of mammalian food items was based on macroscopic (color, length, texture) and microscopic (medulla and cuticular pattern) characteristics of hair, using the authors’ reference hair collection of domestic and wild Greek mammals, and the aid of published manuals (Debrot et al. 1982, Teerink 1991). Ability of the authors to identify prey species from hair was tested by means of a blind test performed on 50 artificial scats containing 13 mammal prey items with different combinations (Bassi et al. 2012). Hair identification in scats was performed by one of the authors (MP), who scored 100% in the test.

Diet composition analysis

For each food item, we calculated: (i) % frequency of occurrence (FO%= ni/N×100, where ni is the number of scats containing a given item i and N is the total number of scats), and (ii) % average volume (AV%=Vi/N×100, where Vi is the total volume of a given food item i, and N is the total number of scats), as in other studies (e.g. Klare et al. 2011, Bassi et al. 2012). We considered as “trace” food items those constituting <5% of scat volume and we excluded them from further analysis, as suggested by Klare et al. (2011). We subsequently calculated % biomass index (BM%) for each mammalian food item i using the standard formula BMi%=(Yi×AVi%/sum Yij×AVij%)×100, where Yi is Weavers’ correction factor (Weaver 1993), for estimating kilograms of each mammalian food item i per scat (based on the linear regression model Yi=0439+0.008Xi, Xi being mean mammalian weight in kilograms). We used mammalian mean weights reported in the literature (Rogdakis 2006, Barja 2009): goat 40 kg, sheep 57 kg, cattle (Limousin breed) 180 kg, pig 95 kg, wild boar 75 kg, fox 8.1 kg, badger 12 kg, cat 4.3 kg, hare 3.8 kg, small rodents 0.02 kg. For livestock, we used female mean weight, because herds mainly consisted of females. For cattle, we used weights corresponding to the 4–7-month age class, which is more susceptible to wolf attacks, as assessed by field work and interviews with cattle breeders. For wild prey, we used adult mean weight per species. To estimate the BM%, we used a 90% proportion of each species’ mean weight, with the exception of rodents, representative of the average large mammal consumption rate by wolves when excluding non-consumptive parts (most of skeleton, horns, paunch contents, hooves, etc.; Jedrzejewski et al. 2002, Álvares 2011). For domestic pig, we used instead the 20% proportion of the mean pig weight, because pigs were available only as carrion parts (skin, heads, etc.) at the dumping site.

We used a t-test to compare seasonal differences (summer=May–Oct, winter=Nov–Apr) on AV%, per food item. We used the Holm method to control p-valued multiple comparison errors (Holm 1979). Residuals were tested for normality. Comparisons were performed only for food items which appeared more than 10 times in total.

Prey selection considering livestock was calculated with Ivlev’s selectivity index D (Jacobs 1974) for both seasons:

D=rpr+p2rp

where r is the proportion of occurrence of each livestock species in the scat sample, and p is the proportion of their numeric availability in the study area. Ivlev’s selectivity index ranges from −1 (strong avoidance), through 0 (selection according to availability) to +1 (strong preference).

Results

We successfully analyzed 123 wolf scats (76 winter samples and 47 summer samples) corresponding to the January 2010–January 2012 period, after excluding amorphous (n=13) and probable dog scats (n=11). Successful genetic analysis on 41 scats resulted in a high rate of correctly identified wolf scats (88%). A total number of 41 profiles were produced which corresponded to 25 individual consensus profiles. STRUCTURE assigned the profiles into 20 wolves and five dogs at K=2 with a clear intraspecific population distinction for wolves. The average membership coefficient (Q) was 0.986 for wolves and 0.996 for dogs whilst no hybridization was detected.

Overall, our scat sample contained 11 food items: 74% contained one food item, 19% contained two, 4% contained three and only 1% of scats contained four items. Domestic ungulates composed the bulk of wolf diet (FO%=73.5%, AV%=84.8%, BM%=97.2%). Wild ungulates were nearly absent (FO%=0.5%, AV%=0.8%, BM%=1.2%), whereas other smaller prey represented only a low proportion (FO%=4%, AV%=2.6%, BM%=1.7%). Plant matter, mainly grass, was often consumed (FO%=19.5%, AV%=11.0%). Results are summarized in Table 1.

Table 1:

Composition of wolf diet in central Greece based on scat analysis (n=123, January 2010–January 2012).

Food itemTotal (n=123)Winter (n=76)Summer (n=47)
FO%AV%BM%FO%AV%BM%FO%AV%BM%
Domestic ungulates
 Goat46.061.264.945.258.362.547.165.868.4
 Sheep11.59.011.212.29.111.410.68.910.8
 Cattle4.54.512.43.54.512.45.94.612.5
 Pig (carrion)11.510.18.713.013.711.99.44.33.6
73.584.897.273.985.598.272.983.695.3
Wild ungulates
 Wild boar0.50.81.20.00.00.01.22.13.0
 Roe deer0.00.00.00.00.00.00.00.00.0
0.50.81.20.00.00.01.22.13.0
Other prey
 Hare0.50.20.20.00.00.01.20.60.6
 Badger0.50.40.30.90.70.50.00.00.0
 Domestic cat0.50.60.40.00.00.01.21.61.1
 Fox0.50.70.50.91.10.80.00.00.0
 Small rodents1.00.50.31.70.80.50.00.00.0
 Birds1.00.20.90.11.20.3
4.02.61.74.42.61.83.52.51.7
Plant material
 Grasses18.510.819.111.117.610.3
 Seeds1.00.20.90.21.20.2
19.511.020.011.318.810.5
 Garbage2.50.81.70.53.51.2
Total100.0100.0100.0100.0100.0100.0100.0100.0100.0
  1. Data are expressed as frequency of occurrence (FO%), average volume (AV%) and biomass (BM%) using the linear regression model by Weaver (1993).

Goat was the most important prey species (FO%=46.0%, AV%=61.2%, BM%=64.9%). The rest of livestock species were also frequently consumed, with sheep (FO%=11.5%, AV%=9.0%), pig (FO%=11.5%, AV%=10.1%) and cattle (FO%=4.5%, AV%=4.5%) ranking behind (Table 1). In contrast, based on BM%, cattle (BM%=12.4%) consisted the second most important prey closely followed by sheep (BM%=11.2%) and domestic pig (BM%=8.7%). Undigested garbage remains (e.g. plastic and paper) were not very common (FO%=2.5%, AV%=0.8%). However, when domestic pig (available only as carrion parts in the dumping site) was added to the garbage category, the FO% proportion was raised to 14%.

We did not detect any seasonal differences in overall livestock consumption rates (AV%, t-test: F=3.62, p=0.118; Figure 2). However, when considering livestock species separately, only the consumption of pig carrion parts significantly increased during winter (t-test: F=12.99, p=0.007; Figure 2).

Figure 2: Seasonal differences in wolf diet composition between winter (n=76) and summer (n=47) based on average volume (AV%) in central Greece.
Figure 2:

Seasonal differences in wolf diet composition between winter (n=76) and summer (n=47) based on average volume (AV%) in central Greece.

Regarding livestock selection, goats were strongly selected by wolves during both summer and winter (summer: D=0.71, winter: D=0.70; Figure 3). Cattle were moderately selected in both seasons, but to a greater extent during summer (winter: D=0.12, summer: D=0.35; Figure 3). Lastly, sheep were strongly avoided in both seasons (winter D=−0.73, summer D=−0.79; Figure 3). Wild prey was rarely consumed. Wild boar was found only in one scat (FO%=0.5%, AV%=0.8%, BM=1.2%) and roe deer was totally absent. Other prey (e.g. hare, mesocarnivores, small rodents, birds) was also rare (FO%=0.5–1.0%, AV%=0.2–0.7%, BM%=0.2–0.5%; Table 1).

Figure 3: Wolf livestock prey selection in central Greece during summer and winter (Ivlev’s index D).
Figure 3:

Wolf livestock prey selection in central Greece during summer and winter (Ivlev’s index D).

Discussion

The current study provides important information on wolf food habits in a human-dominated landscape, with high availability of livestock, a large number of guardian dogs and low densities of wild ungulates. The large number of guardian dogs compelled us to be extremely cautious and selective during scat collection and subsequent genetic analyses gratified our precautions, as correct identification of wolf scats reached 88%. This further underlines the difficulties in collecting wolf scats in highly anthropogenic environments, rendering the present study quite valuable for assessing wolf diet in human-dominated areas. The present results confirm our prediction that wolf diet in central Greece would be dominated by domestic ungulates, comprising up to 85% of volume (rising up to 95.8% when excluding plant material) and 97.2% of biomass. Moreover, our results confirmed the hypothesis that wolves do not simply prey on the most numerous livestock species but rather the most vulnerable.

Previous studies have also reported the use of domestic ungulates as an important food resource by wolves. This is quite evident in Italy (Meriggi et al. 1996, 2015, Imbert et al. 2016, Ciucci et al. 2018) and Spain (Nores et al. 2008, Lagos and Barcena 2018). However, comparable high dependence on livestock has been only recorded in northwestern Spain (Llaneza and Lopez-Bao 2015), Portugal (Vos 2000, Álvares 2011, Torres et al. 2015) and Greece (Papageorgiou et al. 1994, Migli et al. 2005).

In our study area, wolf diet consisted almost exclusively of four species of domestic ungulates: goat, sheep, cattle and pig (as carrion in the dumping site). Among them, goats were strongly and positively selected all year round, as was also the case in Portugal (Vos 2000, Torres et al. 2015). These results also corroborate previous findings in Greece, revealing goats as the main prey of wolves (Papageorgiou et al. 1994, Iliopoulos et al. 2009), or an important one (Migli et al. 2005); in this study, domestic pigs scored higher but grazed unguarded throughout the day). Although goats were not the most available livestock species, wolf preference for goats is very likely related to their easier accessibility. Goats tend to utilize denser forested areas and steeper terrains, which are favored by wolves for resting (Ciucci et al. 1997, Llaneza et al. 2016) and home sites (Ciucci et al. 1997, Iliopoulos et al. 2014, Sazatornil et al. 2016), increasing wolf-livestock encounter rates. Moreover, goat surveillance is harder because they tend to scatter extensively, providing wolves with more opportunities to approach and attack (Iliopoulos et al. 2009).

Sheep also represented a main domestic prey category (Table 1). However, our findings showed limited selection by wolves, despite their high availability. Limited selection of sheep has been previously recorded in Greece (Iliopoulos et al. 2009), as well as in Portugal (Vos 2000, Torres et al. 2015). Only, in central Italy, sheep was the preferred livestock species, but there, goat availability was very low (Ciucci and Boitani 1998). Limited selection of sheep may be related to the fact that they graze in open pasture areas and tend to group in dense and compact flocks, rendering them a difficult target (see also Torres et al. 2015).

Cattle was also a significant prey for wolves (especially when considering BM%) and were moderately selected. This finding is consistent with the results from other southern European countries, such as Portugal (Pimenta et al. 2017), Spain (Llaneza and Lopez-Bao 2015) or Italy (Ciucci et al. 2018) and is important considering the high economic value of cattle. Their selection was probably related to the fact that cattle, including calves, grazed unattended during day and night. Moreover, their relatively higher consumption during summer may be due to their increased availability during the warm grazing season (Meriggi et al. 1991, Ciucci et al. 2018).

Pig carrion and garbage were well-represented in our sample. These findings corroborate our GPS telemetry data, which revealed that wolves were feeding at the dumping site once every 2 days during both seasons. Wolves feeding in dumping sites have been also documented in Spain and Italy (Ciucci et al. 1997, Lagos and Barcena 2018). Dumping sites, usually located near villages and towns, represent a stable and easily accessible food source that is valuable for wolves when forested areas do not provide enough prey (Ciucci et al. 1997). Moreover, use of dumping sites was apparently higher during winter, as revealed by the increased consumption of pig carrion during winter compared to that of summer (Table 1). This seasonal increase may be linked to the fact that livestock is often kept indoors during the winter and/or to the actual higher winter availability of pig carrion. Carrion availability and consumption may have a significant impact on prey selection and depredation by wolves; European Union (EU) sanitary regulations on livestock carcass disposal in Spain resulted in increased wild ungulate predation, together with an increase in wild ungulate populations (Lagos and Barcena 2015). Permitting wolf access to carrion where and when wild prey is scarce, so as to mitigate wolf-livestock conflict, could be a highly controversial issue. Under low wild ungulate abundance, closure of local dumping sites should be considered very cautiously as this could result in an increase of livestock depredation (Newsome and van Eeden 2017). On the other hand, freely accessible livestock carcasses may attract wolves close to farms resulting in livestock depredation (Lagos and Barcena 2015). In any case, this should be considered for future research and under specific local contexts.

Our results demonstrated that wolves in central Greece largely depend on livestock year round. In general, high occurrence of livestock consumption has been related to low densities and decreased availability of wild ungulates (Sidorovich et al. 2003, Meriggi et al. 2011, Imbert et al. 2016), increased accessibility to livestock (Miller et al. 2016) and lack of effective preventive measures (Miller et al. 2016, Eklund et al. 2017). It is very likely that the low presence of wild ungulates (roe deer and wild boar) in our study site may also account for the high livestock consumption we observed. However, more data on wild ungulate diversity, range, density and population trends will provide more conclusive associations. Apart from that, livestock selection was related to increased accessibility rather than numerical availability. In effect, livestock accessibility depends on the preventive measures applied on husbandry practices and can be notably decreased by introducing effective protection techniques (Miller et al. 2016, Eklund et al. 2017). Well-trained guardian dogs, increased supervision, night fencing, visual and auditory deterrents, synchronized livestock births and exclusion of births and young vulnerable animals from pastures represent effective tools that can reduce livestock losses (Iliopoulos et al. 2009, Dondina et al. 2015, Miller et al. 2016, Eklund et al. 2017, Pimenta et al. 2017). Moreover, we believe that for goats, the number of guardian dogs per 100 animals should be higher compared to that for sheep. Additionally, goatherd supervision needs to be intensified so as to maintain herd cohesion and proper protection during pasture grazing in dense forested habitats. Current research in Greece focuses on the relation of these measures to livestock depredation. Moreover, the adoption of innovative tools, such as predation risk modeling, will identify high-priority “conflict hotspots” and facilitate implementation of intervention actions (Miller 2015).

Wolves in Greece represent the southernmost part of the European population. Long-term conservation of this marginal sub-population depends on efforts to reduce livestock depredation and on successful restoration of wild ungulate populations. These two goals form a necessary prerequisite to ensure independence of wolves from human-related food sources, such as livestock and garbage, and to resolve serious conflicts with rural communities (Chapron et al. 2014).

Acknowledgments

The authors would like to greatly thank E. Bassi, E. Donaggio and M. Apollonio for their warm hospitality and training on wolf diet analysis, S. Roque for her advice on livestock data handling, B. Vilela for help in the statistical analysis, M. Psaralexi for linguistic help with the manuscript and I. Hatzimichael and Y. Mertzanis for their support throughout the study. The study was carried out in the framework of the project “Mitigating habitat fragmentation by the new high speed railway and E65 highway considering wolf and other large mammals”, ERGOSE, TERNA-AKTOR Domokos Consortium, Argyropoulos SA and Callisto NGO.

References

Álvares, F. 2011. Ecologia e conservação do lobo (Canis lupus, L.) no noroeste de Portugal. PhD Thesis, Universidade de Lisboa, Portugal.Search in Google Scholar

Barja, I. 2009. Prey and prey-age preference by the Iberian wolf Canis lupus signatus in a multiple-prey ecosystem. Wildl. Biol. 15: 147–154.10.2981/07-096Search in Google Scholar

Bassi, E., E. Donaggio, A. Marcon, M. Scandura and M. Apollonio. 2012. Trophic niche overlap and wild ungulate consumption by red fox and wolf in a mountain area in Italy. Mamm. Biol. 77: 369–376.10.1016/j.mambio.2011.12.002Search in Google Scholar

Blanco, J.C., S. Reig and L. de la Cuesta. 1992. Distribution, status and conservation problems of the wolf Canis lupus in Spain. Biol. Conserv. 60: 73–80.10.1016/0006-3207(92)91157-NSearch in Google Scholar

Carter, N.H. and J.D.C. Linnell. 2016. Co-adaptation is key to coexisting with large carnivores. Trends Ecol. Evol. 31: 575–578.10.1016/j.tree.2016.05.006Search in Google Scholar PubMed

Chapron, G., P. Kaczensky, J.D.C. Linnell, M. von Arx, D. Huber, H. Andre, J.V. López-Bao, M. Adamec, F. Álvares, O. Anders, L. Balciauskas, V. Balys, P. Bedo, F. Bego, J.C. Blanco, U. Breitenmoser, H. Brøseth, L. Bufka, R. Bunikyte, P. Ciucci, A. Dutsov, T. Engleder, C. Fuxjäger, C. Groff, K. Holmala, B. Hoxha, Y. Iliopoulos, O. Ionescu, J. Jeremic, K. Jerina, G. Kluth, F. Knauer, I. Kojola, I. Kos, M. Krofel, J. Kubala, S. Kunovac, J. Kusak, M. Kutal, O. Liberg, A. Majic, P. Männil, R. Manz, E. Marboutin, F. Marucco, D. Melovski, K. Mersini, Y. Mertzanis, R.W. Myslajek, S. Nowak, J. Odden, J. Ozolins, G. Palomero, M. Paunovic, J. Persson, H. Potocnik, P.-Y. Quenette, G. Rauer, I. Reinhardt, R. Rigg, A. Ryser, V. Salvatori, T. Skrbinšek, A. Stojanov, J.E. Swenson, L. Szemethy, A. Trajçe, E. Tsingarska-Sedefcheva, M. Vána, R. Veeroja, P. Wabakken, M. Wölfl, S. Wölfl, F. Zimmermann, D. Zlatanova, L. Boitani. 2014. Recovery of large carnivores in Europe’s modern human-dominated landscapes. Science 346: 1517–1519.10.1126/science.1257553Search in Google Scholar PubMed

Ciucci, P. and L. Boitani. 1998. Wolf and dog depredation on livestock in central Italy. Wildl. Soc. Bull. 26: 504–514.Search in Google Scholar

Ciucci, P., L. Boitani, E.R. Pelliccioni, M. Roco and I. Guy. 1996. A comparison of scat-analysis methods to assess the diet of the wolf Canis lupus. Wildl. Biol. 2: 37–48.10.2981/wlb.1996.006Search in Google Scholar

Ciucci, P., L. Boitani, F. Francisci and G. Andreoli. 1997. Home range, activity and movements of a wolf pack in central Italy. J. Zool. 243: 803–819.10.1111/j.1469-7998.1997.tb01977.xSearch in Google Scholar

Ciucci, P., L. Artoni, F. Crispino, E. Tosoni and L. Boitani. 2018. Inter-pack, seasonal and annual variation in prey consumed by wolves in Pollino National Park, southern Italy. Eur. J. Wildl. Res. 64: 5.10.1007/s10344-018-1166-1Search in Google Scholar

Debrot, S., G. Fivaz, C. Mermod and J.M. Weber. 1982. Atlas des poils de mammifères d’Europe. Institut de Zoologie de l’Université de Neuchâtel, Peseux, Switzerland.Search in Google Scholar

Dondina, O., A. Meriggi, V. Dagradi, M. Perversi and P. Milanesi. 2015. Wolf predation on livestock in an area of northern Italy and prediction of damage risk. Ethol. Ecol. Evol. 27: 200–219.10.1080/03949370.2014.916352Search in Google Scholar

Eklund, A., J.V. López-Bao, M. Tourani, G. Chapron and J. Frank. 2017. Limited evidence on the effectiveness of interventions to reduce livestock predation by large carnivores. Sci. Rep. 7: 2097.10.1038/s41598-017-02323-wSearch in Google Scholar PubMed PubMed Central

ELGA. 2011. Animal insurance regulation. Hellenic Farmers Insurance Organization. FEK 1669/B/27-7-2011. Athens: 1–12 (in Greek).Search in Google Scholar

Fick, S.E. and R.J. Hijmans. 2017. WorldClim 2: new 1-km spatial resolution climate surfaces for global land areas. Int. J. Climatol. 37: 4302–4315.10.1002/joc.5086Search in Google Scholar

Gazzola, A., C. Capitani, L. Mattioli and M. Apollonio. 2008. Livestock damage and wolf presence. J. Zool. 274: 261–269.10.1111/j.1469-7998.2007.00381.xSearch in Google Scholar

Holm, S. 1979. A simple sequentially rejective multiple test procedure. Scand. J. Stat. 6: 65–70.Search in Google Scholar

Iliopoulos, Y. 2010. Wolf (Canis lupus) packs territory selection in Central Greece. Habitat selection, movement patterns and effects on livestock. PhD thesis. Department of Zoology, School of Biology, Aristotle University of Thessaloniki: Thessaloniki, Greece.Search in Google Scholar

Iliopoulos, Y., S. Sgardelis, V. Koutis and D. Savaris. 2009. Wolf depredation on livestock in central Greece. Acta Theriol. 54: 11–22.10.1007/BF03193133Search in Google Scholar

Iliopoulos, Y., A. Gannakopoulos, M. Petridou, Y. Lazarou, C. Pylidis and I. Aravidis. 2010. Monitoring the impact on wolves Canis lupus from the construction of the motorway of Central Greece – E65 south part. TERNA-AKTOR Domokos Consortium, Argyropoulos SA, Callisto NGO (in Greek).Search in Google Scholar

Iliopoulos, Y., M. Petridou, K. Selinides, Y. Lazarou, A. Giannakopoulos, C. Pylidis and I. Aravidis. 2013. Wolf monitoring project during the construction phase of high speed railway in Central Greece 2009–2012: proposals for mitigating effects on habitat fragmentation. ERGOSE, Callisto NGO (in Greek).Search in Google Scholar

Iliopoulos, Y., D. Youlatos and S. Sgardelis. 2014. Wolf pack rendezvous site selection in Greece is mainly affected by anthropogenic landscape features. Eur. J. Wildl. Res. 60: 23–34.10.1007/s10344-013-0746-3Search in Google Scholar

Iliopoulos, Y., C. Astaras, M. Petridou and E. Sideri. 2015. Total deliverables for wolf monitoring. In: (C. Papamichail, T. Arapis and K. Petkidis, eds.) Monitoring and assessment of the conservation status of species of mammals of community interest in Greece. YPEKA, Athens.Search in Google Scholar

Imbert, C., R. Caniglia, E. Fabbri, P. Milanesi, E. Randi, M. Serafini, E. Torretta and A. Meriggi. 2016. Why do wolves eat livestock? Factors influencing wolf diet in northern Italy. Biol. Conserv. 195: 156–168.10.1016/j.biocon.2016.01.003Search in Google Scholar

Jacobs, J. 1974. Quantitative measurement of food selection. Oecologia 14: 413–417.10.1007/BF00384581Search in Google Scholar PubMed

Jedrzejewski, W., K. Schmidt, J. Theuerkauf, B. Jedrzejewska, N. Selva, K. Zub and L. Szymura. 2002. Kill rates and predation by wolves on ungulate populations in Białowieża Primeval Forest (Poland). Ecology 83: 1341–1356.10.2307/3071948Search in Google Scholar

Jedrzejewski, W., M. Niedzialkowska, M.W. Hayward, J. Goszczynski, B. Jedrzejewska, T. Borowik, K.A. Barton, S. Nowak, J. Harmuszkiewicz, A. Juszczyk, T. Kalamarz, A. Kloch, J. Koniuch, K. Kotiuk, R.W. Myslajek, M. Nedzynska, A.O. Marta and T.M. Wojtulewicz. 2012. Prey choice and diet of wolves related to ungulate communities and wolf subpopulations in Poland. J. Mammal. 93: 1480–1492.10.1644/10-MAMM-A-132.1Search in Google Scholar

Klare, U., J.F. Kamler and D.W. Macdonald. 2011. A comparison and critique of different scat-analysis methods for determining carnivore diet. Mammal. Rev. 41: 294–312.10.1111/j.1365-2907.2011.00183.xSearch in Google Scholar

Kojola, I., O. Huitu, K. Toppinen, K. Heikura, S. Heikkinen and S. Ronkainen. 2004. Predation on European wild forest reindeer (Rangifer tarandus) by wolves (Canis lupus) in Finland. J. Zool. 263: 229–235.10.1017/S0952836904005084Search in Google Scholar

Lagos, L. and F. Bárcena. 2015. EU sanitary regulation on livestock disposal: implications for the diet of wolves. Environ. Manage. 56: 890–902.10.1007/s00267-015-0571-4Search in Google Scholar PubMed

Lagos, L. and F. Bárcena. 2018. Spatial variability in wolf diet and prey selection in Galicia (NW Spain). Mamm. Res. 63: 1–15.10.1007/s13364-018-0352-6Search in Google Scholar

Lanszki, J., M. Márkus, D. Újváry, Á. Szabó and L. Szemethy. 2012. Diet of wolves Canis lupus returning to Hungary. Acta Theriol. 57: 189–193.10.1007/s13364-011-0063-8Search in Google Scholar PubMed PubMed Central

Llaneza, L. and J.V. López-Bao. 2015. Indirect effects of changes in environmental and agricultural policies on the diet of wolves. Eur. J. Wildl. Res. 61: 895–902.10.1007/s10344-015-0966-9Search in Google Scholar

Llaneza, L., E.J. García, V. Palacios, V. Sazatornil and J.V. López-Bao. 2016. Resting in risky environments: the importance of cover for wolves to cope with exposure risk in human-dominated landscapes. Biodivers. Conserv. 25: 1515–1528.10.1007/s10531-016-1134-6Search in Google Scholar

Meriggi, A. and S. Lovari. 1996. A review of wolf predation in southern Europe: does the wolf prefer wild prey to livestock? J. Appl. Ecol. 33: 1561–1571.10.2307/2404794Search in Google Scholar

Meriggi, A., P. Rosa, A. Brangi and C. Matteucci. 1991. Habitat use and diet of the wolf in northern Italy. Acta Theriol. 36: 141–151.10.4098/AT.arch.91-11Search in Google Scholar

Meriggi, A., A. Brangi, C. Matteucci and O. Sacchi. 1996. The feeding habits of wolves in relation to large prey availability in northern Italy. Ecography 19: 287–295.10.1111/j.1600-0587.1996.tb01256.xSearch in Google Scholar

Meriggi, A., A. Brangi, L. Schenone, D. Signorelli and P. Milanesi. 2011. Changes of wolf (Canis lupus) diet in Italy in relation to the increase of wild ungulate abundance. Ethol. Ecol. Evol. 23: 195–210.10.1080/03949370.2011.577814Search in Google Scholar

Meriggi, A., V. Dagradi, O. Dondina, M. Perversi, P. Milanesi, M. Lombardini, S. Raviglione and A. Repossi. 2015. Short-term responses of wolf feeding habits to changes of wild and domestic ungulate abundance in Northern Italy. Ethol. Ecol. Evol. 27: 389–411.10.1080/03949370.2014.986768Search in Google Scholar

Migli, D., D. Youlatos and Y. Iliopoulos. 2005. Winter food habits of wolves in central Greece. J. Biol. Res. 4: 217–220.Search in Google Scholar

Miller, J.R. 2015. Mapping attack hotspots to mitigate human–carnivore conflict: approaches and applications of spatial predation risk modeling. Biodivers. Conserv. 24: 2887–2911.10.1007/s10531-015-0993-6Search in Google Scholar

Miller, J.R., K.J. Stoner, M.R. Cejtin, T.K. Meyer, A.D. Middleton and O.J. Schmitz. 2016. Effectiveness of contemporary techniques for reducing livestock depredations by large carnivores. Wildl. Soc. Bull. 40: 806–815.10.1002/wsb.720Search in Google Scholar

Newsome, T.M. and L. van Eeden. 2017. The effects of food waste on wildlife and humans. Sustainability 9: 1269.10.3390/su9071269Search in Google Scholar

Newsome, T.M., L. Boitani, G. Chapron, P. Ciucci, C.R. Dickman, J.A. Dellinger, J.V. Lòpez-Bao, R.O. Peterson, C.R. Shores, A.J. Wirsing and W.J. Ripple. 2016. Food habits of the world’s grey wolves. Mammal Rev. 46: 255–269.10.1111/mam.12067Search in Google Scholar

Nores, C., L. Llaneza and Á. Álvarez. 2008. Wild boar Sus scrofa mortality by hunting and wolf Canis lupus predation: an example in northern Spain. Wildl. Biol. 14: 44–51.10.2981/0909-6396(2008)14[44:WBSSMB]2.0.CO;2Search in Google Scholar

Nowak, S., R.W. Mysłajek, A. Kłosińska and G. Gabryś. 2011. Diet and prey selection of wolves (Canis lupus) recolonising Western and Central Poland. Mamm. Biol. 76: 709–715.10.1016/j.mambio.2011.06.007Search in Google Scholar

Papageorgiou, N., C. Vlachos, A. Sfougaris and E. Tsachalidis. 1994. Status and diet of wolves in Greece. Acta Theriol. 39: 411–416.10.4098/AT.arch.94-48Search in Google Scholar

Pimenta, V., I. Barroso, L. Boitani, P. Beja. 2017. Wolf predation on cattle in Portugal: assessing the effects of husbandry systems. Biol. Conserv. 207: 17–26.10.1016/j.biocon.2017.01.008Search in Google Scholar

Pritchard, J.K., M. Stephens and P. Donnelly. 2000. Inference of population structure using multilocus genotype data. Genetics 155: 945.10.1093/genetics/155.2.945Search in Google Scholar

Rogdakis, E. 2006. General zootechnics. Stamouli Press, Athens (in Greek).Search in Google Scholar

Sazatornil, V., A. Rodríguez, M. Klaczek, M. Ahmadi, F. Álvares, S. Arthur, J.C. Blanco, B.L. Borg, D. Cluff, Y. Cortés, E.J. García, E. Geffen, B. Habib, Y. Iliopoulos, M. Kaboli, M. Krofel, L. Llaneza, F. Marucco and J.V. López-Bao. 2016. The role of human-related risk in breeding site selection by wolves. Biol. Conserv. 201: 103–110.10.1016/j.biocon.2016.06.022Search in Google Scholar

Sidorovich, V.E., L.L. Tikhomirova and B. Jedrzejewska. 2003. Wolf Canis lupus numbers, diet and damage to livestock in relation to hunting and ungulate abundance in northeastern Belarus during 1990–2000. Wildl. Biol. 9: 103–111.10.2981/wlb.2003.032Search in Google Scholar

Teerink, B. 1991. Hair of West European mammals: atlas and identification key. Cambridge University Press, Cambridge.Search in Google Scholar

Torres, R.T., N. Silva, G. Brotas and C. Fonseca. 2015. To eat or not to eat? The diet of the endangered Iberian wolf (Canis lupus signatus) in a human-dominated landscape in central Portugal. PLoS One 10: e0129379.10.1371/journal.pone.0129379Search in Google Scholar PubMed PubMed Central

van Asch, B., C. Alves, L. Santos, R. Pinheiro, F. Pereira, L. Gusmão and A. Amorim. 2010. Genetic profiles and sex identification of found-dead wolves determined by the use of an 11-loci PCR multiplex. Forensic Sci. Int. Genet. 4: 68–72.10.1016/j.fsigen.2009.05.003Search in Google Scholar PubMed

Vos, J. 2000. Food habits and livestock depredation of two Iberian wolf packs (Canis lupus signatus) in the north of Portugal. J. Zool. 251: 457–462.10.1111/j.1469-7998.2000.tb00801.xSearch in Google Scholar

Wagner, C., M. Holzapfel, G. Kluth, I. Reinhardt and H. Ansorge. 2012. Wolf (Canis lupus) feeding habits during the first eight years of its occurrence in Germany. Mamm. Biol. 77: 196–203.10.1016/j.mambio.2011.12.004Search in Google Scholar

Weaver, J.L. 1993. Refining the equation for interpreting prey occurrence in gray wolf scats. J. Wildl. Manag. 57: 534–538.10.2307/3809278Search in Google Scholar

Received: 2018-02-09
Accepted: 2018-12-21
Published Online: 2019-02-21
Published in Print: 2019-11-26

©2019 Walter de Gruyter GmbH, Berlin/Boston

Downloaded on 2.6.2024 from https://www.degruyter.com/document/doi/10.1515/mammalia-2018-0021/html
Scroll to top button