Open Access
How to translate text using browser tools
1 June 2016 Roles of Raptors in a Changing World: From Flagships to Providers of Key Ecosystem Services
José A. Donázar, Ainara Cortés-Avizanda, Juan A. Fargallo, Antoni Margalida, Marcos Moleón, Zebensui Morales-Reyes, Rubén Moreno-Opo, Juan M. Pérez-García, José A. Sánchez-Zapata, Iñigo Zuberogoitia, David Serrano
Author Affiliations +
Abstract

Birds of prey have been, in comparison to other avian groups, an uncommon study model, mainly due to the limitations imposed by their conservative life strategy (low population density and turnover). Nonetheless, they have attracted a strong interest from the point of view of conservation biology because many populations have been close to extinction and because of their recognised role in ecosystems as top predators and scavengers and as flagship species. Today, after more than a century of persecution, and with the exception of some vultures still very much affected by illegal poisoning, many populations of birds of prey have experienced significant recoveries in many regions of Spain and the European Mediterranean. These changes pose new challenges when addressing the conservation of raptors in the coming decades. On this basis, and from a preferentially Mediterranean perspective, we have focused our attention on the need of describing and quantifying the role of these birds as providers of both regulating (rodent pest control and removal of livestock carcasses) and cultural ecosystem services. Moreover, we revisited persisting conflicts with human interests (predation of game species) and call attention to the emergence of new conflicts with a strong social and media component such as the predation on live cattle by vultures. Also, the rampant humanization of the environment determines the need for new solutions to the growing, yet scarcely explored, problem of accidents in new infrastructures such as mortality in wind farms. Finally, we explored in depth the ecological response of birds of prey to large-scale habitat changes such as urbanisation and abandonment of marginal lands that are also expected to increase in the near future. We urgently need more scientific knowledge to provide adequate responses to the challenge of keeping healthy populations of avian predators and scavengers in a rapidly changing world.

Introduction: Raptors multiplying and its ecological and conservation aftermath

Anthropogenic activities are leading to broad-scale non-random changes in bird community compositions (Bonebrake et al., 2010; Le Viol et al., 2012). This is not a new phenomenon as human pressures, mainly through persecution and modification of habitats, have had historical consequences for bird populations all around the world. For example, it is estimated that up to 1,300 avian species could have been wiped out in only a few centuries by the colonising Pacific islanders (see among others Steadman, 1995; Duncan et al., 2013). What is new is the current speed of the process. The rampant humanization of the planet following the industrial revolution has led to a new epoch, the “anthropocene”, which involves the rapid disappearance of species, the so-called “sixth extinction” (Barnosky et al., 2011; Dirzo et al., 2014). Although bird extinctions have been exceptional in Europe during the last century generalised declines are evident in many groups. For instance, those birds associated with farmlands and/or the transaharan migrants have suffered clear declines since the middle of the XXth century. In contrast, forest birds have been relatively well conserved, due to the widespread tendency of land abandonment in marginal areas (Donald et al., 2001; Gregory et al., 2007; BirdLife International, 2013; Reif, 2013).

Within this general scenario of biotic homogenisation, environmental impoverishment and decline of many primary consumers it would be expected that predatory birds also would show negative population trends. However, this is not so. In fact, birds of prey (Accipitriformes and Strigiformes) have shown striking changes during the last few decades, switching in many cases from an almost “terminal” status (see e.g., Geroudet, 1964) to spectacular recoveries in both numbers and distribution ranges (see below). It is well-known that since the end of the XIXth century, birds of prey were intensively persecuted in Europe to which was added the mode of the scientific collecting (see e.g., Chapman and Buck, 1893, 1910; Hiraldo et al., 1979). In Spain, it was in the middle of the XXth century, following the creation of the “Juntas de Extinción de Animales Dañinos” (Boards for the extinction of harmful animals), when slaughtering was implemented as a systematic practice (Garitacelaya, 2003; Paulos, 2006). A similar picture was found in the rest of Europe and in many regions of North America and Australia (Bijleveld, 1974; Newton, 1979; Olsen, 2006). Following the legal protection of all species of birds of prey, many populations have shown a notable recovery in Europe (Deinet et al., 2013) with the notable exception of some scavenger species whose decline still continues, especially in the Eastern Mediterranean, mainly due to illegal poisoning (Donázar et al., 2009). The shift is clearly linked to the stopping of non-natural mortality by direct killing and the progressive correction of other limiting factors, such as the banning of some pollutants and the correction of intrusively positioned power lines (Newton, 1979), as well as to broadscale habitat protection and the adoption of species-specific conservation measures and broad-scale habitat protection (Donald et al., 2007; Williams et al., 2012; Kolecek et al., 2014) (fig. 1).

But the struggle continues. During the last half century raptor populations have suffered successive crises with potentially very negative consequences such as the irruption of the European wild rabbit haemorrhagic disease at the end of the XXth century and the restrictive EU sanitary measures limiting the disposal of livestock carcasses to scavengers at the beginning of the XXIth century (Telia, 2001; Donázar et al., 2009) which required the immediate reaction of scientists, managers and conservationinsts (Margalida et al., 2012). For future decades the main challenge is to determine how populations of birds of prey, along with other organisms, will envisage their fitting into a world shaped by anthropogenic activities and how healthy populations of raptors and human interests can coexist. This is our main objective in this essay (fig. 2). Within a global scenario, our perspective focuses mainly on the Mediterranean basin, where, as in the rest of the world, studies of ecology and conservation of raptors flourished from the mid-90s of the last century, but that currently shows a certain delay that requires correction (fig. 3). First, and because successful conservation must be based on scientific knowledge, we examine the position of raptor research in the field of vertebrate ecology. Then we examine in depth the role played by birds of prey as flagship and umbrella species within general nature conservation strategies. After, a review of the role of raptor as ecosystem services providers is followed by an analysis of the conflicts arising between raptor conservation and human interests. Finally, we approach the future of birds of prey in light of the prospective changes in ecosystems and landscapes.

Fig. 1.

Non-natural mortality caused by human activities is the main limiting factor for birds of prey populations in Europe. Poisoning is causing the precipitous regression of different scavenger species (a. a poisoned Egyptian vulture) whereas direct killing is losing relative importance but still is a menace factor for some migratory routes (b, radiography of a shot Spanish imperial eagle). New developments such as wind farms are gaining importance during the last decades, being responsible now for the deaths of thoushands of birds of prey (c. a griffon vulture mutilated by a wind turbine) and other flying vertebrates. Accidents in power lines are still probably the main factor of mortality for many species of raptors (d. an electrocuted eagle owl). Credits: a: EBD-CSIC: b: Centro de Análisis y Diagnóstico de la Fauna Silvestre-C.A.D. Consejería de Medio Ambiente y Ordenación del Territorio: c: Eugenio Montelío: d) Juan M. Pérez-García.

[La mortalidad no natural causada por actividades humanas es el principal factor limitante para poblaciones de rapaces en Europa. Los envenenamientos son causa del abrupto declive de varias especies de carroñeros (a. alimoche común envenenado), mientras que la persecución directa ha perdido importancia relativa pero es todavía un factor de amenaza en algunas vías migratorias (b, radiografía de un águila imperial ibérica tiroteada). Nuevos desarrollos como los parques eólicos están ganado im-portancia durante las últimas décadas, siendo responsables de la muerte de miles de grandes rapaces (c. a buitre leonado mutilado por una turbina eólica). Los accidentes en tendidos eléctricos son todavía el principal factor de mortalidad para muchas especies de rapaces (d. búho real electrocutado). Creditos: a: EBD-CSIC; b: Centro de Análisis y Diagnóstico de la Fauna Silvestre-C.A.D. Consejería de Medio Ambiente y Ordenación del Territorio; c: Eugenio Montelío; d) Juan M. Pérez-García.]

f01_181.jpg

Fig. 2.

Main benefits and challenges for conserving raptors in a changing world. Raptors are fruitful study models for several ecological and evolutionary disciplines, are key elements in many biodiversity conservation schemes that can exert positive effects to a wide array of organisms and perform crucial ecosystem functions that translate into important ecosystem services. On the other hand, conflicts with hunters and, to a lesser extent, farmers may compromise the future of some raptor populations. Also, raptors living in Mediterranean countries will need to cope with changes in their habitats due to increasing urbanization around big cities and abandonment of rural areas.

[Principales beneficios y retos en la conservación de aves de presa en un mundo en cambio. Las rapaces son buenos modelos de estudio en determinadas disciplinas dentro de la ecología y evolución, son además elementos clave en muchas estrategias dentro de la biología de la conservación, ejerciendo efectos positivos sobre muchos otros organismos ecológicos y de conservación, y llevan a cabo funciones ecosistémicas clave que se traducen en servicios ecosistémicos. Desde otro punto de vista, los conflictos con cazadores y, en menor medida, con ganaderos, pueden comprometer el futuro de algunas poblaciones de aves de presa. Las rapaces que viven en países mediterráneos deben afrontar fuertes cambios de hábitat asociados a la urbanización en torno a grandes ciudades y ai abandono de áreas rurales.]

f02_181.jpg

Fig. 3.

Number of raptor-related papers published in peer-reviewed journals between 1960 and 2015. It is distinguished between total (black) and those focused in the Mediterranean region (grey). Based on a compilation from the Scopus search engine.

[Número de artículos sobre aves rapaces publicados en revistas científicas con revisión por pares entre 1960 y 2015. Se distingue entre el total (línea negra) y aquellos que se centran en la región mediterránea (línea gris). Basado en una búsqueda a través de Scopus.]

f03_181.jpg

Raptors in animal ecology research

Birds of prey might not seem the most ideal study models in animal ecology as they often occur at low densities over large ranges, breed in sites of difficult access, show elusive behaviour, and have long generation times, all of which which makes it difficult to obtain large sample sizes, follow entire cohorts of marked individuals, and apply experimental designs. At the same time, they encompass very different lifestyles, from territorial to colonial, strictly sedentary to vagile or migratory, or from arthropod- or snail-eaters to obligate scavengers, opening substantial opportunities for the study of a wide range of ecological and evolutionary questions. Animal ecology in its classical sense deals with the relationships between animals and the environment and how these relationships shape animal distribution and abundance, with fundamental contributions of birds of prey in the literature (e.g. trophic ecology and predator-prey dynamics, factors shaping abundance and distribution, population dynamics of long-lived animals). Nowadays, this discipline includes many functional and evolutionary aspects that exceed its initial meaning, and birds of prey have been employed as study models to advance decisively in some of them (e.g. signalling, sexual selection, sibling competition, parental and sexual investment, lifehistory strategies). However, it is the traditional view of animal ecology that has a more direct and obvious link with applied aspects such as management and conservation, and therefore it is within this context that we focus our attention in this review.

The first ecological studies on raptors emerged in parallel with the advent of animal ecology as a formalised discipline (see e.g. Breckenridge, 1935). However, raptor ecology actually began to flourish after the pesticide crisis in the 1950s and 1960s and the inspiring book by Ian Newton (Newton, 1979). In this book, Newton addressed many aspects of raptor ecology and paved the way for most modem research. Many of these aspects were then in its infancy and are much better known nowadays, while others remain rather underinvestigated. In this section, we briefly discuss some of the most significant advances and promising avenues in three fundamental and related areas of animal ecology: population, spatial and movement ecology.

Population ecology and dynamics

Understanding population dynamics and regulation is pivotal in animal ecology. Pattern-oriented approaches (Coulson et al., 2000) investigate what regulates raptor populations by studying the feedback between density and population size or fecundity (e.g. Ferrer and Donázar, 1996; Krüger and Lindström, 2001). Whether the frequently reported negative relationship is due to the preemptive occupation of the best sites (the habitat heterogeneity hypotheses), to interference between individuals (the individual adjustment hypothesis), or to other causes such as changes in the age-structure of populations is however difficult to infer without additional information, making this approach controversial (Balbontín and Ferrer, 2008; Beja and Palma, 2008; Carrete et al., 2008; Ferrer et al., 2008). Clearly, experimental approachess are needed, but they are often ethically inadmissible or hardly feasible in raptors. Fruitful ways of further advance in this field should include more detailed behavioural information on habitat selection, territory defence and social interactions (Serrano and Telia, 2007), as well as information on how territory size and quality change with density and population size (Both and Wiser, 2000; Carrete et al., 2006), and to what extent variations in fecundity are motivated by intrinsic properties of habitats and/or to phenotypic traits of individuals (Krüger et al., 2012). In addition, floaters (the non-breeding fraction of the population) may play a key but largely overlooked role in the regulation of breeding populations (Penteriani et al., 2005), not only because they compensate any population decrease by moving to a breeder position, but also by potentially affecting fecundity of breeders in multiple ways (e.g. by decreasing feeding rates or increasing stress or exposure to predators of territory owners, see López-Sepulcre and Kokko, 2005). However, it should be noted that regulatory processes do not only act throughout productivity, but the key regulatory mechanisms (e.g. density dependent competition by interference) may also operate on survival, even in the presence of densitydependent fecundity depression (Nevoux et al., 2011). For these reasons, progress in our understanding of population regulation and limitation in raptors will be only possible if we effectively move from a pattern- to a process-oriented framework in which the responsible mechanisms of density-dependent feedbacks are directly investigated.

In practice, this change of paradigm implies that emphasis should be put on how intrinsic (population composition in terms of age, morphology, behaviour or any other phenotypic trait) and extrinsic (food, predators, competitors, diseases, parasites, weather) factors affect demographic rates (survival, recruitment, reproduction, dispersal) that will ultimately explain population dynamics (Frederiksen et al., 2014). This approach requires high-quality longitudinal data on known-age individuals, which means in practice that marking and monitoring programs need to be maintained in the long-term. Once the determinants of vital rates are known, they can be integrated in population models that provide estimates of deterministic and stochastic population growth rates. Stageand age-structured matrix models (Caswell, 2001) have been among the most popular, particularly, to assess the effect of current and future limiting factors on population trajectories and long-term viability (e.g. Altwegg et al., 2006; Millon et al., 2014; Sanz-Aguilar et al., 2015), and to identify the most influential demographic parameters (e.g. Hiraldo et al., 1996). Although the time is long past when populations were considered geographically closed, sensu lato metapopulation dynamics remains largely underinvestigated in raptors. This is a key element in ensuring effective conservation actions, as the scarce evidence to date have proven that connectivity plays a critical role in the persistence of both local populations and the entire metapopulation, and that local management decisions may have range-wide consequences (Kauffman et al., 2004; Schaub, 2006; Hernández-Matías et al., 2013; Schumaker et al., 2014). Finally, new mathematical models are fuelling a more efficient and effective use of data and we will certainly see more progress in the future. Integrated population models are a powerful tool that make use of multiple and independent data (e.g. annual population counts and demographic data), providing more precise parameters and projections with the full range of uncertainties (Frederiksen et al., 2014) . These models have proved useful in furnishing robust evidence about the paramount role of immigration on raptor population growth and persistence (Abadi et al., 2010; Altwegg et al., 2014; Tempel et al., 2014; Lieury et al., 2015).

Spatial ecology: raptor-habitat relationships

Habitat use, preference, and selection have been usually confounded in the literature (Johnson, 1980), although the goal of most studies, regardless of the scale of analysis, is to characterise habitat suitability. This has been usually approached by modelling data on presence-absence, presence only, density, and/or abundance as a function of environmental covariates (see e.g. Donázar et al., 1993; Meyer et al., 1998). Inspired by the complexity and dynamic nature of animal distributions and abundances, some studies have dealt with this topic at multiple temporal and spatial scales or at different activityspecific environments for nesting, roosting or feeding (e.g. Thompson and McGarigal, 2002; Martinez et al., 2003; D'Elia et al., 2015). These approaches have generated an important body of knowledge with a wide range of applied uses, such as understanding the ecological requirements or biogeography of species, predicting species distributions, identifying reintroduction sites and priority conservation areas, and predicting the response of populations to climate change or habitat loss (Franklin, 2009). However, distribution and ecological niche models have limitations because underlying mechanisms are indirectly inferred from patterns of distribution or habitat use, so they may be confounded and are not always clear or generalizable (Martinez et al., 1999; Bamford et al., 2009). For example, a given species may be absent in some localities because individuals avoid certain habitat features, but also because they are systematically killed by man or predators. Further, in avoiding competition or disturbance birds may be forced to use habitats they do not actually prefer, decoupling distribution from habitat preferences. In some circumstances, the behaviour and use of habitat resources may be directly observed by recording habitats where individuals forage or roost (Bakaloudis et al., 1998; Ursúa et al., 2005), although biases associated with, for example, habitat-mediated detectability are sometimes unavoidable. These problems can be circumvented with telemetry-based data which are being incorporated routinely in to the study of raptor-habitat relationships since the pioneer work of Southern (1964). Telemetry has provided a pile of information on habitat use and selection since then to study, for example, daily and seasonal variations, amplitude of home ranges in relation to territory quality, prey abundance or habitat fragmentation, and differences between sex- and age-clases (e.g. Bechard, 1982; Telia et al., 1998; Carrete and Donázar, 2005; Sunde and Redpath, 2006; Fernández et al., 2009; Campioni et al., 2013; Tanferna et al., 2013; Rivers et al., 2014). The understanding of the mechanisms underlying individual variability in habitat use is however poorly known and very limited as yet. A few studies on raptors have shown individual differences in settlement areas depending on previous experience during the wandering stage (Delgado et al., 2010), or in prey selection and diet breadth in relation to foraging habitat selection (Terraube et al., 2014). Davis and Stamps (2004) proposed that experience in the natal habitat may shape habitat selection during adulthood, but information on birds of prey is controversial (Mannan et al., 2006; Delgado et al., 2010). Apart from this scarce evidence, very little is known about whether genetic, cultural or ontogénie factors shape between and within individual differences and to what extent they may be flexible in their habitat preferences. This may well be one of the most promising and productive areas of research on raptor-habitat relationships in the future.

Movement ecology: dispersal and migration

Dispersal and migration involve the longest movements performed by birds throughout their life-cycles. Dispersal is the glue that binds populations and allows the expansion of distribution ranges and colonization of new areas, with a key role in population dynamics and evolutionary processes. Most research on dispersal in birds of prey has focused on the causes motivating it and on the cues used by individuals to settle in a new location, both in natal and breeding dispersal contexts.

Using capture-recapture techniques, a wide array of individual, environmental and social correlates has been identified (e.g. Korpimäki, 1993; Wiklund, 1996; Serrano et al., 2001), but few studies have evaluated comprehensively the fitness payoffs of adopting different dispersal strategies (Forero et al., 2002; Serrano and Telia, 2012). Additionally, molecular methods have increasingly been used to infer effective dispersal and connectivity among subpopulations (Martinez-Cruz et al., 2004; Brito, 2007; Hull et al., 2008; Alcaide et al., 2009). Although both capture-recapture and molecular data provide valuable information on realised dispersal patterns, their main handicap is that they rarely give in-depth insight into the underlying movement process. For this reason, one of the major gaps of knowledge about raptor dispersal is the transient stage separating emigration from a site and immigration to another (see Penteriani and Delgado, 2009). This is particularly important in natal dispersal of long-lived raptors with delayed sexual maturity because they usually spend several years in non-breeding areas in which they may be exposed to differential pressures. Much progress in this respect has been made possible thanks to the advent of modem tracking tools, showing that in some species, such as large eagles and territorial vultures, pre-breeders wander over vast geographic areas and settle temporarily in sites with high food resources, usually located far away from breeding grounds (Balbontín, 2005; Cadahía et al., 2010; Margalida et al., 2013). In other species, such as large owls, young individuals try to settle as soon as possible in well-defined home ranges, and temporary settling areas used are often located near breeding territories of conspecifics (Rohner, 1997; Delgado et al., 2010; Penteriani and Delgado, 2012). Further, floaters and failed-breeders may prospect different areas to gather information on habitat quality to optimise future settlement decisions (Serrano et al., 2003; Sergio and Penteriani, 2005), although the precise pathways covered and the cues used during prospecting remain poorly known (Nathan et al., 2012; Therrien et al., 2015).

Biotelemetry and experimental approaches have also been useful to demonstrate that dispersal in birds of prey is determined by the interplay between endogenous factors (e.g. hormones and body condition) and the social environment or the spatial configuration of the landscape (see Belthoff and Dufty, 1998; Delgado et al., 2010).

Migratory raptors are in turn exposed to different environmental conditions during breeding, migration and wintering, so patterns observed at one stage may be affected by conditions experienced in another (Grande et al., 2009; Mihoub et al., 2010). New tracking devices and other technological innovations have provided detailed information on migration routes, wintering, summering and stopover areas, causes and rates of mortality experienced en route, and migration strategies that would have been unthinkable until recently (Rodriguez et al., 2009; Bohrer et al., 2012; Klaassen et al., 2014; Themen et al., 2014). Moreover, they have provided information on within and between individual variability in speed and amplitude of movements, flying performance in relation to food resources and weather conditions (Lanzone et al., 2012; Dodge et al., 2014), timing and routes of migration (López-López et al., 2014), and their link with life-history strategies and demographic issues (Sergio et al., 2014).

However, carry-over effects, i.e. processes occurring in one season that affect individual performance in another season (Harrison et al., 2011), remain largely unknown in birds of prey, although they have been well documented in other migratory birds (e.g. Norris et al., 2004) and constitute an interesting field of open research. Finally, it has been shown that resident and migrant raptors may segregate spatially in sympatry (e.g. Cardador et al., 2015), so another promising line of research is partial migration (populations composed of a mixture of resident and migratory individuals). Partial migration is ubiquitous in animals but has been scarcely explored in spite of its potential to further our understanding of the evolution of migratory behaviours, its ecological consequences, and its applied implications in a context of global change (Chapman et al., 2001).

From a general point of view, movement ecology is a growing field with an increasingly better established theoretical framework (Nathan et al., 2008). From the first studies of long-distance dispersal and migratory movements investigated by following radiotracked raptors with aircraft (Cochran, 1975; Beske, 1982), the use of satellite telemetry has revolutionised the study of movement ecology in birds of prey. Now the challenge is to visualise and analyse the overwhelming volume of available data (Gurarie et al., 2016), as well as to identify how individuals take movement decisions under different environmental conditions depending on their own state and previous experience. This will certainly enrich our understanding of raptor ecology and conservation in the years ahead.

The role of raptors in nature conservation

Raptors within modern progress in wildlife conservation

Nature conservation has experienced a continued transition in relation to the commitments of authorities to promote it worldwide. After the first private initiatives funded by global organizations for the protection of nature (i.e. IUCN, WWF, ICPB) that took place in the mid twentieth century, it was not until the 1970s when environmental awareness acquired a legal status in Europe (De Klemm and Shine, 1993). The milestone that initially launched these global environmental commitments came in 1972 with the UN Convention on the Human Environment and the establishment of the United Nations Environmental Program ( www.unep.org/); further, the endorsement of the first international conventions and treaties on the protection of species and habitats (RAMSAR, OSPAR, Barcelona) was triggered. The decade ended with the adoption of the Convention on the Conservation of European Wildlife and Natural Habitats in Europe (Bern Convention  www.coe.int/en/web/bern-convention/home). It was the first to include a list of strictly protected species establishing a new and binding approach, and forcing countries to take active measures to protect them. The Convention on the Conservation of Migratory Species (Bonn Convention  www.cms.int/) and the Birds (2009/147/EC) and Habitats (43/92/EEC) Directives also deepen the need for protecting some species, prioritised against others depending on their conservation status, scientific interest, rarity or cultural value.

Within these different conventions, treaties and regulations, birds of prey have had a notorious presence almost monopolising the lists of protected species. Along with waterfowl (Kirby et al., 2008) and reptiles, raptors have been the group with a higher level of protection at the European level in different reference texts (table 1). Thus, there are significant differences in the proportion of protected species between raptors and other taxonomic groups in both the Bern Convention (fi01_181.gif = 20,532.4; p < 0.001) and the European Directives (fi01_181.gif = 20,764.3; p < 0.001). Likewise, birds of prey have also been the subject of priority conservation and recovery projects. Taking as reference the 740 LIFE projects awarded by the EU between 1992 and 2014 for species of wild flora and fauna ( http://ec.europa.eu/environment/life/project/Projects/), 14.0% were aimed to one or several species of raptors, being these projects about a third (31.9%) of those devoted to the protection of birds.

Raptors as flagship species and other surrogate concepts

Why do protection and devotion are so strongly biased to raptors? Why do raptors enthral society to become a priority group? Several reasons have been highlighted within this regard that could justify their importance: 1) the attractiveness that their behaviour provokes in people, their aesthetic attributes and symbolism (Douglas and Verísimo, 2013), 2) the traditional persecution that raptors have suffered due to interactions with anthropic interests, and the resulting environmentalist awareness awakened in different social sectors (Redpath et al., 2013), 3) their rarity or the overall unfavourable conservation status of several species (Burfield, 2008), and 4) their role as top-predators and their ecological requirements whose fulfilment depends on the environmental health of the ecosystems they inhabit (Sergio et al., 2008).

Table 1

Number and percentage (in brackets) of species of different taxonomic groups, protected under the Bern Convention and the Birds 2009/147/CE and Habitats 92/43/CEE Directives, in relation to the total number of existing species. Sources: Flora: Euro + Med PlantBase, 2006–2011  http://www.emplantbase.org/home.html; Invertebrates: Wieringa, 1995; Fishes: Freyhof and Brooks, 2011; Amphibians: Temple and Cox, 2009; Reptiles: Cox and Temple, 2009; Birds: BirdLife International, 2015; Mammals: Temple and Terry, 2007.

[Número y porcentaje (paréntesis) de especies pertenecientes a diferentes grupos taxonómicos protegidos bajo la Convención de Berna y las Directivas de Aves (2009/147CE) y de Hábitats (92/43/CEE) en relación con el número total de especies existente. Fuentes: Flora: Euro + Med PlantBase, 2006–2011  http://www.emplantbase.org/home.html; Invertebrados: Wieringa, 1995; Peces: Freyhof and Brooks, 2011 ; Anfibios: Temple and Cox, 2009; Reptiles: Cox and Temple, 2009; Aves: BirdLife International, 2015; Mamíferos: Temple and Terry, 2007.]

t01_181.gif

The leading role of raptors in active conservation initiatives during the last 40 years granted them the tag of “flagship species” (Sergio et al., 2008), being organisms for which protection has been prioritised by all public administrations, NGOs and media, similar to the level of large carnivore mammals (Sergio et al., 2006). As a result, a large number of actions have been developed like reintroduction projects, land stewardship, scientific research, solving threatening factors or environmental awareness (see “Other search options” in  www.iucnredlist.org). These great investments and efforts towards raptor protection in Europe has had positive effects on those species with a more unfavorable conservation status and prioritised in different regulations. Thus, of the 24 endangered species of bird of prey at the European level and/or with ongoing EU action plans, more than half have improved their status in the last 20–40 years, 29.1% are stable or no clear information on their trend is available, and 16.6% show a declining trend (table 2).

Table 2

Population status (breeding pairs) of the 24 diurnal and nocturnal raptors included in priority lists of European protected species (“Source”: a = IUCN Red List, in the categories NT; VU; EN; CR; b = European Union's Species Action Plans for Birds; c = Spanish Catalogue of Threatened Species; and d = Other flagship raptor species), between first available global census and most recent ones, and their trend (+: positive; negative; stable or unknown according to BirdLife International, 2005).

[Estatus poblacional (número de parejas) de las 24 rapaces diurnas y nocturnas incluídas como prioritarias en las listas europeas de especies protegidas. (“Source”: a = Libro Rojo de la UICN, en las categorías NT; VU; EN; CR; b = Planes europeos de acción para aves; c = Catálogo español de especies amenazadas; y d = otras rapaces “paraguas”). Se muestra el cambio entre el primer censo global disponible y el más reciente, así como la tendencia ( + : positiva; negativa; estable o desconocida; de acuerdo con BirdLife International, 2005).]

t02_181.gif

An important body of scientific literature has been published regarding the effects that active raptor conservation has had on other components of the environment, following the “surrogate” concept (i.e. species used as sensitive indicators of a conservation problem, to track population changes of other species or to locate areas of high biodiversity; Caro and O'Doherty, 1999). There are heterogeneous results regarding the relationships between presence and relative abundance of birds of prey and geographically co-occurring environmental indices such as diversity, species richness and threatening factors. On the one hand, several authors have demonstrated the role of raptors as “biodiversity indicators”, establishing a positive and direct relationship between richness and abundance of raptors and other animal or plant species in the same areas (Sergio et al., 2006; Martín and Ferrer, 2013; Burgas et al., 2014). However, this principle is far from being widespread because some biomes such as semidesert do not show any of the mentioned relationships with raptors (see Estrada and Rodríguez-Estrella, 2016). Also, links between raptors and species with different habitat requirements or with less inter-specific attraction seem lower (Senzaki and Yamaura, 2015). Otherwise, raptors properly act as “sentinels” of different local and large-scale environmental changes and global threats to biodiversity, such as the dynamic of their prey populations, the levels of pollutants and illegal poaching activities (García-Fernández et al., 2008; Helander et al., 2008, Molina-López et al., 2011). However, a direct effect on the performance of raptor populations from change and habitat alteration processes when they occur in low or moderate intensity has been analysed and ruled out (Rodríguez-Estrella et al., 1998), possibly due to their plasticity in satisfying several of their ecological requirements (Donázar et al., 2010; Sullivan et al., 2016). Finally, less clear is their role as “umbrella species” (Branton and Richardson, 2011). There is a small number of conclusive studies (i.e. Senzaki et al., 2015), and therefore a lower overall consensus on the reliability of considering birds of prey such as drivers that meet the complete needs of all co-occurring species in all ecosystems and habitats, given the difficulty in standardising their ecological requirements (land use, ecosystem functionality, ecological processes, etc) genetically (Carrete and Donázar, 2005; Sergio et al., 2008). Overall, the diversity of life strategies of raptors and the different conservation status within the same species at the regional level make their role as sentinel, indicator or umbrella species highly context-dependent.

In any case, conservation plans based on raptors presence have shown positive effects in providing global environmental benefits derived from the increase in the land surface covered with management plans, the enhanced environmental awareness and the fight against global threats for biodiversity. Thus, their consideration as “flagship species” allows that the protection and proper management of the territory inhabited by raptors also impact on a large amount but not all the species and habitats, thus optimising efforts and performance of the financial investment prioritised towards these top-predators and the achievement of global environmental goals (Sergio et al., 2008). In relation to adaptive management (Salafsky et al., 2001), the accurate knowledge of the ecology and threats of birds of prey (table 2) in comparison to other taxa has been widely used as a tool for updating the assessment of their conservation status (i.e. BirdLife International, 2015) and, as a consequence, the application of a wide range of protection techniques regularly assessed from a scientific perspective (i.e. power line retrofitting, poisoning prosecution, habitat management; Balbontín, 2005; Margalida et al., 2014; Chevallier et al., 2015; Hernández-Matías et al., 2015; Matthiopoulos et al., 2015).

Raptors as providers of ecosystem services

Biodiversity loss and alteration of the environment are increasingly threatening ecosystem functions and associated ecosystem services worldwide (e.g. Loreau et al., 2001; Hooper et al., 2005). Ecosystem services can be defined as the benefits people obtain from ecosystems and the species that make them up (MEA, 2003). Millennium Ecosystem Assessment establishes four broad types of ecosystem services: provisioning services that relate to the products obtained from ecosystems including for example water or food; regulating services that are the benefits obtained from the regulation of ecosystem processes, for instance, climate regulation; cultural services, that refer to the nonmaterial benefits people obtain from ecosystems through spiritual enrichment, cognitive development, reflection, recreation, and aesthetic experiences; and supporting services, which are those that are necessary for the production of all other ecosystem services, such as soil formation or nutrient cycling (MEA, 2003). Although the idea that humans can benefit from nature is not new, the concept of ecosystem service, which is increasingly used by scientists, managers, and general public (Daily et al., 2009), has allowed us to better understand the consequences of ecosystems change and biodiversity loss for human well-being.

Avian scavengers as providers of supporting, regulating and cultural services

Birds are an ideal study models in ecosystem services research (Şekercioğlu et al., 2004; Şekercioğlu, 2006; Whelan et al., 2008; Green and Elmberg, 2014). Within this field, a “classical” example are the ecosystem services provided by vultures and avian facultative scavengers, but it has received little scientific attention until recently (Moleón et al., 2014), when the number of studies has raised (see Markandya et al., 2008; Gangoso et al., 2013) partially due to the precipitous decline of vultures and other avian scavengers in many regions of the world and specially in Africa and Asia where they were formerly extremely abundant (Ogada et al., 2012, 2016).

Quantification, not only in economic terms, of ecosystem services provided by avian scavengers will be a major future challenge (Wenny et al., 2011). In the Indian subcontinent, the catastrophic vulture decline led to increases in the abundance of mammal scavengers, i.e. rats Rattus sp. and feral dogs Canis lupus familiaris (e.g. Pain et al., 2003; Markandya et al., 2008), which are important reservoirs of pathogens causing bubonic and rabies diseases, respectively. This also increased rates of infection to humans and domestic species (Markandya et al., 2008; Ogada et al., 2012). In this scenario, health costs attributable to vulture declines, i.e. the loss of regulating services, were estimated at more than $2 billion per year (Markandya et al., 2008). In Europe, the outbreak of bovine spongiform encephalopathy (BSE) led to the approval of a sanitary regulation that forced the collection and transport of carcasses of extensive livestock from farms to be destroyed in authorised plants. Apart from negative impacts on vulture behaviour and populations (Donázar et al., 2009; see below), this measure led to the decline of the services provided by vultures (Margalida and Colomer, 2012). In Spain, for instance, it was estimated that the Spanish vulture populations removed between 134 and 200 t of bones and between 5551 and 8326 t of rotten meat every year (Margalida and Colomer, 2012). In addition, supplanting carrion consumption by scavengers with artificial removal of livestock carcasses meant the emission of over 77,000 metric tons of CO, eq. to the atmosphere per year and the annual payment of ca. $50 million to insurance companies by farmers and administrations (Morales-Reyes et al., 2015).

Humans and vultures have been closely related to each other during the course of human history, especially since the rise of agriculture and animal domestication around 10,500 years ago (Moleón et al., 2014). Since then, vultures have been revered and become part of religious symbolism of numerous human cultures, as derived from the ancient Egyptian goddesses Mut (a woman with vulture's wings) and Nekhbet (a vulture), and monuments from the Maya civilisation (Morelli et al., 2015). Even today, the Parsees in Asia depend upon funeral services provided by vultures, which remove their dead relatives placed in the ‘Towers of Silence’ (Pain et al., 2003).

Recently, the observation of birds (i.e. birdwatching) has become one of the most promising branches of ecotourism, which provides important economic profits and may favour conservation of natural areas (Şekercioglu, 2002). However, the economic benefits of these cultural services have been rarely quantified. In the United States there were 47 million birdwatchers in 2011. They spent $41 billion in trip and equipment expenditures, generating $107 billion in total, and creating 666,000 jobs (Carver, 2013). Specifically, the observation of vultures and other large raptors is an increasing activity and supplies important recreational services. For example, in a natural reserve of Israel, the griffon vulture Gyps fulvus watching provided benefits around $1.1–1.2 million per year (Becker et al., 2005).

Scavenging plays a fundamental role in the maintenance of the stability, structure and dynamics of the food-webs (Wilson and Wolkovich, 2011). Although supporting services provided by scavengers still remain relatively unexplored it seems clear that avian carrion eaters may contribute to the nutrient redistribution among ecosystems (DeVault et al., 2003; Beasley et al., 2015) thanks to their ability to travel great distances. Nutrient cycling benefits plants that then produce oxygen, food, medicine or erosion control (MEA, 2003; Wenny et al., 2011). Finally, carcass recycling implies an increase in nutrient availability in the soil (Melis et al., 2007), which could have effects on microbes, plants and invertebrates.

Biological control of rodent agricultural pests by raptors

The size and growth rate of human population has required huge environmental transformations from heterogeneous landscapes with complex ecological interactions to homogeneous and ecologically simplified areas devoted to food production, settlement and rubbish treatment-deposition. These new ecosystems are almost free of predators, competitors, parasites and pathogens and represent optimal breeding grounds plenty of food for a number of human commensal organisms, such as some rodent species. The high capacity of adaptation of microtine voles, murine rats and mice favoured by their high intrinsic reproductive potential has promoted the rapid colonisation of these novel habitats. In addition, current agricultural practices, such as cropping intensification, increasing irrigation and cultivation of green herbaceous plants throughout the year, create ideal conditions for rodents to expand and breed continuously (Singleton et al., 2010; Jareño et al., 2015). Consequently, many small, extremely fertile, rodent species have experienced population outbreaks becoming pests in agricultural areas and villages. These rodent pests have provoked considerable damage in crop production and caused human health problems (Singleton et al., 2010; Buckle and Smith, 2015).

Rodent pest control

The first strategies used to combat rodent pests were developed in the 1950s and were based on the use of chemical poisons, including anticoagulants (Hadler and Buckle, 1992). This was followed by a period of reassessment of chemical procedures due on the one hand to the resistance developed by rodents to chemicals and, on the other hand, to the social awareness about the environmental impact of chemical poisons (Singleton et al., 2010). Although some old-fashioned minds leading regional administrations persist in the use of this technique, pest-control experts recognise that the use of rodent-killing materials alone is not enough to control rodent pests (Meyer and Kaukinen, 2015). The recent action lines for rodent pest control are based on a set of simultaneous methods. The idea of integrating different methods (including biological control) to combat rodent pests goes back a long way (Howard, 1967), although it began to create interest in the 1990s through the so-called Integrated Pest Management (IPM) scheme (Kogan, 1998; Singleton et al., 2010). The IPM plan aims to find durable strategies that combine biological and cultural controls with limited pesticide use. From this scheme arises what was called the Ecologically Based Rodent Management (EBRM; Singleton, 1999).

Raptors as biological control of rodent pests

One of the measures pursued by the IPM and EBRM is the increase of rodent predator populations. Both diurnal and nocturnal raptors are avian predators whose abundances seem to play a significant role in the population dynamics of rodents (Lima et al., 2002; Fargallo et al., 2009). Furthermore, there are abundant empirical studies about how raptor abundance or raptor rodent consumption relates to rodent population density (e.g. Korpimäki, 1985; Salamolard et al., 2000; Bernard et al., 2009; Millon et al., 2014). Thus, it is plausible to think that raptors have the potential to regulate or limit rodent abundance (Şekercioğlu, 2006). However, the evidence and the number of studies finding causation are notably scarce. Experimental studies are not conclusive (Norrdahl and Korpimäki, 1995; Torre et al., 2007; Maron et al., 2010) and recent reviews about predation effects have shown that whereas experimental mammalian predator removal clearly increases prey population densities, the effect is not so clear in the case of avian predators (Holt et al., 2008; Salo et al., 2010). Similar conclusions can be extrapolated to technical and scientific studies aimed to evaluate the effectiveness of the biological control by raptors (BCR) in rodent pests. Raptors, and particularly barn owls Tyto alba, have a long tradition in their use as a management technique to control rodent pests in palm and rice plantations of Asia (Fall, 1977; Duckett, 1991). The relatively easy methods to attract raptors to plantations by providing artificial nest sites and perches facilitate the direct predation of rodents (Askham, 1990; Duckett, 1991; Motro, 2011). Yet, the efficiency of this measure has been questioned. Criticisms come from the scarcity of reports showing measurable effects of BCR on rodent numbers and/or crop damage (Marsh, 1998; Hygnstrom et al., 1994; Wood and Fee, 2003). The truth is that even today there are very few properly designed studies performed to evaluate the efficiency of BCR. Nevertheless, there are also promising results in the application of BCR to diminish both plantation damage (Murua and Rodríguez, 1989; Hafidzi and Mohd, 2003) and rodent densities in reduced areas around the nests and perches (Paz et al., 2013). These few and contrasting results highlight the need to undertake more properly designed studies to evaluate the efficiency of BCR as an applicable management technique in agro-sylviculture.

To assess the effectiveness of BCR on rodent pests a deep knowledge about the raptor-rodent population dynamics is needed as the capacity of raptors to regulate rodent populations will depend upon the relative abundance of the prey species (Sinclair et al., 1990). Furthermore, the efficiency of biological control techniques has been favoured by the increasing landscape complexity in agricultural systems through the integration of croplands with natural habitats to help the persistence of populations and guilds of predators that reduce pest densities (see Bianchi et al., 2006; Maas et al., 2015), but again this measure has not been investigated in the case of raptors. Therefore information about the predator community composition, its trophic ecology, intraguild competition and habitat requirements is also needed.

Current agricultural view in many developed countries still makes its commitment to cropping intensification and overuse of chemical phytosanitaries (herbicides, fungicides, insecticides, rodenticides, etc) and fertilizers. It is expected that predators will leave the area if they succeed in reducing the population of the only prey species (targeted rodent species) inhabiting the place, as there are no other resources abundant enough to sustain predator populations. The modern IPM scheme offers a valuable framework within which to think about agriculture as a less aggressive environmental exploitation able to maintain a reasonable level of biodiversity, to exploit other economical resources, such as hunting or tourism and benefit from ecological services.

Human-raptor conflicts

Predation on game species

Background

Hunters and raptors preying on game species represent one of the more pervasive human-wildlife conflicts (fig. 4). The general assumption behind this conflict is that raptors compete with hunters for a shared resource, i.e. game species, which could lead to reduced quotas and subsequent economic losses to the hunting lobby (Kenward 1999; Thirgood et al., 2000a; Valkama et al., 2005; Park et al., 2008). Tensions are mainly associated to small-game hunting, an important socio-economic activity in rural areas of developed countries (Grado et al., 2001; Virgos and Travaini, 2005). The perceived (and mostly unfounded, see below) role of birds of prey in depleting hunting resources entails conservation consequences: raptor persecution (e.g., in the form of poisoning or intentional shooting) has been common practice among hunters for decades (e.g., Whitfield et al., 2003), which strongly ignores national and international regulations regarding the protection of raptor populations (e.g., European “Bird Directive” no. 79/409/CEE and “Habitat Directive” no. 92/43/CEE). This is of special concern for the most endangered species, for which continued killing contributed to reduce their populations to the point of local and regional extermination in some cases (Whitfield et al., 2004; Carrete et al., 2007). Unfortunately, species not feeding on game may also suffer from poisoning, as well as from direct killing by indiscriminate shooting (e.g., Villafuerte et al., 1998). Thus, political and social action, as well as scientific attention, has often been encouraged to reconcile raptor conservation and hunting interests (Valkama et al., 2005; Moleón, 2007a; Thirgood and Redpath, 2008).

Assessing the impact of raptor predation on game populations

The key question from a scientific perspective is whether predation exerted by raptors is compensatory or additive to other prey losses (Moleón, 2012). Raptor predation would be compensatory if raptor-induced mortality does not entail a net increase in the mortality rate of the prey population. This may occur when mortality by raptors is rapidly compensated by a parallel reduction of deaths caused by other factors, or by enhanced reproduction, longevity and recruitment of prey species (Krebs, 2002). In this case, preyed individuals are usually of low quality (e.g., ill, injured, young or old) and constitute the so called “doomed surplus” of the population (Errington, 1946). In contrast, raptor predation would be additive if raptor-induced mortality results in a parallel increase of net mortality (Krebs, 2002). However, rather than fully compensatory or additive, predation exerted by raptors is likely characterised by a mixture of both components (Aebischer, 1991).

Fig. 4.

Conflicts between hunters and conservationists are still present and deteriorate because game populations have decreased in many regions of Europe whereas birds of prey populations recover. Booted eagle Aquila pennata with red-legged partridge Alectoris rufa. Credit: Eugenio Martínez-Noguera.

[Los conflictos entre cazadores y conservacionistas son todavía habituales e incluso empeoran porque las poblaciones de especies cinegéticas han declinado en muchas regiones de Europa mientras que las rapaces se recuperaban. Aguililla calzada Aquila pennata con una perdiz roja Alectoris rufa. Crédito: Eugenio Martínez-Noguera.]

f04_181.jpg

But, how can we empirically assess the impact of raptor predation on game prey populations? Predator-removal experiments could provide excellent insights. In practice, however, manipulation of raptor densities is largely prevented due to ethical and logistic reasons, especially on large spatiotemporal scales. An alternative approach consists of calculating the predation impact, i.e., the number of individuals (kill rate) or the proportion of the prey population (predation rate) taken by the predator (e.g., Valkama et al., 2005; Moleón et al., 2012). However, although the assessment of both parameters may be an important basic step, it results insufficient to infer limitation by the studied predator on the prey population because the nature of predation (i.e., what percentage is actually additive) remains unresolved. Answering this crucial question requires a profound knowledge of prey demography (including mortality rates and causes). Predation patterns and demographic information can then be integrated into prey population dynamics models to estimate changes of the hunting bag (which is highly dependent on the hunting form and socio-economical context) in relation to raptor abundance and predation rates (Marcström et al., 1988; Thirgood et al., 2000a). To accurately do this, we should recognise that raptor abundance and predation rates might change according to prey density. Thus, a recommended first step is to estimate the raptor's total response (i.e., how the total prey consumed by the total raptor population varies with changing prey densities), which results from summing the functional response (a function of the per capita intake) and the numerical response (a function of the total raptor population; Gilg et al., 2006). Correlational methods, e.g., comparing prey population with predator population and diet over different years or regions, could also provide an indication of the influence of raptor predation on smallgame prey population dynamics (Moleón et al., 2008, 2013; see Park et al., 2008 for a list of advantages and disadvantages associated with different methods).

Impact of raptor predation on game populations: scientific evidences

Several reviews have dealt with the issue of raptor predation on game populations (Newton, 1993, 1998; Kenward, 1999; Valkama et al., 2005; Park et al., 2008). In Europe, ca. 62% of diurnal and nocturnal raptor species do not include (or do so only occasionally) gamebirds in their diets (Valkama et al., 2005). For the other 20 species, few studies have addressed the limiting potential of raptor predation on game breeding populations, and even less on game pre-harvest numbers, which is the main concern of hunters. The main conclusion that can be taken from these studies is that the impact of raptors on game populations and hunting bags is low in general, with some possible exceptions at the local level for certain raptor-gamebird systems (Valkama et al., 2005; Park et al., 2008). Scientific evidence on the best-known case study, i.e., that represented by hen harrier Circus cyaneus and red grouse Lagopus lagopus scoticus in Scotland (Thirgood et al., 2000a; Thirgood and Redpath, 2008), indicate that: a) harriers show a sigmoidal, type III functional response (typical of generalist predators) to changing red grouse densities, but harrier breeding density is primarily determined by the abundance of passerines and rodents, not red grouse (Redpath and Thirgood, 1999); b) predation by hen harriers can reduce grouse harvests when the former are present at high densities and the latter at low densities, but hardly otherwise (Thirgood et al., 2000b; 2000c), which results in annual commercial losses of ca. £100,000 (estimation for 1996; Redpath and Thirgood, 1997); and c) several science-based management measures, including diversionary feeding, could significantly reduce the conflict (e.g., Redpath et al., 2001). Unfortunately, this profound scientific knowledge has been insufficient to date to reconcile conservationist and hunting interests in Scottish grouse moors because the lack of political willingness to implement solutions and, especially, the entrenched position of game managers, so hen harrier persecution is still widespread there (Thirgood and Redpath, 2008).

What is the state of the art in other systems? Studies in the Mediterranean Basin are scarcer and largely limited to Spain. Here, the system formed by Bonelli's eagle Aquila fasciata and two of its main prey in W Europe, the European rabbit Oryctolagus cuniculus and the red-legged partridge Alectoris rufa (Moleón et al., 2009), has received recent attention. In an area of SE Spain, Bonelli's eagle showed a hyperbolic, type II functional response to rabbit and partridge density changes, thus behaving as a specialist predator (Moleón et al., 2012). Although no numerical response to rabbits or partridges was found in this study, another work considering additional study areas and a longer temporal series found a positive relationship between the proportion of rabbits in the diet and Bonelli’s eagle productivity at the territory scale. On the population scale, greater consumption of rabbits also improved productivity, adult survival and population growth rate (Resano-Mayor et al., 2016). A positive effect between rabbit availability and demographic parameters has been recorded in other diurnal (e.g., golden eagle Aquila crysaetos; Fernández, 1993) and nocturnal (e.g., Eurasian eagle owl Bubo bubo; Martínez and Zuberogoitia, 2001) Iberian raptors.

Mean kill rate per Bonelli's eagle territory (i.e., breeding pair plus fledglings) during the 100-day reproductive season (i.e., when the eagles’ food requirements are higher) was ca. 49 rabbits and ca. 40 partridges in SE Spain. This resulted in a mean predation rate of 7.6–18.5% (rabbit) and 4.6–22.4% (partridge), depending on the considered territory radius of the eagle (Moleón et al., 2012). At the population level, the predation rate by Bonelli's eagle was < 2.5% for both prey, either for the breeding or the non-breeding period. This suggests that predation by Bonelli’s eagles could scarcely affect the number of rabbits and partridges available for shooting in autumn, as well as the partridge population in spring hunted using call lures. In fact, Bonelli's eagle predation rates on partridge were notably lower than those reported for other European raptor-gamebird systems, probably because of the lower population density of Bonelli's eagles due to their greater territories (Moleón et al., 2011a). In an area of NE Spain, Northern goshawk Accipiter gentilis predation was estimated to reduce the partridge huntable stock by 22%, although it was unclear how much of these losses was additive. Moreover, goshawk predation hardly provoked changes in partridge breeding numbers between consecutive years (Mañosa, 1991). Given the lower pair density of large eagles in relation to smaller raptors (Newton, 1979) and the superpredation events by the former on the latter (Sergio and Hiraldo, 2008; Lourenço et al., 2011), small game populations in hunting states with breeding eagles could be subject overall to less predation pressure if compared with those in which eagles have been exterminated (Moleón et al., 2011a).

Although the predation impact by Bonelli's eagle was very low in general, the enhanced predation pressure exerted by all shared predators (i.e., those species preying on both rabbits and partridges) on partridges after rabbit haemorrhagic disease (RHD) had decimated rabbit populations 25 years ago was probably enough to force a subsequent partridge population decline in Spain (Moleón et al., 2008; 2013). The population crash of these important small game species caused by RHD, either directly via infection (rabbit) or indirectly via hyperpredation (partridge), led to increased predator persecution in Spain (e.g., red kites Milvus milvus; Villafuerte et al., 1998).

Finally, it is interesting to note that not all preyed individuals are similar in quality. For instance, Penteriani et al. (2008) found that Eurasian eagle owls prefer to capture substandard rabbits. Also, unpaired sex ratios in prey populations could enhance or buffer the negative effects of raptor predation. In spring, Bonelli's eagle and eagle owl predation on partridges is biased towards males (Donázar and Castién, 1989; Moleón et al., 2011a). However, female partridges are probably easier to capture by terrestrial predators during this period, which could counteract the effects of raptor predation on the sex ratio of partridge populations (Moleón et al., 2011a). In addition, prey can exhibit behavioural responses to avoid excessive predation. In dispersal areas of the Spanish imperial eagle Aquila adalberti in Spain, the continued presence of foraging immature individuals in reduced areas induced changes in the activity and space use pattern of rabbits, which forced dispersing eagles to abandon the area after a few days (Ferrer, 1993).

The main lesson learned after decades of research effort is that determining the impact of raptor predation on small-game populations, with particular emphasis on quantifying the additive component of predation, and offering efficient tools for sport hunting management is a complex issue that cannot be resolved by using simple and partial approaches. We need fine estimations of game bag loss, which could be accompanied by associated economic quantification. However, this is not the last step.

Let us imagine that raptor predation is identified as an important limiting factor of a given game prey population. In such a case, a new key question would arise: can we conclude that removing raptors is the most efficient management measure to guarantee sustained hunting profitability? Apart from legal and ethical considerations, the response is not, because several factors other than predation (mostly, human-induced factors such as habitat degradation, overhunting and the spread of emerging infectious diseases) can exert a much higher limiting effect on the game prey population (e.g., Newton, 1998). It is also pivotal to differentiate between the proximate and ultimate causes of prey declines in order to design the most efficient measures (Moleón et al., 2013). Thus, improved scientific arguments, together with responsible attitudes among policy makers, stakeholders and hunters are indispensable ingredients to reduce human pressure on raptors while maximising hunting profitability.

Finally, we should bear in mind that raptors need healthy prey populations to persist in the wild, while sport hunting in general is not a vital activity for humans. This is not a trivial question, especially in light of the large amounts of money and huge efforts that Europe is nowadays devoting to the conservation of raptors (see above).

Scavengers and livestock

Although human-wildlife conflicts likely have an ancient origin (Anderson, 1997), it is during the recent decades, coinciding with the increase of the human population and the need to give value to natural resources, that the confrontations between wildlife and humans are becoming habitual, increasing the interest of environmentalists, managers and policy-makers (Treves et al., 2009; Gehring et al., 2010). Among the conflicts between wildlife and humans, the commonest scenario is related to the coexistence between herbivores of economic importance and threatened predators (mainly carnivores and raptors) (Peterson et al., 2010; Redpath et al., 2013, Redpath et al., 2015). However, when the conflict arises in an unexpected scenario as, for example, with species not considered a menace for human interests, the management actions to solve this interaction are unknown and difficult to implement. This is the case of the attribution of attacks of Eurasian griffon vultures on livestock, an apparent emerging conflict that was identified in southern Europe during the 1990s and which has increased progressively during the subsequent decade (Margalida et al., 2011; Margalida et al., 2014). This conflict causes major unrest among farmers but has received scarce attention from the scientific community (Margalida et al., 2011; Margalida et al., 2014). The opportunistic killing of small and medium-sized vertebrates has been reported as relatively frequent in some New World vulture species (Lowney, 1999; Avery and Cummings, 2004) but has only been reported occasionally in the Old World vultures (Houston, 1994).

The temporal context of the conflict

The first reports of vultures attacking livestock were documented in northern Spain in the middle of 1990s (Navarra, western Pyrenees) but it was not until middle of the 2000s when the conflict was generalised across Spain and southern France (see the review in Margalida et al., 2014). In general, most of the presumed attacks took place in spring (from April to June), coinciding with birthing dates in areas of extensively reared livestock (mainly sheep and cows that constitute 80% of the livestock affected, see Margalida and Campión, 2009; Margalida et al., 2014). Analysing the temporal pattern, most of the complaints took place from 2006 onwards, coinciding with the food shortages provoked by the sanitary policies after the bovine spongiform encephalopathy (BSE) (Donázar et al., 2009; Margalida et al., 2010). This facilitated that the general social perception of the problem, also misinterpreted by the media and people, attributed immediately a causal relationship between food shortages and vulture behaviour. Indeed, the sudden trophic reduction provoked some shifts in the diet (Donázar et al., 2010), behaviour (Zuberogoitia et al., 2010), demographic parameters (Margalida et al., 2014) and foraging movements of vultures. However, some preliminary results suggested that there is no relation between the frequency of complaints of alleged attacks and both the spatial availability of trophic resources and the density of vulture populations (Margalida and Campión, 2009). To support this, it must also be taken into account that many complaints took place before food shortage regulations (1996–2006, see Margalida et al., 2014). Thus, the link between “vulture attacks” and a reduction in food availability is unclear, at least with respect to the breeding population.

A magnification of the problem

Despite the relatively low economic cost of the conflict (Margalida et al., 2014), the social and media impact has been substantial. This provoked pressure on administrations that were obliged to explore the problem with more detail. Figure 5 shows the trend of the between 2006 and 2015. As we can see from the figure, this trend increased suddenly from 2006 to 2010, then diminished from 2011 onwards (fig. 5) coinciding with the approval and application of new and more flexible regulations that allowed access to more food resources by avian scavengers (Margalida et al., 2012). However, the real application was effective after 2013 when most of the administrations returned to a pre-BSE scenario, providing more natural food (livestock losses) and creating feeding stations for vultures. Thus, probably with the approval of new regulations, the social alarm diminished and the importance of the conflict showed the regressive trend observed. However, as occurs with the previous hypothesis that “food shortages” equals “increase of attacks”, other explanations are possible. It seems that after the important increase in the number of complaints between 2006 and 2010, a more rigorous control and assessment of the complaints by administrations increased the percentage of rejection. For example, in Catalonia, as a case model that applied specific and rigorous controls during 2010–2015, the complaints compensated by administration decreased from 40–62% during 2008–2010 to only 2–9% during 2011–2015. Thus, a careful monitoring showed that the problem was overestimated and most of the complaints were independent of the vultures' behaviour in the scenario of the problem. In this regard, some preliminary conclusions obtained during in situ visits to disentangle the true factors explaining the presumed attack were obtained. Firstly, changes in husbandry practices in areas in which livestock range freely or remain unattended in large fenced enclosures even at the time of lambing has been considered as one of the main factors. These circumstances increase the chance of predation by other animals such as common ravens Corvus corax, which have been considered responsible of several aggressions. Secondly, individuals were mainly affected during calving and most of them presented problems (young individuals with first birth, breech birth…). Finally, in most cases, the complaint is plainly argued by showing pictures with the rests of the livestock presumably attacked (skin debris and bone remains), but with direct observation or additional evidence, just relying on the only affirmation that “vultures were present in the livestock carrion”.

Fig. 5.

Trend in the number of complaints related with European griffon vulture attacks to livestock in Northern Spain (Catalonia, Aragón, Guipuzcoa and Álava) during the period 1996–2015.

[Tendencia en el número de reclamaciones relacionadas con ataques de buitres leonados a ganado en el norte de España (Cataluña, Aragón, Guipúzcoa y Álava) durante el periodo 1996–2015.]

f05_181.jpg

Raptors and human infrastructures

Transport and energy infrastructures are growing exponentially worldwide. Their increases derive from such human demand of resources and are also pointed out as drivers of global change (Vitousek et al., 1997). The development of infrastructures has been exponential in Mediterranean Europe during the last decades as a result of the growing financial support of EU development funds. For example, France has the largest network of roads (106 km) and Spain the largest network of motorways (14,000 km) in Europe ( http://www.irfnet.eu/images/Statistics/ER_Statistics_Final_2012.pdf). Thus, the major socioeconomic changes in the Mediterranean are driving a shift in the mortality factors affecting birds of prey from direct persecution to casualties associated to infrastructures (Martínez-Abraín et al., 2009) (fig. 1).

Transportation: Roads, railways and airports

The increasing network of roads for the transportation of people and goods has important effects on habitats, mainly through fragmentation (Forman and Alexander, 1998; Trombulak and Frissel, 2000; Meunier et al., 2000) but also on the behaviour and survival of raptors and owls (Bautista et al., 2004; Lambertucci et al., 2009). Although other terrestrial vertebrates such as reptiles and mammals suffer higher road mortality rates (Glista et al., 2007), some diurnal raptors and owls might also follow population declines because of reduced survival. For example, in Mediterranean habitats two declining owl species, barn owls and little owls Athene noctua, might be particularly sensitive to road casualties (Hernández, 1988; Frías, 1999; Fajardo, 2001). In contrast, different raptor species such as common kestrels Falco tinnunculus, black kites Milvus migrans or common buzzards Buteo buteo are attracted to road verges that provide with food resources (small mammals and rabbits) and perches (Meunier et al., 2000), but nevertheless these species as well as other forest-dwelling raptors suffer a significantly lower vehicle collisions (Zuberogoitia et al., 2015). On the other hand, increasing traffic load has been described as a factor influencing foraging behaviour of raptors affecting particularly large and endangered species such as the Spanish imperial eagle and cinereous Aegypius monachus and griffon vultures that decrease their activity close to roads on weekends (Bautista et al., 2004). Surprisingly although the Mediterranean region holds dense railway networks including the largest high-speed railway network in the world (UIC, 2014), there are no published studies on its potential effects on raptors.

Interactions between birds of prey and airports have been mainly assessed from the perspective of bird collision with aircrafts. Although raptors only constitute a minor percentage of the airfield's avifauna, this group caused the majority of fatal accidents in the Mediterranean region (Thorpe, 2005; Kitowski, 2011). Due to its large size and low flight maneuverability, griffon vulture was the species that caused the most serious accidents (Kitowski, 2011). Other species involved were Eurasian kestrels, common buzzards, marsh harriers Circus aeruginosus or Montagu's harriers C. pygargus. Among owls, little owls are particularly sensitive to this type of of accidents, being involved in half of all the bird strikes reported in some airports (Kitowski, 2011). Most of the bird strike reports remain unavaible to the public due to airport policies which prevent the study of their incidence. Despite this, the relative importance of mortality from collision with aircraft seems likely to be low, but precise studies focussing on its effect on bird of prey populations are needed. Mitigation measures for bird of prey strikes included vegetation management and habitat modification to make airfield less desirable for their prey species (ICAO, 2012; De Vault et al., 2013) and real time radar surveillance to monitor bird movements and manage air traffic (Leshem et al., 2010). So far, the effectiveness of other widely used methods as bird deterrents or falconry has not been properly assessed. In recent years, some hazard bird strike models based on the structure of landscape, abundance and sensitivity of bird species or aircraft movements have been implemented (Soldatini et al., 2010; Coccon et al., 2015). This could be an effective tool to evaluate and manage the risk of bird collision with aircrafts.

Energy production: power lines, wind farms and solar facilities

Electrocution and collision with power lines are among the main causes of population declines for many bird species (Lehman et al., 2007) and are considered a major source of human induced mortality for raptors in Southern Europe since the pioneering studies in late 1980's (Ferrer et al., 1991; Ferrer and Negro, 1992; Rubolini et al., 2005). In fact, electrocution might be considered as the major modern source of non-natural mortality for an array of medium and large-sized birds of prey such as the Spanish imperial eagle (González et al., 2007; López-López et al., 2011), Bonelli's eagle (Real et al., 2001; Hernández-Matías et al., 2015), booted Aquila pennata and short-toed snake Circaetus gallicus eagles (Martínez et al., 2016) or the Eurasian eagle owl (Sergio et al., 2004; Martínez et al., 2006). Retrofitting of power lines has been proved to reduce mortality and improve population viability of endangered species (López-López et al., 2011; Chevallier et al., 2015).

Wind farms are also a major source of mortality for birds and bats worldwide (Sánchez-Zapata et al., 2016). In Mediterranean habitats there is increasing concern on the demographic consequences of collisions in wind farms for long lived raptors such as Egyptian Neophron percnopterus, cinereous and griffon vultures (Carrete et al., 2009, 2012; Sanz-Aguilar et al., 2015; Vasilakis et al., 2016). Other raptor species such as short-toed snake eagles and common kestrels are also frequently found dead in wind farms although their effects on population dynamics are unknown.

Solar facilities are relatively new and their effects on biodiversity have been scarcely documented (DeVault et al., 2014). On a local scale, impacts are mostly associated with habitat transformation and to a lesser extent with wildlife mortality (Lovich and Ennen, 2011; Hernández et al., 2014). There are no detailed published results on the effects of these infrastructures on raptors in Mediterranean habitats although steppe raptors such as Montagu’s harriers and lesser kestrels Falco naumanni are more likely to be affected because of the spatial coincidence of these infrastructures with their breeding and foraging habitats.

Birds of prey in novel habitats: adaptations and consequences of living in cities

Habitat alteration and loss are one of the major threats for biodiversity and unfortunately their effects are reaching a global scale (see i.e. Donald et al., 2013). Nonetheless, the same processes that are causing the habitat loss are simultaneously originating novel ecosystems, defined as ecosystems created deliberately or inadvertently by human activity that show a species composition and abundance previously unknown in that bioma (Hobbs et al., 2006). Urbanization involves one of the most extreme forms of landscape change, since natural habitats are fragmented, isolated and finally lost by urban development at varying spatial scales, which generally lead to a deep restructuring of affected communities (Marzluff et al., 2001; Alberti, 2005; González-Oreja, 2011). Species diversity is lower in animal communities inhabiting urban landscapes, although human habitation offers a variety of intentional or incidental subsidies to wildlife, such as provision of supplementary foods (Fuller et al., 2008); and some well-adapted species opportunistically exploit such resources and reach high abundances. These high densities of potential prey attract birds of prey to hunt and even to breed in urban environments (Chace and Walsh, 2006; Rutz, 2008; Solonen and Ursin, 2008). While it is evident that urbanization is causing the restructuration of faunal assemblages, the effect of urbanization on individual species is more complex than initially perceived (González-Oreja, 2011; Isaac et al., 2014).

High prey concentrations act as powerful attraction for predators, mainly raptors and owls, which are able to occupy surrounding natural areas and move into urban areas for hunting (Zuberogoitia et al., 2002). Raptors breeding near urban areas largely increase the consumption of certain species included in the range of optimal prey mass, which positively affect breeding success and brood size (Drewitt and Dixon, 2008; Zuberogoitia et al., 2013; Martínez-Hesterkamp, 2015). The abundance of feeding resources also attracts floaters, wintering and non-territorial birds, which directly occupy cities (Pirovano et al., 2000; Cade and Burnham, 2003; Isaac et al., 2014). However, urban environments may not contain the full complement of resources required by a species (Chace and Walsh, 2006). Nest site availability is one of the main limiting factors in urban areas and those that already exist are usually of poor quality (Altwegg et al., 2014). One of the consequences may be the establishment of long-term non-breeding territories, which would act as ecological traps (see Battin, 2004; Isaac et al., 2014; Fasciolo et al., 2016), ultimately reducing fitness components (Schlaepfer et al., 2002; Remes, 2003).

In some cases, birds of prey are able to find nesting resources in cities, because they reuse stick-platforms of corvids, occupy old structures or artificial ledges, or are intentionally favoured by setting nest-boxes (Cade et al., 1996; Ranazzi et al., 2000; Solonen and Ursin, 2008; Fiuczynski, 2011; Zuberogoitia, 2011; Lövy and Riegert, 2013; Rodríguez et al., 2013; Altwegg et al., 2014). Some raptor species produce fewer eggs and fledglings in urban habitats than they do in rural ones (Tella et al., 1996). However, other species respond with higher reproductive success (Bird et al., 1996) (fig. 6). In urban environments, breeding tawny owls Strix a lu co seem to be relatively free from the pronounced regional abundance fluctuations of small mammals, and probably gain from the relatively stable food resources (such as rats and common pigeons Columba livia), as well as the warmer microclimate of towns compared to rural habitats that largely govern the breeding of owls elsewhere (Solonen and Ursin, 2008). The abundance and stability of prey species and the behavioural adjustments to human disturbance result in higher breeding success and productivity in urban goshawks Accipiter gentilis and sparrowhawks Accipiter nissus (Papp, 2011; Solonen, 2014a). However, some breeding urban raptors, mainly cliff-nesting species, suffer conflicts with the human occupants of the building on which the birds have decided to breed and are expelled as a consequenc, reducing their reproductive performance (Atwegg et al., 2014).

Home ranges of urban raptors and owls are larger than those monitored in natural areas, because they probably enlarge territories for exploiting suitable hunting grounds, which are sparsely distributed within urbanised areas and surrounded by a high proportion of unused developed areas (Tella et al., 1996; Henrioux, 2000; Riegert et al., 2007; Lövy and Riegert, 2013). However, this is not a regular rule for every species; avian forest-dwelling predators seem to have lower home ranges in cities than non-urban con-specifics, possibly due to high concentrations of prey species, mainly birds, of optimal size in parks (Bloom and McCrary, 1996; Warkentin and Oliphant, 1990; McGrady, 1991; Sodhi and Oliphant, 1992; Rutz, 2006; Fiuczynski, 2011; Sulkava et al., 2014). Moreover, urban raptors exhibit higher levels of hunting success and lower flight activity than non-urban ones (Rutz, 2006).

Fig. 6.

Peregrine falcon Falco peregrinus breeding in a building in Valladolid (central Spain) where it preys mainly on domestic pigeons Columba livia. Urban raptors may have comparatively higher breeding success in relation to rural areas thanks to the exploitation of locally abundant food resources. Credit: Juan Sagardía.

[Halcón peregrino Falco peregrinus nidificando en un edificio de Valladolid (España central) donde preda fundamentalmente sobre palomas domésticas Columba livia. Las rapaces urbanas pueden tener éxitos reproductores comparativamente altos en relación con zonas rurales gracias a la explotación de recursos localmente abundantes. Crédito: Juan Sagardía.]

f06_181.jpg

Birds born in urban environments seem prone to reoccupy these habitats; even secretive species such as goshawk and sparrowhawks show a remarkable tolerance to human disturbance (Rutz, 2003; Papp, 2011). However, fear of humans remains highly constant throughout the adult lifespan of a bird species (i.e. burrowing owl, Athene cunicularia; Carrete and Tella, 2013) and there is strong individual consistency in the response to disturbances (Carrete and Tella, 2010). Thus, it is more likely that a permanent perturbation factor like human disturbance would be selecting those individuals that are able to better tolerate humans (Rebolo-Ifrán et al., 2015). Therefore, differences among urban and rural populations are more probably a consequence of selective pressures (Carrete and Tella, 2013). Urbanised areas may be selecting individuals with adaptive abilities (so called behavioural syndromes or personalities) which are able to deal with urban challenges, human presence included (Rebolo-Ifrán et al., 2015). The contribution of captive bred birds in reintroduction programmes may also affect individual behavioural adjustments and favour changes in the relationships between raptors and humans in cities (Cade and Burnham, 2003; Cugnasse, 2004).

From a negative point of view, urban raptors are frequently affected by risks associated with vehicle and window collisions as well as electrocutions (Chace and Walsh, 2006; Hager, 2009), by premature fledging and by direct persecution (Cade and Bird, 1990). Riegert et al. (2010) conclude that the apparent advantage of suitable nesting sites for common kestrels in the city centre can be counterbalanced by new possible risks, such as threats from humans. Moreover, the high population densities of some prey species that positively affect raptors can elevate contact rates within and among species, and favour the spread of parasites by direct contact or oral-fecal routes (Bradley and Altizer, 2006). Increased interspecific competition in some reduced urban areas (i.e. foraging and resting sites) has been linked to chronic stress, which can lower resistance to infection and intensify the harmful effects of pathogens through effects on the host immune system (Sol et al., 1998; Padgett and Glaser, 2003). Furthermore, some heavy metal and pesticide pollutants become concentrated in the surroundings of developed areas and can be highly detrimental to vertebrate health (Bradley and Altizer, 2006) and adversely affect urban raptors (e.g. Newsome et al., 2010).

Overall, the combined effect of these various factors in many modem cities seems to be positive for peregrine falcons Falco peregrinus, goshawks, sparrowhawks, long-eared owls Asio otus, tawny owls and others (Cade et al., 1996; Rutz, 2008; Solonen, 2014a, b; Lövy and Riegert, 2013). However, there are still open questions related to the medium- and long-term consequences of living in urban habitats. On one hand, it is not clear that the urban population may maintain a stable trend without immigration. On the contrary, demographic data from a long-term monitored population of urban peregrines suggest that the population would have declined in the absence of immigration (Atwegg et al., 2014). Finally, reproductive performance is high in most species, but further research is needed in order to assess the effects on survival and long-term inter-generational fitness. Lövy and Riegert (2013) also suggested that it would be of interest to clarify raptors' use of habitat networks in urban environments, i.e. how fragmentation and/or connectivity of focal habitats influence patterns of birds' movement.

The challenge of rewilding for raptors

As we have previously emphasised throughout the text, besides direct and indirect persecution, degradation and the loss of natural habitats remain as one of the major threats for raptors conservation worldwide. Nonetheless, the projections on land-use changes modelled for Europe reveal a different picture (Pereira et al., 2010). After centuries of deforestation, the continent has followed a trend of land abandonment and consequently an increase of scrub and forested areas (Kaplan et al., 2009; van Vuuren et al., 2006). Although frequently linked to the idea of releasing charismatic species (see details in Deinet et al., 2013; Stokstad, 2015) a “rewilding” process is more properly understood as the passive management of ecological succession through the reduction of human control of landscapes. This encompasses opportunities to restore natural ecological processes, as well as species and the ecosystems services that might be provided (Gillson et al., 2011; Navarro and Pereira, 2012; Pereira and Navarro, 2015). During the last decade this issue is motivating a great deal of interest among scientists and policy makers who are trying to predict the response of species and the capacity to adapt to new environments on the basis of the different projections of socio-economic alternative scenarios.

According to Navarro and Pereira (2012), rewilding triggers profound biodiversity changes with some loser and winner species (see below). In this context raptors are not an exception. The study of the effects of land-use abandonment and its consequences on the viability of the ecosystems (and the species inhabiting there) is a challenge in environmental sciences. On the one hand, rewilding processes reduce human presence in many areas thus increasing the availability of suitable habitat for those species being historically persecuted (Enserik and Vogel, 2006; and see above). Conversely, farmland abandonment may have negative consequences on species historically linked to traditional agro-grazing exploitations and rural constructions (Fuller, 1987; Lab aune and Magnin, 2002; Laiolo et al., 2004; see below). Within this context, European raptors are a paradigmatic study case due to their position within food webs as top predators and scavengers. However, there is little knowledge on how raptors would face rewilding. Focusing on carrion-eaters, it seems likely that land-abandoment process may affect important fractions of the populations of vultures living in the Iberian Peninsula in the coming decades (fig. 7; see also García-Barón, 2014). It has been argued that rewilding may favour the availability of carrion for large avian scavengers thanks to both the expansion of wild ungulates (and their abundances) and the recovery of large predators. Both issues would represent an increase in the availability of carcasses in the wild (Selva et al., 2003; Wilmers and Post, 2006). However, although rewilding occurs in some mountain ranges of southern Europe it must be taken into account that large scavengers forage over huge areas so they may be still very dependent on live-stock carcasses and food supply at artificial feeding stations (Margalida et al., 2011).

Many raptor species (especially large eagles) and some scavengers like the cinereous (Aegypius monachus) and Egyptian vultures are very dependent on wild rabbits, a keystone species of the Mediterranean biome (Delibes-Mateos et al., 2008). Rewilding future scenarios predict the dominance of closed forest and scrubland undermining this lagomorph which prefers mixed habitats of scrubs-forest and open grassland for concealment and feeding (see references in Smith and Boyer, 2008). Thus, García-Barón (2014) showed that up to 50% of the breeding and potential foraging areas for cinereous vultures in the Iberian Peninsula may be affected by predicted land abandonment under variable socio-economic scenarios. Conversely, and because the rewilding processes would favour the size and the number of patches of woodland, a higher number of old and tall trees would also increase and thus enhance the availability of suitable breeding areas for forest-breeding raptors (Zuberogoitia et al., 2013).

Fig. 7.

Hotspots of land abandonment (in green) vs. the distribution of the breeding areas of four vultures (griffon, Egyptian, cinereous and bearded) and two eagles (Bonelli's and golden) based on information available in the Spanish Atlas of Breeding Birds (del Moral and Martí, 2004). The map in the centre shows areas categorized as “agriculture” in 2000 that are projected to become abandoned or afforested in 2030 (Verburg and Overmars, 2009). (Species drawings: Juan Varela). Figure modified from Cortés-Avizanda, et al. (2015).

[Mapas de abandono de usos del suelo (en verde) vs la distribución de las zonas de reproducción de 1os cuatro buitres ibéricos (alimoche, quebrantahuesos, buitre negro y leonado) y dos águilas (perdicera y real) basados en la información disponible en el Atlas Español de Aves Reproductoras (del Moral y Martí, 2004). El mapa en el centro muestra las áreas categorizadas como “agricultura” en el año 2000 que se prevé que se convertirán en abandonadas o de expansión forestal para 2030 (Verburg y Overmars, 2009). (Dibujos de especies: Juan Varela). Figura modificada a partir de Cortés-Avizanda, et al. (2015).]

f07_181.jpg

In general, it can be expected that land abandonment and the subsequent changes in the abundance and distribution of resources will have asymmetrical effects on birds of prey (and other organism) populations. How these processes can shape the structure of raptor assemblages and communities remains an outstanding question. Besides, diffuse competition (Bock et al., 1992; Piana and Stuart, 2012), a scarcely-studied process, competitive exclusion may determine that most aggressive and/or social species can displace other birds of prey from preferred breeding areas (Krüger, 2002; Sergio et al., 2005) although high variability can be expected depending on site-specific contexts (see e.g. Treynis et al., 2011). Also, when a group of species exploits simultaneously the same food resource, as happens with avian scavengers, monopolisation by dominants (e.g., Eurasian griffon vultures) may appear when their relative abundance increases (Donázar et al., 2010; see also Carrete et al., 2010), and negative effects on populations of the subordinate avian scavenger species, as well as on other organisms, are predicted (De Vault et al., 2003; Shivik, 2006; Selva and Cortés-Avizanda, 2009; Cortés-Avizanda et al., 2012). Finally, intraguild predation (the killing of species using the same resources) appears as a generalised phenomenon having potential consequences at individual and population levels, but also having the potential of shaping predator communities (Sergio and Hiraldo, 2008). For example, the spread of forests may benefit the larger body-sized tawny owl in detriment of smaller species as barn and little owls which are more associated with open areas (Mikkola, 1983).

The counterpart of land-use abandonment is the intensification of agricultural practices in much more cost-efficient other areas. This fact leads to an increase of contrasting landscapes of large surfaces covered by dense forests and scrublands and intensively cultivated crops. These changes in the landscape are not new but has accelerated in recent decades when the process also directly affects many birds of prey species linked to environments under agro-grazing traditional practices which are not as hostile (Donázar et al., 1996). The abandonment of traditional practices may negatively affect species such as the threatened lesser kestrel because of the disappearance of older, rural buildings for breeding. Similarly, great interest has arisen in the study of the effects of rewilding on the abundance of arthropods (an important item in the diet of these and other openhabitat birds) where population fluctuations could be largely influenced by both the modification of land uses.

In general, the abandonment of the rural areas would reduce human presence and thereby relax human disturbance, thus enabling breeding success of raptors as well as modifying their feeding or foraging behaviours (see review in Martínez-Abraín et al., 2009, 2010). However, the abandonment of traditional agro-grazing activities does not necessarily mean the disappearance of all forms of humanisation and their impact. In fact, those wilder areas would become of interest for recreational activities which are currently considered as an increasing concerning topic in wildlife conservation (Martínez-Abraín et al., 2010). In addition, after the abandonment of the land, owners can search profitability by installing certain industries, such as wind farms that can have serious consequences on populations of raptors, other birds and bats (Sánchez-Zapata et al., 2016). Consequently, it seems erroneous to establish a linear relationship between land abandonment and human intervention on landscapes and life assemblages.

Future avenues

The paradigm of nature protection through the prioritisation of threatened or flagship species has had valuable results in Europe over the last 40 years benefiting specifically most birds of prey species occupying the continent. Thus, most of the emblematic taxa have experienced spectacular recoveries so that they have substantially expanded both the areas of distribution and population sizes. However, it is necessary to advance new approaches which improve the conservation status of both birds of prey as well as other elements of biodiversity. Increasing awareness of general public and policy-makers in this subject would overcome the dilemma between protected vs. non-protected territories, improving the capacity for sectorial policies integration —especially the agricultural ones (Hodge et al., 2015)—, the mobilisation of financial resources and the achievement of the objectives derived from the international legal commitments on wildlife protection and sustainable development.

It seems desirable to highlight the role of raptors as ecosystem services providers with clear benefits to human welfare and ecosystems sustainability (Whelan et al., 2015). It is increasingly acknowledged among scientists that avian scavengers yield important regulatory and cultural services but its quantification remains to be addressed in detail. In addition, studies aimed at the mapping of ecosystem services on a large scale would be very helpful (Naidoo et al., 2008). Moreover, additional research about supporting services (i.e., nutrient cycling) provided by scavengers is necessary. Studies about the perception of stakeholders (i.e., shepherds) and evaluation of the ecosystem services and providers (i.e., vultures and facultative scavengers) would be essential for the establishment of management proposals for the conservation of the scavenger populations and the maintenance of the services provided. Also, higher social recognition of these benefits seems desirable (Daily et al., 2009). This is particularly true for Spain where the bulk (> 90%) of the European vulture populations are found (Margalida et al., 2010).

Birds of prey also provide regulating services through pest control (chiefly rodents) in agricultural systems. There is however, a striking paucity of studies in this field, mainly because reliable investigation needs long-term experimental approaches in large predator-excluded and control areas with replicates, which is time- and personneldemanding and expensive, thus unaffordable for local small research groups. IPM (integrated pest management) and BCR (biological control by raptors) implementation requires the involvement of regional and national administrations to provide areas, personnel and funds for investigation on these promising, non-aggressive, cheap and sustainable procedures of agricultural pests more and more demanded by the public.

Increasing populations of birds of prey may result in conflicts with human interests. Perhaps the most entrenched is that arising from the predation on game species which often results in direct killing and indirect persecution (poisoning) (Villafuerte et al., 1998). Available scientific evidence indicates that raptors are rarely able to limit game prey populations (see reviews in Valkama et al., 2005; Park et al., 2008). However, current scientific knowledge on this topic is still very scarce and suffers from serious biases that prevent inferring robust conclusions. The methodological complexity (including considerable field effort) of determining the impact of predation on game populations, which increases in complex ecosystems, probably deters further advance in this field (Valkama et al., 2005; Park et al., 2008; Moleón, 2012). Future research is particularly desirable in Mediterranean regions, which combine high hunting pressure and important raptor populations (Valkama et al., 2005; Moleón et al., 2011a). Studies that explore simultaneously the raptor functional and numerical responses are rare (Valkama et al., 2005; Moleón et al., 2012), as well as multi-scale (e.g., population plus territories within the population) approaches of predation rates (Moleón, 2007b). Also, mammalian game species, which can be highly valuable for hunters (e.g. European rabbit is the most frequently killed game species in Spain; Moleón et al., 2011a) have received much less attention than gamebirds (e.g., Newton, 1993, 1998; Kenward, 1999; Valkama et al., 2005; Park et al., 2008; Thirgood and Redpath, 2008). Moreover, studies including the floating segment of the raptor population are virtually absent. This is a relevant gap in the case of species such as large eagles because floaters concentrate in large numbers in juvenile dispersal areas during the non-breeding period (e.g., Moleón et al., 2011b), which normally coincides with the main hunting period. Floater persecution in such areas and period is elevated, which may result in negative demographic effects for the raptor whole raptor population (Penteriani et al., 2008). The impact of raptor-caused mortality on game species releases for population reinforcement (Kenward et al., 2001), as well as the effect of raptors on game prey behaviour (Ferrer, 1993), also provide ample room for further research. Finally, multispecies approaches, preferably considering the whole raptor (and mammalian carnivore) community, are highly encouraged against one predator-one prey approaches (Valkama et al., 2005).

Damages blamed on raptors are not limited to game species but also reach the binomial vultures-livestock. This emergent conflict is widely publicised and magnified by the broadcast media (Margalida et al., 2014) and its mitigation requires dialogue between scientists, farmers, managers, policy-makers and disseminators. The lack of scientific data and the role of the media increased the social alarm and political pressures to take management decisions not based on scientific evidence. Practitioners generally have centred their attention on reducing negative interactions, rather than on increasing positive relations between humans and wildlife (Frank, 2016). In the case of griffon vultures, taking into account the important services provided by these species (Margalida and Colomer, 2012; Morales-Reyes et al., 2015) and the lack of sound evidences of a true and well defined problem, we need to face the issue under this angle to reduce the uncertainty of an overestimated conflict.

Conflicts between raptors and human interests are not uni- but bidirectional. It is largely known that raptors may be killed in power lines but new infrastructures such as wind farms charge a growing toll on birds of prey populations. Adequate planning and adoption of mitigation measures should be, in this order, the main strategies aimed to minimize these undesirable effects. Fortunately, raptors are among the best studied group of vertebrates in Europe and elsewhere with detailed information on the distribution and status for most of the species (Sánchez-Zapata, 2012). Besides, some of them (particularly large eagles and scavengers) have been long-term monitored so population viability analyses could be applied to evaluate the effects of infrastructures (Hernández-Matías et al., 2013; Chevallier et al., 2015; Sanz-Aguilar et al., 2015). Raptors might be used as indicators and/or umbrella species useful for both for planning new infrastructures and for evaluating and managing mitigation measures (Moleón et al., 2007; Pérez-García et al., 2011, 2016). Further research is required on the impact of emergent infrastructures such as wind and solar facilities, on the conflicts with transportation (particularly air and rail) but also on the effectiveness of the mitigation measures undertaken.

One of the most challenging prospective questions is how birds, and raptors in particular, will be affected by the impending changes in the European landscapes. Because novel ecosystems, (like large cities and conurbations) result from human actions, management is required to guide their development. How we manage these new ecosystems effectively is a point for debate: what should the goals be and how should these systems fit with other systems along the wildintensively managed gradient (Atwegg et al., 2014)? What is even more interesting, does a new combination of species maintain similar functional properties with respect to the old species pool? Do they alter the original network of mutualistic and antagonistic interactions, and what are the consequences for community organisation (Hobbs et al., 2006)? On the other hand, concentration of humans in urban areas leads to the progressive depopulation of rural regions and the loss of traditional agro-grazing systems. In a Europe under severe global change, land abandonment might be an opportunity to preserve specific habitats and therefore to conserve specific organisms and ecological processes.

Overall, a multidisciplinary approach is required in order to get further understandings on the responses of raptors to different land-abandonment “rewilding” scenarios, to define new research avenues, addressing the costs and benefits and developing monitoring schemes within those future divergent landscapes. Future perspectives should rely on to couple the numerous studies about the biology, habitat selection, diet, concerns, etc. (as well as the historical information accumulated on raptor distribution and abundance) with the socio-economic projections on future land-uses which, from our point of view, will allow researchers and policy managers to know how populations and species fit into new habitats and landscapes. Indeed, we consider that those studies would facilitate the establishment of future conservation measures guarantying the preservation of raptors and their relationship with humans.

At the end, and it can not be otherwise, advances in the conservation of birds of prey within the context of current global change can hardly be dissociated from ecological knowledge. As in many areas of science, important conceptual advances in animal ecology have been stimulated by technological innovations combined with a parallel increase of statistical and mathematical methods tools (Cagnacci et al., 2010; Borger, 2016). Modern technologies not only allow collecting data on precise location of individual animals on the planet, but a plethora of sensors can be used to collect information on physiology and behaviour, and animals themselves can be used as sampling platforms to measure external variables as they move (Wilmers et al., 2015). Bio-logging innovations and practical analytical methods are increasingly formalising the close relationship between individual behaviour, habitat selection, species abundance, and population dynamics (Morales et al., 2010; Matthiopoulos et al., 2015; Van Moorter, 2015). This integrated and comprehensive framework will fuel the incorporation of new ideas and theories relating within- and between-individual variation to (meta) population-level consequences. For example, studies on vertebrates have shown the great importance of consistent individual differences in key personality traits such as boldness or aggressiveness on immediate survival prospects and dispersal propensity (e.g. Dingemanse et al., 2003; Bremner-Harrison et al., 2004), and how the coupling of some of these components have far-reaching consequences for range expansions and the colonisation of novel environments (Duckworth et al., 2007). This kind of integrative approach is of crucial importance for understanding eco-evolutionary dynamics at all scales from individuals to (meta) populations, and probably is the most fertile ground for future research in raptor ecology.

Acknowledgements.

M. M. acknowledges financial support through the Severo Ochoa Program for Centres of Excellence in R+D+I (SEV-2012-0262), Z. M. R. was supported by a pre-doctoral grant FPU12/00823. A. M. was supported by a Ramón y Cajal research contract (Ministry of Economy and Competitiveness, RYC-2012-11867). Juan A. Amat and an anonymous reviever greatly improved the original version of the manuscript. ACA was supported by a Post-Doctoral grant from the Fundação para a Ciência e a Tecnologia (FCT) (SFRH / BPD / 91609 / 2012) and a contract Juan de la Cierva Incorporación IJCI-2014-20744 of the Ministry of Economy and Competitiveness. JAD, JMPG, AM, ZM and JASZ were funded by project CGL2015-66966-C2-1-2-R, MINECO/FEDER, UE. Andy Paterson kindly revised the English.

Bibliography

1.

Abadi, F. , Giménez, O. , Ullrich, B. , Arlettaz, R. and Schaub, M. 2010. Estimation of immigration rate using integrated population models. Journal of Applied Ecology , 47: 393–400. Google Scholar

2.

Aebischer, N. 1991. Sustainable yields: gamebirds as a harvestable resource. Gibier Faune Sauvage , 8: 335–351. Google Scholar

3.

Alberti, M. 2005. The effects of urban patterns on ecosystem function. International Regional Sciencific Review , 28: 168–192. Google Scholar

4.

Alcaide, M. , Serrano, D. , Tella, J. L. and Negro, J. J. 2009. Strong philopatry derived from capture-recapture records does not lead to fine-scale genetic differentiation in lesser kestrels. Journal of Animal Ecology , 78: 468– 475. Google Scholar

5.

Altwegg, R. , Jenkins, A. and Abadi, F. 2014. Nestboxes and immigration drive the growth of an urban peregrine falcon Falco peregrinus population. Ibis , 156: 107–115. Google Scholar

6.

Altwegg, R. , Roulin, A. , Kestenholz, M. and Jenni, L. 2006. Demographic effects of extreme winter in the barn owls. Oecologia , 149: 44–51. Google Scholar

7.

Anderson, K. 1997. A walk on the wild side: a critical geography of domestication. Progress in Human Geography , 21: 463–485. Google Scholar

8.

Askham, L. R. 1990. Effect of Artificial Perches and Nests in ttracting Raptors to Orchards. Proceedings of the 14h Vertebrate Pest Conference. University of California. Davis. Google Scholar

9.

Avery, M. L. and Cummings, J. L. 2004. Livestock depredations by Black Vultures and Golden Eagles. Sheep & Goat Research Journal , 19: 58–63. Google Scholar

10.

Bakaloudis, D. D. , Vlachos, C. G. and Holloway, G. J. 1998. Habitat use by Shorttoed Eagles Circaetus gallicus and their reptilian prey during the breeding season in Dadia Forest (northeastern Greece). Journal of Applied Ecology , 35: 821–828. Google Scholar

11.

Balbontín, J. 2005. Identifying suitable habitat for dispersal in Bonelli's eagle: an important issue in halting its decline in Europe. Biological Conservation , 126: 74–83. Google Scholar

12.

Balbontín, J. and Ferrer, M. 2008 Densitydependence by habitat heterogeneity: individual quality versus territory quality. Oikos , 117: 1111–1114. Google Scholar

13.

Bamford, A. J. , Monadjem, A. , Anderson, M. D. , Anthony, A. , Borello, W. D. , Bridgeford, M. , Hancokc, P. , Howells, B. , Wakelin, J. and Hardy, I. C. W. 2009. Trade-offs between specificity and regional generality in habitat association models: a case study of two species of African vulture. Journal of Applied Ecology , 46: 852–860. Google Scholar

14.

Barnosky, A. D. , Matzke, N. , Tomiya, S. , Wogan, G.O.U. , Swartz, B. , Ouental, T. B. , Marshall, C. , McGuire, J. L. , Lindsey, E. L. , Maguire, K. C. , Merset, B. and Ferrer, A. 2011. Has the Earth's sixth mass extinction already arrived? Nature , 471: 51–57. Google Scholar

15.

Battin, J. 2004. When good animals love bad habitats: Ecological traps and the conservation of animal populations. Conservation Biology , 18: 1482–1491. Google Scholar

16.

Bautista, L. M. , García J.T. , Calmaestra, R. G. , Palacín, C. , Martín, C. A. , Morales, M. B. , Bonal, R. and Viñuela, J. 2004. Effect of weekend road traffic on the use of space by raptors. Conservation Biology , 18: 726–732. Google Scholar

17.

Bayle, P. 1999. Preventing birds of prey problems at transmission lines in Western Europe. Journal of Raptor Research , 33: 43–48. Google Scholar

18.

Bechard, M. J. 1982. Effect of vegetative cover on foraging site selection by Swainson's Hawks. Condor , 84: 153–159. Google Scholar

19.

Becker, N. , Inbar, M. , Bahat, O. , Choresh, Y. , Ben-Noon, G. and Yaffe, O. 2005. Estimating the economic value of viewing griffon vultures Gyps fulvus: a travel cost model study at Gamla Nature Reserve, Israel. Oryx , 39: 429–434. Google Scholar

20.

Beja, P. and Palma, L. 2008. Limitations of methods to test density-dependent fecundity hypothesis. Journal of Animal Ecology , 77: 335–340. Google Scholar

21.

Belthoff, J. R. and Dufty, A. M. Jr. 1998. Corticosterone, body condition and locomotor activity: a model for dispersal in screech-owls. Animal Behaviour , 55: 405–415. Google Scholar

22.

Bernard, N. , Michelat, D. , Raoul, F. , Quéré, J. P. , Delattre, P. and Giraudoux, P. 2010. Dietary response of Barn Owls (Tyto alba) to large variations in populations of common voles (Microtus arvalis) and European water voles (Arvicola terrestris). Canadian Journal of Zoology , 88: 416–426. Google Scholar

23.

Beske, A. E. 1982. Local and migratory movements of radio-tagged juvenile harriers. Journal of Raptor Research , 16: 39–53. Google Scholar

24.

Bianchi, F. J. J. A. , Booij, C. J. H. and Tscharntke, T. 2006. Sustainable pest regulation in agricultural landscapes: a review on landscape composition, biodiversity and natural pest control. Philosophical Transactions of the Royal Society of London. B , 273: 1715–1727. Google Scholar

25.

Bijleveld, M. 1974. Birds of Prey in Europe. Mac Millan Press Ltd. London and Basingstore. Google Scholar

26.

Bird, D. M. , Varland, D. E. and Negro, J. J. 1996. Raptors in Human Landscapes: Adaptations to Built and Cultivated Environments. Academic Press. London. Google Scholar

27.

Birdlife International . 2013. Europe-wide monitoring schemes highlight declines in widespread farmland birds. Presented as part of the BirdLife State of the world’s birds website. Available from:  http://www.birdlife.org/datazone/sowb/casestudy/62. Checked: 02/04/2016 Google Scholar

28.

BirdLife International. 2015. European Red List of Birds. Office for Official Publications of the European Communities. Luxembourg. Google Scholar

29.

Bloom, P. H. and McCrary, M. D. 1996. The urban buteo: red-shouldered hawks in Southern California. In D. M. Bird , D. E. Varland and J. J. Negro . Raptors in Human Landscapes: Adaptations to Built and Cultivated Environments. pp. 31–39. Academic Press. London. Google Scholar

30.

Bock, C. E. , Cruz J. R. , A., Grant, M. C. , Aid, C. S. and Strong, T. R. 1992. Field experimental evidence for diffuse competition among southwestern riparian birds. American Naturalist , 140: 815–828. Google Scholar

31.

Bohrer, G. , Brandes, D. , Mandel, J. T. , Bildstein, K. L. , Miller, T. A. , Lanzone, M. , Katzner, T. , Maisonneuve, C. and Tremblay, J. A. 2012. Estimating updraft velocity components over large spatial scales: contrasting migration strategies of golden eagles and turkey vultures. Ecology Letters , 15: 96–103. Google Scholar

32.

Bonebrake, T. C. , Christensen, J. , Boggs, C. L. and Ehrlich, P. R. 2010. Population decline assessment, historical baselines, and conservation. Conservation Letters , 3: 371–378. Google Scholar

33.

Börger, L. 2016. Stuck in motion? Reconnecting questions and tools in movement ecology. Journal of Animal Ecology , 85: 5–10. Google Scholar

34.

Both, C. and Wisser, M. E. 2000. Breeding territory size affects fitness: an experimental study on competition at the individual level. Journal of Animal Ecology , 69: 1021–1030. Google Scholar

35.

Bradley, C. A. and Altizer, S. 2006. Urbanization and the ecology of wildlife diseases. Trends in Ecology & Evolution , 22: 95–102. Google Scholar

36.

Branton, M. and Richardson, J. S. 2011. Assessing the value of the umbrella-species concept for conservation planning with metaanalysis. Conservation Biology , 25: 9–20. Google Scholar

37.

Breckenridge, W. J. 1935. An ecological study of some Minnesota marsh hawks. Condor , 37: 268–276. Google Scholar

38.

Bremner-Harrison, S. , Prodohl, P. A. and Elwood, R. W. 2004. Behavioural trait assessment as a release criterion: boldness predicts early death in a reintroduction programme of captive-bred swift fox (Vulpes velox). Animal Conservation , 7: 313–320. Google Scholar

39.

Brito, P. H. 2007. Contrasting patterns of mitochondrial and microsatellite genetic structure among Western European populations of tawny owls (Strix aluco). Molecular Ecology , 16: 3423–3437. Google Scholar

40.

Buckle, A. P. and Smith, R. H. 2015. Rodent Pests and Their Control. CAB International. Wallinford. U. K. Google Scholar

41.

Burfield, I. J. 2008. The conservation status and trends of raptors and owls in Europe. AMBIO , 37: 401–407. Google Scholar

42.

Burgas, D. , Byholm, P. and Parkkima, T. 2014. Raptors as surrogates of biodiversity along a landscape gradient. Journal of Applied Ecology , 51: 786–794. Google Scholar

43.

Cadahía, L. , López-López, P. , Urios, V. and Negro, J. J. 2010. Satellite telemetry reveals individual variation in juvenile Bonelli's eagle dispersal areas. European Journal of Wildlife Research , 56: 923–930. Google Scholar

44.

Cade, T. J. and Bird, D. M. 1990. Peregrine falcons, Falco peregrinus, nesting in an urban environment—a review. Canadian Field-Naturalist , 104: 209–218. Google Scholar

45.

Cade, T.J. and Burnham, W. (Eds). 2003. Return of the Peregrine. A North America Saga of Tenacity and Teamwork. The Peregrine Fund, Idaho. Google Scholar

46.

Cade, T.J. , Martell, M. , Redig, P. , Septon, G. and Tordoff, H. 1996. Peregrine falcons in urban North America. In, D. M. Bird , D. E. Varland and J. J. Negro (Eds). Raptors in Human Landscapes: Adaptations to Built and Cultivated Environments , pp. 3–13. Academic Press. London. Google Scholar

47.

Cagnacci, F. , Boitani, L. , Powell, R. A. and Boyce, M. S. 2010. Animal ecology meets GPS-based radiotelemetry: a perfect storm of opportunities and challenges. Philosophical Transactions of the Royal Society of London B , 365: 2157–2162. Google Scholar

48.

Campioni, L. , Delgado, M. M. , Lourenço, R. , Bastianelli, G. , Fernández, N. and Penteriani, V. 2013. Individual and spatio-temporal variations in the home range behaviour of a long-lived, territorial species. Oecologia , 172: 371–385. Google Scholar

49.

Cardador, L. , Navarro, J. , Forero, M. G. , Hobson, K. A. and Mañosa, S. 2015. Breeding origin and spatial distribution of migrant and resident harriers in a Mediterranean wintering area: insights from isotopic analyses, ring recoveries and species distribution modelling. Journal of Ornithology , 156: 247–256. Google Scholar

50.

Caro, T. M. and O'Doherty, G. 1999. On the use of surrogate species in conservation biology. Conservation biology , 13: 805–814. Google Scholar

51.

Carrete, M. and Donázar, J. A. 2005. Application of central-place foraging theory shows the importance of Mediterranean dehesas for the conservation of the cinereous vulture, Aegypius monachus. Biological Conservation , 126: 582– 590. Google Scholar

52.

Carrete, M. , Donázar, J. A. and Margalida, A. 2006. Density-dependent productivity depression in Pyrenean bearded vultures: implications for conservation. Ecological Applications , 16: 1674–1682. Google Scholar

53.

Carrete, M. , Grande, J. M. , Tella, J. L. , Sánchez-Zapata, J. A. , Donázar, J. A. , Díaz-Delgado, R. and Romo, A. 2007. Habitat, human pressure, and social behavior: Partialling out factors affecting large-scale territory extinction in an endangered vulture. Biological Conservation , 136: 143–154. Google Scholar

54.

Carrete, M. , Lambertucci, S. A. , Speziale, K. , Ceballos, O. , Travaini, A. , Delibes, M. , Hiraldo, F. and Donázar, J. A. 2010. Winners and losers in human-made habitats: interspecific competition outcomes in two Neotropical vultures. Animal Conservation , 13: 390–398. Google Scholar

55.

Carrete, M , Sánchez-Zapata, J. A. , Benítez, J. R. , Lobón, M. and Donázar, J. A. 2009. Large scale risk-assessment of wind-farms on population viability of a globally endangered long-lived raptor. Biological Conservation , 142: 2954–2961. Google Scholar

56.

Carrete, M. , Sánchez-Zapata, J. A. , Benítez, J. R. , Lobón, M. , Montoya, F. and Donázar, J. A. 2012. Mortality at wind farms is positively related to bird abundances. Biological Conservation , 145: 102–108. Google Scholar

57.

Carrete, M. and Tella, J. L. 2010. Individual consistency in flight initiation distances in burrowing owls: a new hypothesis on disturbance-induced habitat selection. Biology Letters , 6: 167–170. Google Scholar

58.

Carrete, M. and Tella, J. L. 2013. High individual consistency in fear of humans throughout the adult lifespan or rural and urban burrowing owls. Scientific Reports , 3: 3524. Google Scholar

59.

Carrete, M. , Tella, J. L. , Sánchez-Zapata, J. A. , Moleón, M. and Gil-Sánchez, J. M. 2008. Current caveats and further directions in the analysis of density-dependent population regulation. Oikos , 117: 1115–1119. Google Scholar

60.

Carver, E. 2013. Birding in the United States: a Demographic and Economic Analysis. Report 2011-1. U.S. Fish and Wildlife Service. Arlington. Google Scholar

61.

Caswell, H. 2001. Matrix Population Models. Sinauer Associates. Sunderland, Massachusetts. Google Scholar

62.

Chace, J. F. and Walsh, J. J. 2006. Urban effects on native avifauna: a review. Landscape and Urban Planing , 74: 46–69. Google Scholar

63.

Chapman, A. and Buck, W. J. 1893. Wild Spain. Gurney and Jackson. London. Google Scholar

64.

Chapman, A. and Buck, W. J. 1910. Unexplored Spain. Edward Arnold. London. Google Scholar

65.

Chapman, B. B. , Brönmark, C. , Nilsson, J-.A. and Hansson, L.-A. 2011. The ecology and evolution of partial migration. Oikos , 120: 1764–1775. Google Scholar

66.

Chevallier, C. , Hernández-Matías, A. , Real, J. , Vincent-Martin, N. , Ravayrol, A. and Besnard, A. 2015. Retrofitting of power lines effectively reduces mortality by electrocution in large birds: an example with the endangered Bonelli's eagle. Journal of Applied Ecology , 52: 1465–1473. Google Scholar

67.

Coccon, F. , Zucchetta, M. , Bossi, G. , Borrotti, M. , Torricelli, P. and Franzoi, P. 2015. A land-use perspective for birdstrike risk assessment: the attraction risk index. PLoS ONE , 10: e0128363. Google Scholar

68.

Cochran, W. W. 1975. Following a migrating peregrine from Wisconsin to Mexico. Hawk Chalk , 14: 28–37. Google Scholar

69.

Cortés-Avizanda, A. , Donázar, J. A. and Pereira, H. M. 2015. Top scavengers in a wilder Europe. In, H. M. Pereira and L. Navarro (Ed.): Rewilding European Landscapes , pp. 87–106. Springer. London. Google Scholar

70.

Cortés-Avizanda, A. , Jovani, R. , Carrete, M. and Donázar, J. A. 2012. Resource unpredictability promotes species diversity and coexistence in an avian scavenger guild: a field experiment. Ecology , 93: 2570–2579. Google Scholar

71.

Coulson, T. , Milner-Gulland, E. J. and Clutton-Brock . 2000. The relative roles of density and climatic variation on population dynamics and fecundity rates in three contrasting ungulate species. Philosophical Transactions of the Royal Society of London B , 267: 1771–1779. Google Scholar

72.

Cox, N. A. and Temple, H. J. 2009. European Red List of Reptiles. Office for Official Publications of the European Communities. Luxembourg. Google Scholar

73.

Cugnasse, J. M. 2004. Le faucon pèlerin Falco peregrinus devient-il anthropophile? Alauda , 72: 107–124. Google Scholar

74.

D'Elia, J. , Haig, S. M. , Johnson, M. , Marcot, B.G. and Young, R. 2015. Activity-specific ecological niche models for planning reintroductions of California condors (Gymnogyps californianus). Biological Conservation , 184: 90–99. Google Scholar

75.

Daily, G. C. , Polasky, S. , Goldstein, J. , Kareiva, P. M. , Mooney, H. A. , Pejchar, L. , Ricketts, T. H. , Salzman, J. and Shallenberger, R. 2009. Ecosystem services in decision making: time to deliver. Frontiers in Ecology and the Environment , 7: 21–28. Google Scholar

76.

Davis, J. M. and Stamps, J. A. 2004. The effect of natal experience on habitat preferences. Trends in Ecology & Evolution , 19: 411–416. Google Scholar

77.

De Klemm, C. and Shine, C. 1993. Biological Diversity Conservation and the Law. IUCN. Gland. Cambridge. Google Scholar

78.

Deinet, S. , Ieronymidou, C. , McRae, L. , Burfield, I. J. , Foppen, R. P. , Collen, B. and Böhm, M. 2013. Wildlife Comeback in Europe: The Recovery of Selected Mammal and Bird Species. ZSL, BirdLife International and EBCC. London. Google Scholar

79.

Delgado, M. M. , Penteriani, V. , Revilla, E. and Nams, V. O. 2010. The effect of phenotypic traits and external cues on natal dispersal movements. Journal of Animal Ecology , 79: 620–632. Google Scholar

80.

Del Moral, J. C. and Martí, R. 2004. Atlas de aves reproductoras de España. Dirección General de Conservación de la Naturaleza-SEO/BirdLife. Madrid. Google Scholar

81.

Delibes-Mateos, M. , Delibes, M. , Ferreras, P. and Villafuerte, R. 2008. Key role of European rabbits in the conservation of the Western Mediterranean basin hotspot. Conservation Biology , 22: 1106–1017. Google Scholar

82.

DeVault, T. L. , Blackwell, B. F. and Belant, J. L. 2013. Wildlife in Airport Environments: Preventing Animal-Aircraft Collisions through Science-Based Management. Johns Hopkins University Press. Baltimore. Google Scholar

83.

DeVault, T. L. , Rhodes Jr., O. E. and Shivik, J. A. 2003. Scavenging by vertebrates: behavioral, ecological, and evolutionary perspectives on an important energy transfer pathway in terrestrial ecosystems. Oikos , 102: 225–234. Google Scholar

84.

DeVault, T. L. , Seamans, T. W. , Schmidt, J. A. , Belant, J. L. , Blackwell, B. F. , Mooers, N. Tyson, L. A. and Van Pelt, L. 2014. Bird use of solar photovoltaic installations at US airports: implications for aviation safety. Landscape and Urban Planning , 122: 122–128. Google Scholar

85.

Dingemanse, N. J. , Both, C. , van Noordwijk, A. J. , Rutten, A. I. and Drent, P. J. 2003. Natal dispersal and personalities in great tits (Parus major). Philosophical Transactions of the Royal Society of London B , 270: 741–747. Google Scholar

86.

Dirzo, R. , Young, H. S. , Galetti, M. , Ceballos, G. , Isaac, N. J. B. and Collen, B. 2014. Defaunation in the Anthropocene. Science , 345:401–406. Google Scholar

87.

Dodge, S. , Bohrer, G. , Bildstein, K. , Davidson, S. C. , Weinzierl, R. , Bechard, M. J. , Barber, D. , Kays, R. Brandes, D. , Han, J. and Wikelski, M. 2014. Environmental drivers of variability in the movement ecology of turkey vultures (Cathartes aura) in North and South America. Philosophical Transactions of the Royal Society of London B , 369: 20130195. Google Scholar

88.

Donald, P. F. , Green, R. E. and Heath, M. F. 2001. Agricultural intensification and the collapse of Europe's farmland bird populations. Philosophical Transactions of the Royal Society of London B , 268: 25–29. Google Scholar

89.

Donald, P. F. , Collar, N. J. , Marsden, S. J. and Pain, D. J. 2013. Facing extinction. The world's rarest birds and the race to save them. Christopher Helm. London. Google Scholar

90.

Donald, P. F. , Sanderson, F. J. , Burfield, I. J. , Bierman, S. M. , Gregory, R. D. and Waliczky, Z. 2007. International conservation policy delivers benefits for birds in Europe. Science , 317: 810–813. Google Scholar

91.

Donázar, J. A. and Castién, E. 1989. Predación del búho real (Bubo bubo) sobre la perdiz roja (Alectoris rufa): selección de edad y sexo. Doñana, Acta Vertebrata , 16: 219–225. Google Scholar

92.

Donázar, J. A. , Cortés-Avizanda, A. and Carrete, M. 2010. Dietary shifts in two vultures after the demise of supplementary feeding stations: consequences of the EU sanitary legislation. European Journal of Wildlife Research , 56: 613–621. Google Scholar

93.

Donázar, J. A. , Hiraldo, F. and Bustamante, J. 1993. Factors influencing nest site selection, breeding density and breeding success in the bearded vulture (Gypaetus barbatus). Journal of Applied Ecology , 30: 504–514. Google Scholar

94.

Donázar, J. A. , Margalida, A. and Campión, D. 2009. Buitres, Muladares y Legislación Sanitaria: Perspectivas de un Conflicto y sus Consecuencias desde la Biología de la Conservación. Munibe 29 (Suplemento). Sociedad de Ciencias Aranzadi. San Sebastián. Google Scholar

95.

Donázar, J. A. , Margalida, A. , Carrete, M. and Sánchez-Zapata, J. A. 2009. Too sanitary for vultures. Science , 326: 664. Google Scholar

96.

Douglas, L. R. and Veríssimo, D. 2013. Flagships or battleships: deconstructing the relationship between social conflict and conservation flagship species. Environment and Society: Advances in Research , 4: 98–116. Google Scholar

97.

Drewitt, E. J. A. and Dixon, N. 2008. Diet and prey selection of urban-dwelling peregrine falcons in southwest England. British Birds , 101: 58–67. Google Scholar

98.

Duckett, J. E. 1991. Management of the barn owl (Tyto alba javanica) as a predator of rats in oil palm (Elaeis quineensis) plantations in Malaysia. Birds of Prey Bulletin , 4: 11–24. Google Scholar

99.

Duckworth, R. A. and Badyaev, A. V. 2007. Coupling of dispersal and aggression facilitates the rapid range expansion of a passerine bird. Proceedings of the National Academy of Sciences of the United States of America , 104: 15017–15022. Google Scholar

100.

Duncan, R. P. , Boyer, A. G. and Blackburn, T. M. 2013. Magnitude and variation of prehistoric bird extinctions in the Pacific. Proceedings of the National Academy of Sciences of the United States of America , 110: 6436–6441. Google Scholar

101.

Enserik, M. and Vogel, G. 2006. The carnivore comeback. Science , 314: 746–749. Google Scholar

102.

Errington, P. L. 1946. Predation and vertebrate populations. Quarterly Review of Biology , 21: 221–245. Google Scholar

103.

Estrada, C. G. and Rodríguez-Estrella, R. 2016. In the search of good biodiversity surrogates: are raptors poor indicators in the Baja California Peninsula desert? Animal Conservation , https://doi.org/10.1111/acv.12252Google Scholar

104.

Fajardo, I. 2001. Monitoring non-natural mortality in the barn owl (Tyto alba), as an indicator of land use and social awareness in Spain. Biological Conservation , 97: 143–149. Google Scholar

105.

Fall, M. W. 1977. Rodents in tropical rice. Rodent Research Center, University of the Philippines at Los Baños, College, Laguna. Technical Bulletin , 36: 1–39. Google Scholar

106.

Fargallo, J. A. , Martínez-padilla, J. , Viñuela, J. , Blanco, G. , Torre, I. , Vergara, P. and De Neve, L. 2009. Kestrel-prey dynamic in a Mediterranean region: the effect of generalist predation and climatic factors. PLoS ONE , 4: e4311. Google Scholar

107.

Fasciolo, A. , Delgado M. M. , Soutullo, A. , Cortés G. and Penteriani, V. 2016. Limited prospecting behaviour of juvenile Eagle Owls Bubo bubo during natal dispersal: implications for conservation. Bird Study , 63: 128–135. Google Scholar

108.

Fernández, C. 1993. Effect of the viral haemorrhagic pneumonia of the wild rabbit on the diet and breeding success of the golden eagle Aquila chrysaëtos (L.). Revue d'Ecologie-La Terre et laVie , 48: 323–329. Google Scholar

109.

Fernández, M. , Oria, J. , Sánchez, R. , González, L. M. and Margalida, A. 2009. Space use of adult Spanish imperial eagles Aquila adalberti. Acta Ornithologica , 44: 17–26. Google Scholar

110.

Ferrer, M. 1993. Reduction in hunting success and settlement strategies in young Spanish imperial eagles. Animal Behaviour , 45: 406– 408. Google Scholar

111.

Ferrer, M. , de la Riva, M. and Castroviejo, J. 1991. Electrocution of raptors on power lines in Southern Spain. Journal of Field Ornithology , 62: 54–69. Google Scholar

112.

Ferrer, M. and Donázar, J. A. 1996. Density-dependent fecundity by habitat heterogeneity in an increasing population of Spanish imperial eagles. Ecology , 77: 69–74. Google Scholar

113.

Ferrer, M. and Negro, J. J. 1992. Tendidos eléctricos y conservación de aves en España. Ardeola , 39: 23–27. Google Scholar

114.

Ferrer, M. , Newton, I. and Casado, E. 2008. Density dependence hypotheses and the distribution of fecundity. Journal of Animal Ecology , 77: 341–345. Google Scholar

115.

Fiuczynski, K. D. 2011. Two raptor species, European kestrel Falco tinunculus and Eurasian hobby Falco subbuteo in the forest of Berlin, Germany. In, I. Zuberogoitia and J. E. Martínez (Eds): Ecology and conservation of European forest-dwelling raptors , pp. 99–105. Diputación Foral de Bizkaia. Bilbao. Google Scholar

116.

Forero, M. G. , Donázar, J. A. and Hiraldo, F. 2002. Causes and fitness consequences of natal dispersal in a population of black kites. Ecology , 83: 858–872. Google Scholar

117.

Forman, R. T. and Alexander, L. E. 1998. Roads and their major ecological effects. Annual Review of Ecology and Systematics , 29: 207–231. Google Scholar

118.

Frank, B. 2016. Human-wildlife conflicts and the need to include tolerance and coexistence: an introductory comment. Society & Natural Resources , 29: 738–743. Google Scholar

119.

Franklin, J. 2009. Mapping Species Distributions: Spatial Inference and Prediction. Cambridge University Press. Cambridge. Google Scholar

120.

Frederiksen, M. , Lebreton, J. -D. , Pradel, R. , Choquet, R. and Giménez, O. 2014. Identifying links between vital rates and environment: a toolbox for the applied ecologist. Journal of Applied Ecology , 51: 71–81. Google Scholar

121.

Freyhof, J. and Brooks, E. 2011. European Red List of Freshwater Fishes. Publications Office of the European Union. Luxembourg. Google Scholar

122.

Frías, O. 1999. Estacionalidad de los atropellos de aves en el centro de España: número y edad de los individuos y riqueza y diversidad de especie. Ardeola , 46: 23–30. Google Scholar

123.

Fuller, R. 1987. The changing extent and conservation interest of lowland grasslands in England and Wales: a review of grassland surveys 1930–1984. Biological Conservation , 40: 281–300. Google Scholar

124.

Fuller, R. A. , Warren, P. H. , Armsworth, P. R. , Barbosa, O. and Gaston, K. J. 2008. Garden bird feeding predicts the structure of urban avian assemblages. Diversity and Distributions , 14: 131–137. Google Scholar

125.

Gangoso, L. , Agudo, R. , Anadón, J. D. , De la Riva, M. , Suleyman, A. S. , Porter, R. and Donàzar, J. A. 2013. Reinventing mutualism between humans and wild fauna: insights from vultures as ecosystem services providers. Conservation Letters , 6: 172–179. Google Scholar

126.

García-Fernández, A. J. , Calvo, J. F. , Martínez-López, E. , María-Mojica, P. and Martínez, J. E. 2008. Raptor ecotoxicology in Spain: a review on persistent environmental contaminants. AMBIO , 37: 432–439. Google Scholar

127.

Garitacelaya, J. 2003. 1961, el año de las alimañas. Navarra Forestal , 6: 29–34. Google Scholar

128.

Geroudet, P. 1964. Les Rapaces Diurnes et Nocturnes d'Europe. Delachaux et Niestlé. Neuchatel. Google Scholar

129.

Gilg, O. , Sittler, B. , Sabard, B. , Hurstel, A. , Sané, R. , Delattre, P. and Hanski, I. 2006. Functional and numerical responses of four lemming predators in high arctic Greenland. Oikos , 113: 193–216. Google Scholar

130.

Gillson, L. , Ladle, R. J. and Araújo, M. B. 2011. Baselines, patterns and process. In, R. J. Ladle and R. J. Whittaker (Eds.): Conservation biogeography , pp. 31–44. Wiley-Blackwell. Oxford. Google Scholar

131.

Glista, D. J. , DeVault, T. L. and DeWoody, A. J. 2008. Vertebrate road mortality predominantly impact amphibians. Herpetological Conservation and Biology , 3: 77–87. Google Scholar

132.

González, L. M. , Margalida, A. , Mañosa, S. , Sánchez, R. , Oria, J. , Molina, J. I. , Caldera, J. Aranda, A. and Prada, L. 2007. Causes and spatio-temporal variations of non-natural mortality in the vulnerable Spanish imperial eagle (Aquila adalberti) during a recovery period. Oryx , 41: 495–502. Google Scholar

133.

González, L. M. , Oria, J. , Margalida, A. , Sánchez, R. , Prada, L. , Caldera, J. , Aranda, A. and Molina, J. I. 2006. Effective natal dispersal and age of maturity in the threatened Spanish imperial eagle Aquila adalberti: Conservation implications. Bird Study , 53: 285–293. Google Scholar

134.

González-Oreja, J. A. , 2011. Birds of different biogeographic origins respond in contrasting ways to urbanization. Biological Conservation , 144: 234–242. Google Scholar

135.

Grado, S. C. , Kaminski, R. M. , Munn, I. A. and Tullos, T. A. 2001. Economic impacts of waterfowl hunting on public lands and at private lodges in the Mississippi Delta. Wildlife Society Bulletin , 29: 846–855. Google Scholar

136.

Grande, J. M. , Serrano, D. , Tavecchia, G. , Carrete, M. , Geballos, O. , Díaz-Delgado, R. , Tella, J. L. and Donázar, J. A. 2009. Survival in a long-lived territorial migrant: effects of life-history traits and ecological conditions in wintering and breeding areas. Oikos , 118: 580–590. Google Scholar

137.

Green, A. J. and Elmberg, J. 2014. Ecosystem services provided by waterbirds. Biological Reviews , 89: 105–122. Google Scholar

138.

Gregory, R. D. , Vorisek, P. , Van Strien, A. , Meyling, A. W. G. , Jiguet, F. , Fornasari, L. , Reif, J. , Chylarecki, P. and Burfield, I. J. 2007. Population trends of widespread woodland birds in Europe. Ibis , 149: 78–97. Google Scholar

139.

Gurarie, E. , Bracis, C. , Delgado, M. , Meckley, T. D. , Kojola, I. and Wagner, C. M. 2016. What is the animal doing? Tools for exploring behavioural structure in animal movements. Journal of Animal Ecology , 85: 69–84. Google Scholar

140.

Hadler, M. R. and Buckle, A. P. 1992. Forty five years of anticoagulant rodenticides: past, present and future trends. In, J. E. Borrecco and M. E. Marsh (Eds.): Proceedings of the 15th Vertebrate Pest Conference , pp. 149–155. University of California. Davis. Google Scholar

141.

Hafidzi, M. N. and Mohd, N. 2003. The use of the barn owl, Tyto alba, to suppress rat damage in rice fields in Malaysia, In, G. Singlenton , L. Hinds , C. Krebs and D. Spratt (Eds.): Rats, Mice and People: Rodent Biology and Management , pp. 233–27. ACIAR. Bruce. Google Scholar

142.

Hager, S. B. 2009. Human-related threats to urban raptors. Journal of Raptor Research , 43: 210–226. Google Scholar

143.

Harrison, X. A. , Blount, J. D. , Inger, R. , Norris, D. R. and Bearhop, S. 2011. Carry-over effects as drivers of fitness differences in animals. Journal of Animal Ecology , 80: 4–18. Google Scholar

144.

Helander, B. , Bignere, A. and Asplund , L. 2008. Using raptors as environmental sentinels: monitoring the white-tailed sea eagle Haliaeetus albicilla in Sweden. AMBIO , 37: 425–431. Google Scholar

145.

Henrioux, F. 2000. Home range and habitat use by the long-eared owl in northwestern Switzerland. Journal of Raptor Research , 34: 93–101. Google Scholar

146.

Hernández, M. 1988. Road mortality of the little owl (Athene noctua) in Spain. Journal of Raptor Research , 22: 81–84. Google Scholar

147.

Hernández, R. R. , Easter, S. B. , Murphy-Mariscal, M. L. , Maestre, F. T. , Tavassoli, M. , Allen, E. B. , Barrows, C. W. , Belnap, J. , Ochoa-Hueso, R. , Ravi, S. and Allen, M. F. 2014. Environmental impacts of utility-scale solar energy. Renewable and Sustainable Energy Reviews , 29: 766–779. Google Scholar

148.

Hernández-Matías, A. , Real, J. , Moleón, M. , Palma, L. , Sánchez-Zapata, J. A. , Pradel, R. , Carrete, M. , Gil-Sánchez, J. M. , Beja, P. , Balbontín, J. , Vincent-Martin, N. , Ravayrol, A. , Benítez, J. R. , Arroyo, B. , Fernández, C. , Ferreiro, E. and García, J. 2013. From local monitoring to a broad-scale viability assessment: a case study for the endangered Bonelli's eagle Aquila fasciata in Western Europe. Ecological Monographs , 83: 239–261. Google Scholar

149.

Hernández-Matías, A. , Real, J. , Parés, F. and Pradel, R. 2015. Electrocution threatens the viability of populations of the endangered Bonelli's eagle (Aquila fasciata) in Southern Europe. Biological Conservation , 191: 110– 116. Google Scholar

150.

Hiraldo, F. , Delibes, M. and Calderón, J. 1979. El Quebrantahuesos Gypaetus barbatus (L.). ICONA, Monografías 22. Ministerio de Agricultura. Madrid. Google Scholar

151.

Hiraldo, F. , Negro, J. J. , Donázar, J. A. and Gaona, P. 1996. A demographic model for a population of the endangered lesser kestrel in Southern Spain. Journal of Applied Ecology , 33: 1085–1093. Google Scholar

152.

Hiraldo, F. and Sergio, F. 2008. Intraguild predation in raptor assemblages: a review. Ibis , 150: 132–145. Google Scholar

153.

Hobbs, R. J. , Arico, S. , Aronson, J. , Baron, J. S. , Bridgewater, P. , Cramer, V. A. , Epstein, P. R. , Ewel, J. J. , Klink, C. A. , Lugo, A. E. , Norton, D. , Ojima, D. , Rochardson, D. M. , Sanderson, E. W. , Valladares, F. , Villa, M. , Zamora, R. and Zobel, M. 2006. Novel ecosystems: theoretical and management aspects of the new ecological word order. Global Ecology and Biogeography , 15: 1–7. Google Scholar

154.

Hodge, I. , Hauck, J. and Bonn, A. 2015. The alignment of agricultural and nature conservation policies in the European Union. Conservation Biology , 29: 996–1005. Google Scholar

155.

Holt, A. R. , Davies, Z. J. , Tyler, G. and Staddon, S. 2008. Meta-analysis of the effects of predation on animal prey abundance: evidence from UK vertebrates. PLoS ONE , 3: e2400. Google Scholar

156.

Hooper, D. U. , Chapin III, F. S. , Ewel, J. J. , Hector, A. , Inchausti, P. , Lavorel, S. , Lawton, J. H. , Lodge, D. M. , Loreau, M. , Naeem, S. , Schmid, B. , Setälä, H. , Symstad, A. J. , Vandermeer, J. and Wardle, D. A. 2005. Effects of biodiversity on ecosystem functioning: a consensus of current knowledge. Ecological Monographs , 75: 3–35. Google Scholar

157.

Houston, D. C. 1994. Family Cathartidae (New World vultures). In, J. del Hoyo , A. Elliott and J. Sargatal (Eds.): Handbook of the Birds of the World (Vol. 2): New World Vultures to guineafowl , pp. 24–41. Lynx Edicions. Barcelona. Google Scholar

158.

Howard, W. E. 1967. Biological control of vertebrate pests. In, J. E. Borrecco and M. E. Marsh (Eds.): Proceedings of the 15th Vertebrate Pest Conference , pp. 137–157. University of California. Davis. Google Scholar

159.

Hull, J. M. , Hull, A. C. , Sacks, B. N. , Smith, J. P. and Ernest, H. B. 2008. Landscape characteristics influence morphological and genetic differentiation in a widespread raptor (Buteo jamaicensis). Molecular Ecology , 17: 810–824. Google Scholar

160.

Hygnstrom, S. E. , Vercauteren, K. C. and Schmaderer, T. R. 1994. Biological management (control) of vertebrate pests-advances in the last quarter century. In, J. E. Borrecco and M. E. Marsh (Eds.): Proceedings of the 15th Vertebrate Pest Conference , pp. 293–300. University of California. Davis. Google Scholar

161.

ICAO, International Civil Aviation Organization. 2012. Airport Services Manual Part 3. Bird Control and Reduction. ICAO. 4th ed. Montreal. Google Scholar

162.

Isaac, B. , Cooke, R. , Lerodiaconou, D. and White, J. 2014. Does urbanization have the potential to create an ecological trap for powerful owls (Ninox strenua)? Biological Conservation , 176: 1–11. Google Scholar

163.

Jareño, D. , Viñuela, J. , Luque-larena, J. J. , Arroyo, L. , Arroyo, B. and Mougeot, F. 2015. Factors associated with the colonization of agricultural areas by common voles Microtus arvalis in NW Spain. Biological Invasions , 17: 2315–2327. Google Scholar

164.

Johnson, D. H. 1980. The comparison of usage and availability measurements for evaluating resource preference. Ecology , 61: 65–71. Google Scholar

165.

Kaplan, J. O. , Krumhardt, K. M. and Zimmermann, N. 2009. The prehistoric and preindustrial deforestation of Europe. Quaternary Science Reviews , 28: 3016–3034. Google Scholar

166.

Kauffman, M. J. , Pollock, J. F. and Walton, B. 2004. Spatial structure, dispersal, and management of a recovering raptor population. American Naturalist , 164: 582–597. Google Scholar

167.

Kenward, R. E. 1999. Raptor predation problems and solutions. Journal of Raptor Research , 33: 73–75. Google Scholar

168.

Kenward, R. E. , Hall, D. G. , Walls, S. S. and Hodder, K. H. 2001. Factors affecting predation by buzzards Buteo buteo on released pheasants Phasianus colchicus. Journal of Applied Ecology , 38: 813–822. Google Scholar

169.

Kirby, J. S. , Stattersfield, A. J. , Butchart, S. H. , Evans, M. I. , Grimmett, R. F. , Jones, V. R. , O'Sullivan, J. , Tucker, G. M. and Newton, I. 2008. Key conservation issues for migratory land-and waterbird species on the world’s major flyways. Bird Conservation International , 18: S49–S73. Google Scholar

170.

Kitowski, I. 2011. Civil and military birdstrikes in Europe: an ornithological approach. Journal of Applied Sciences , 11: 183–191. Google Scholar

171.

Klaassen, R. H. G. , Hake, M. , Strandberg, R. , Koks, B. J. , Trierweiler, C. , Exo, K. -M. , Bairlein, F. and Alerstam, T. 2014. When and where does mortality occur in migratory birds? Direct evidence from long-term satellite tracking of raptors. Journal of Animal Ecology , 83: 176–184. Google Scholar

172.

Kogan, M. 1998. Integrated pest management: historical perspectives and contemporary developments. Annual Review of Entomology , 45: 243–270. Google Scholar

173.

Kolecek, J. , Schleuning, M. , Burfield, I. J. , Báldi, A. , Böhning-Gaese, K. , Devictor, V. , Fernández-García, J. M. , Hořák, D. , Van Turnhout, C. A. M. , Hnatyna, O. and Reif, J. 2014. Birds protected by national legislation show improved population trends in Eastern Europe. Biological Conservation , 172: 109–116. Google Scholar

174.

Korpimäki, E. 1985. Prey choice strategies of the kestrel Falco tinnunculus in relation to available small mammals and other Finnish birds of prey. Annales Zoologici Fennici , 22: 91–104. Google Scholar

175.

Korpimäki, E. 1993. Does nest-hole quality, poor breeding success or food depletion drive the breeding dispersal of Tengmalm‘s owls?. Journal of Animal Ecology , 62: 606–613. Google Scholar

176.

Krebs, C. J. 2002. Beyond population regulation and limitation. Wildlife Research , 29: 1–10. Google Scholar

177.

Krüger, O. 2002. Analysis of nest occupancy and nest reproduction in two sympatric raptors: common buzzard Buteo buteo and goshawk Accipiter gentilis. Ecography , 25: 523–532. Google Scholar

178.

Krüger, O. and Lindström, J. 2001. Habitat heterogeneity affects population growth in goshawk Accipiter gentilis. Journal of Animal Ecology , 70: 173–181. Google Scholar

179.

Krüger,O., Chakarov, N. , Nielsen, J. T. , Looft, V. , Grünkorn, T. , Struwe-Juhl, B. and Møller, A. P. 2012. Population regulation by habitat heterogeneity or individual adjustment? Journal of Animal Ecology , 81: 330–340. Google Scholar

180.

Labaune, C. and Magnin, F. 2002. Pastoral management vs. land abandonment in Mediterranean uplands: impact on land snail communities. Global Ecology and Biogeography , 11: 237–245. Google Scholar

181.

Laiolo, P. , Dondero, F. , Ciliento, E. and Rolando, A. 2004. Consequences of pastoral abandonment for the structure and diversity of the alpine avifauna. Journal Applied Ecology , 41: 294–304. Google Scholar

182.

Lambertucci, S. A. , Speziale, K. L. , Rogers, T. E. and Morales, J. M. 2009. How do roads affect the habitat use of an assemblage of scavenging raptors? Biodiversity and Conservation , 18: 2063–2074. Google Scholar

183.

Lanzone, M. J. , Miller, T. A. , Turk, P. , Brandes, D. , Halverson, C. , Maisonneuve, C. , Tremblay, J. , Cooper, J. , O'Malley, K. , Brooks, R. P. and Katzner, T. 2012. Flight responses by a migratory soaring raptor to changing meteorological conditions. Biology Letters , 8: 710–713. Google Scholar

184.

Le Viol, I. , Jiguet, F. , Brotons, L. , Herrando, S. , Lindström, A. , Pearce-Higgins, J. W. , Reif, J. , Van Turnhout, C. and Devictor, V. 2012. More and more generalists: two decades of changes in the European avifauna. Biology Letters , 8: 780–782. Google Scholar

185.

Lieury, N. , Gallardo, M. , Ponchon, C. , Besnard, A. and Millon, A. 2015. Relative contribution of local demography and immigration in the recovery of a geographicallyisolated population of the endangered Egyptian vulture. Biological Conservation , 191: 349–356. Google Scholar

186.

Lima, M. , Stenseth N. C. and Jaksic, F. M. 2002. Food web structure and climate effects on the dynamics of small mammals and owls in semi-arid Chile. Ecology Letters , 5: 273– 284. Google Scholar

187.

López-López, P. , García-Ripollés, C. and Uríos, V. 2014. Individual repeatability in timing and spatial flexibility of migration routes of trans-Saharan migratory raptors. Current Zoology , 60: 642–652. Google Scholar

188.

López-López, P. , Ferrer, M. , Madero, A. , Casado, E. and McGrady, M. 2011. Solving man-induced large-scale conservation problems: the Spanish Imperial Eagle and power lines. PLoS ONE , 6(3): e17196. Google Scholar

189.

López-Sepulcre, A. and Kokko, H. 2005. Territorial defense, territory size, and population regulation. American Naturalist , 166: 317–329. Google Scholar

190.

Loreau, M. , Naeem, S. , Inchausti, P. , Bengtsson, J. , Grime, J. P. , Hector, A. , Hooper, D. U. , Huston, M. A. , Raffaelli, D. , Schmid, B. , Tilman, D. and Wardle, D. A. 2001. Biodiversity and ecosystem functioning: current knowledge and future challenges. Science , 294: 804–808. Google Scholar

191.

Lourenço, R. , Santos, S. M. , Rabaça, J. E. and Penteriani, V. 2011. Superpredation patterns in four large European raptors. Population Ecology , 53: 175–185. Google Scholar

192.

Lövy, M. and Riegert, J. 2013. Home range and land use of urban long-eared owls. Condor , 115: 551–557. Google Scholar

193.

Lowney, M. S. 1999. Damage by black and turkey vultures in Virginia, 1990–1996. Wildlife Society Bulletin , 27: 715–719. Google Scholar

194.

Maas, B. , Karp, D. S. , Bumrungsri, S. , Darras, K. , Gonthier, D. , Huang, J. C.-C. , Lindell, C. A. , Maine, J. J. , Mestre, L. , Michel, N. L. , Morrison, E. B. , Perfecto, I. , Philpott, S. M. , Şekercioğlu, Ç. H. , Silva, R. M. , Taylor, P. J. , Tscharntke, T. , Van Bael, S. A. , Whelan, C. J. and Williams-Guillén, K. 2015. Bird and bat predation services in tropical forests and agroforestry landscapes. Biological Reviews. ( https://doi.org/10.1111/brv.12211). Google Scholar

195.

Mannan, R. W. , Mannan, R. N. , Schmidt, C. A. , Estes-Zumpf, W. A. and Boal, C. W. 2006. Influence of natal experience on nest-site selection by urban-nesting Cooper's Hawks. Journal of Wildlife Management , 71: 64–68. Google Scholar

196.

Mañosa, S. 1991. Biologia tròfica, ús de l'hàbitat I biologia de la reproducció de l'astor Accipiter gentilis (Linnaeus, 1758) a la Segarra. Ph.D. dissertation. Universitat de Barcelona. Google Scholar

197.

Marcström, V. , Kenward, R. E. and Engren, E. 1988. The impact of predation on boreal tetraonids during vole cycles: an experimental study. Journal of Animal Ecology , 57: 859– 872. Google Scholar

198.

Margalida, A. and Campión, D. 2009. Interacciones agresivas entre buitres leonados (Gyps fulvus) y ganado: aspectos ecológicos y económicos de un conflicto emergente. In, J. A. Donázar , A. Margalida and D. Campión (Eds.): Buitres, muladares y legislación sanitaria: perspectivas de un conflicto y sus consecuencias desde la biología de la conservación , pp. 476–491. Munibe 29 (Suplemento). Sociedad de Ciencias Aranzadi. San Sebastián. Google Scholar

199.

Margalida, A. , Campión, D. and Donázar, J. A. 2011. European vultures' altered behaviour. Nature , 480: 457. Google Scholar

200.

Margalida, A. , Campión, D. and Donázar, J. A. 2014. Vultures vs livestock: Conservation relationships in an emergent human-wildlife conflict. Oryx , 48: 172–176. Google Scholar

201.

Margalida, A. , Carrete, M. , Hegglin, D. , Serrano, D. , Arenas, R. and Donázar, J. A. 2013. Uneven large-scale movement patterns in wild and reintroduced pre-adult bearded vultures: conservation implications. PLoS ONE , 8: e65857. Google Scholar

202.

Margalida, A. , Carrete, M. , Sánchez-Zapata, J. A. and Donázar, J. A. 2012. Good news for European vultures. Science , 335: 284. Google Scholar

203.

Margalida, A. and Colomer, M. À. 2012. Modelling the effects of sanitary policies on European vulture conservation. Scientific Reports , 2: art.753. Google Scholar

204.

Margalida, A. , Colomer, M. À. and Oro, D. 2014. Man-induced activities modify demographic parameters in a long-lived species: effects of poisoning and health policies. Ecological Applications , 24: 436–444. Google Scholar

205.

Margalida, A. , Colomer, M. À and Sanuy, D. 2011. Can wild ungulate carcasses provide enough biomass to maintain avian scavenger populations? An empirical assessment using a bio-inspired computational model. PLoS ONE , 6: e20248. Google Scholar

206.

Margalida, A. , Donázar, J. A. , Bustamante, J. , Hernández, F. J. and Romero-Pujante, M. 2008. Application of a predictive model to detect long-term changes in nest-site selection in the bearded vulture Gypaetus barbatus: conservation in relation to territory shrinkage. Ibis , 150: 242–249. Google Scholar

207.

Margalida, A. , Donázar, J. A. , Carrete, M. and Sánchez-Zapata, J. A. 2010. Sanitary versus environmental policies: fitting together two pieces of the puzzle of European vulture conservation. Journal of Applied Ecology , 47: 931–935. Google Scholar

208.

Markandya, A. , Taylor, T. , Longo, A. , Murty, M. N. , Murty, S. and Dhavala, K. 2008. Counting the cost of vulture decline-An appraisal of the human health and other benefits of vultures in India. Ecological Economics , 67: 194–204. Google Scholar

209.

Maron, J. L. , Pearson, D. E. and Fletcher, R. Jr. 2010. Counter-intuitive effects of largescale predator removal on a mid-latitude rodent community. Ecology , 91: 3719–3728. Google Scholar

210.

Marsh, R. E. 1998. Barn Owl Nest Boxes Offer No Solution to Pocket Gopher Damage. In, J. E. Borrecco and M. E. Marsh (Eds.): Proceedings of the 15th Vertebrate Pest Conference , pp. 414–415. University of California. Davis. Google Scholar

211.

Martín, B. and Ferrer, M. 2013. Assessing biodiversity distribution using diurnal raptors in Andalusia, southern Spain. Ardeola , 60: 15–28. Google Scholar

212.

Martínez, J. A. , Martínez, J. E. , Mañosa, S. , Zuberogoitia, I. and Calvo, J. F. 2006. How to manage human-induced mortality in the eagle owl Bubo bubo. Bird Conservation International , 16: 265–278. Google Scholar

213.

Martínez, J. A. , Serrano, D. and Zuberogoitia, I. 2003. Predictive models of habitat preferences for the Eurasian eagle owl Bubo bubo: a multiscale approach. Ecography , 26: 21–28. Google Scholar

214.

Martínez, J. E. , Zuberogoitia, I. , Jiménez-Franco, M. V. , Mañosa, S. and Calvo, J. F. 2016. Spatio-temporal variations in mortality causes of two migratory forest raptors in Spain. European Journal of Wildlife Research , 62: 109–118. Google Scholar

215.

Martínez-Abraín, A. , Crespo, J. , Jiménez, J. , Gómez, J. A. and Oro, D. 2009. Is the historical war against wildlife over in southern Europe? Animal Conservation , 12: 204–208. Google Scholar

216.

Martínez-Cruz, B. , Godoy, J. A. and Negro, J. J. 2004. Population genetics after fragmentation: the case of the endangered Spanish imperial eagle (Aquila adalberti). Molecular Ecology , 13: 2243–2255. Google Scholar

217.

Martínez-Hesterkamp, S. 2015. Territorialidad y relaciones espaciales en rapaces diurnas. Patrones y procesos a escala global y local. Ph.D. dissertation. Universidad de Alcalá. Alcalá de Henares. Google Scholar

218.

Marzluff, J. M. , Bowman, R. , Donnelly, R. (Eds.) 2001. Avian Ecology and Conservation in an Urbanizing World. Kluwer Academic Press. Norwell. Google Scholar

219.

Matthiopoulos, J. , Fieberg, J. , Aarts, G. , Beyer, H. L. , Morales, J. M. and Haydon, D. T. 2015. Establishing the link between habitat selection and animal population dynamics. Ecological Monographs , 85: 413–436. Google Scholar

220.

McGrady, M. J. 1991. The ecology and breeding behaviours of urban Sparrowhawks (Accipiter nisus) in Edinburgh, Scotland. Ph.D. dissertation. University of Edinburgh. Edinburgh. Google Scholar

221.

MEA (Millennium Ecosystem Assessment). 2003. Ecosystems and Human Wellbeing: a Framework for Assessment. Island Press. Washington, D.C. Google Scholar

222.

Melis, C. , Selva, N. , Teurlings, I. , Skarpe, C. , Linnell, J. D. C. and Andersen, R. 2007. Soil and vegetation nutrient response to bison carcasses in Białowieza Primeval Forest, Poland. Ecological Research , 22: 807–813. Google Scholar

223.

Meunier, F. D. , Verheyden, C. and Jouventin, P. 2000. Use of roadsides by diurnal raptors in agricultural landscapes. Biological Conservation , 92: 291–298. Google Scholar

224.

Meyer, A. N. and Kaukienen D. E. 2015. Rodent control in practice: protection of humans and animal health. In, A. P. Buckle and R. H. Smith (Eds.): Rodent Pests and Their Control , pp. 231–246. CAB International. Wallinford. Google Scholar

225.

Meyer, J. S. , Irwin, L. L. and Boyce, M. S. 1998. Influence of habitat abundance and fragmentation on spotted owls in western Oregon. Wildlife Monographs , 139: 1–51. Google Scholar

226.

Mihoub, J.-B. , Giménez, O. , Philippe, P. and Sarrazin, F. 2010. Challenging conservation of migratory species: Sahelian rainfalls drive first-year survival of the vulnerable lesser kestrel Falco naumanni. Biological Conservation , 143: 839–847. Google Scholar

227.

Mikkola, H. 1983. Owls of Europe. T & AD Poyser. Calton. Google Scholar

228.

Millon, A. S. , Petty, J. , Little, B. , Giménez, O. , Cornulier, T. and Lambin, X. 2014. Dampening prey cycle overrides the impact of climate change on predator population dynamics: a long-term demographic study on tawny owls. Global Change Biology , 20: 1770–1781. Google Scholar

229.

Moleón, M. 2007a. El estudio del impacto de los depredadores sobre las presas cinegéticas: un intento de compatibilizar caza y conservación. In, J. M. Barea-Azcón , M. Moleón , R. Travesí , E. Ballesteros-Duperón , J. M. Luzón and J. M. Tierno de Figueroa (Eds.): Biodiversidad y Conservación de Fauna y Flora en Ambientes Mediterráneos , pp. 743–794. Sociedad Granatense de Historia Natural. Granada. Google Scholar

230.

Moleón, M. 2007b. Interacciones ecológicas entre depredadores y presas. Águilas perdiceras, conejos y perdices. Ph.D. dissertation. Universidad de Granada. Granada. Google Scholar

231.

Moleón, M. 2012. Predator-prey Interactions. A Mediterranean Vertebrate System as a Case Study. Lap Lambert Academic Publishing. Saarbrücken. Google Scholar

232.

Moleón, M. , Almaraz, P. and Sánchez-Zapata, J. A. 2008. An emerging infectious disease triggering large-scale hyperpredation. PLoS ONE , 3: e2307. Google Scholar

233.

Moleón, M. , Almaraz, P. and Sánchez-Zapata, J. A. 2013. Inferring ecological mechanisms from hunting bag data in wildlife management: a reply to Blanco-Aguiar et al. (2012). European Journal of Wildlife Research , 59: 599–608. Google Scholar

234.

Moleón, M. , Bautista, J. , Garrido, J. R. , Martín-Jaramillo, J. , Ávila, E. and Madero, A. 2007. La corrección de tendidos eléctricos en áreas de dispersión de águila-azor perdicera: efectos potenciales positivos sobre la comunidad de aves rapaces. Ardeola , 54: 319–325. Google Scholar

235.

Moleón, M. , Bautista, A. and Madero, A. 2011b. Communal roosting in young Bonelli's eagles (Aquila fasciata). Journal of Raptor Research , 45: 353–356. Google Scholar

236.

Moleón, M. , Sánchez-Zapata, J. A. , Gil-Sánchez, J. M. , Ballesteros-Duperón, E. , Barea-Azcón, J. M. and Virgos, E. 2012. Predator-prey relationships in a Mediterranean vertebrate system: Bonelli's eagles, rabbits and partridges. Oecologia , 168: 679–689. Google Scholar

237.

Moleón, M. , Sánchez-Zapata, J. A. , Gil-Sánchez, J. M. , Barea-Azcón, J. M. , Ballesteros-Duperón, E. and Virgós, E. 2011a. Laying the foundations for a human-predator conflict solution: assessing the impact of Bonelli’s eagle on rabbits and partridges. PLoS ONE , 6: e22851. Google Scholar

238.

Moleón, M. , Sánchez-Zapata, J. A. , Margalida, A. , Carrete, M. , Owen-Smith, N. and Donázar, J. A. 2014. Humans and scavengers: the evolution of interactions and ecosystem services. BioScience , 64: 394–403. Google Scholar

239.

Moleón, M. , Sánchez-Zapata, J. A. , Real, J. , García-Charton, J. A. , Gil-Sánchez, J. M. , Palma, L. , Bautista, J. and Bayle, P. 2009. Large-scale spatio-temporal shifts in the diet of a predator mediated by an emerging infectious disease of its main prey. Journal of Biogeography , 36: 1502–1515. Google Scholar

240.

Molina-López, R. A. , Casal, J. and Darwich, L. 2011. Causes of morbidity in wild raptor populations admitted at a wildlife rehabilitation centre in Spain from 1995–2007: a long term retrospective study. PLoS ONE , 6: e24603. Google Scholar

241.

Morales, J. M. , Moorcroft, P. R. , Matthiopoulos, J. , Frair, J. L. , Kie, J. G. , Powell, R. A. , Merrill, E. H. and Haydon, D. T. 2010. Building the bridge between animal movement and population dynamics. Philosophical Transactions of the Royal Society of London B , 365: 2289–2301. Google Scholar

242.

Morales-Reyes, Z. , Pérez-García, J. M. , Moleón, M. , Botella, F. , Carrete, M. , Lazcano, C. , Moreno-Opo, R. , Margalida, A. , Donázar, J. A. and Sánchez-Zapata, J. A. 2015. Supplanting ecosystem services provided by scavengers raises greenhouse gas emissions. Scientific Reports, 5: art. 7811. Google Scholar

243.

Morelli, F. , Kubicka, A. M. , Tryjanowski, P. and Nelson, E. 2015. The vulture in the sky and the hominid on the land: three million years of human-vulture interaction. Anthrozoos , 28: 449–468. Google Scholar

244.

Motro, Y. 2011. Economic evaluation of biological rodent control using barn owls Tyto alba in alfalfa. Eighteenth European Vertebrate Pest Management Conference. Julius-Kühn-Archiv , 432: 79–80. Google Scholar

245.

Murua, R. and Rodríguez, J. 1989. An integrated control system for rodents in pine plantations in central Chile. Journal of Applied Ecology , 26: 81–88. Google Scholar

246.

Nathan, R. , Getz, W. M. , Revilla, E. , Holyoak, M. , Kadmon, R. , Saltz, D. and Smouse, P. E. 2008. A movement ecology paradigm for unifying organismal movement research. Proceedings of the National Academy of Sciences of the United States of America , 105: 19052– 19059. Google Scholar

247.

Nathan, R. , Spiegel, O. , Fortmann-Roe, S. , Harel, R. , Wikelski, M. and Getz, W. M. 2012. Using tri-axial acceleration data to identify behavioral modes of free-ranging animals: general concepts and tools illustrated for griffon vultures. Journal of Experimental Biology , 215: 986–996. Google Scholar

248.

Navarro, L. M. and Pereira, H. M. 2012. Rewilding abandoned landscapes in Europe. Ecosystems , 15: 900–912. Google Scholar

249.

Nevoux, M. , Giménez, O. , Arlt, D. , Nicoll, M. , Jones, C. and Norris, K. 2011. Population regulation of territorial species: both site dependence and interference mechanisms matter. Proceedings of the Royal Society B , 278: 2173– 2181. Google Scholar

250.

Newsome, S. D. , Park, J. S. , Henry, B. , Holden, A. , Fogel, M. , Linthicum, J. , Chu, V. and Hooper, K. 2010. Polybrominated diphenyl ether (PBDE) levels in peregrine falcon (Falco peregrinus) eggs from California correlate with diet and human population density. Environmental Science and Technology , 44: 5248– 5255. Google Scholar

251.

Newton, I. 1979. Population Ecology of Raptors. T & AD Poyser. Berkhamsted. Google Scholar

252.

Newton, I. 1993. Predation and limitation of bird numbers. Current Ornithology , 11: 143–198. Google Scholar

253.

Newton, I. 1998. Population Limitation in Birds. Academic Press. London. Google Scholar

254.

Norrdahl, K. and Korpimäki, E. 1995. Effects of predator removal on vertebrate prey populations: birds of prey and small mammals. Oecologia , 103: 241–248. Google Scholar

255.

Norris, D. R. , Marra, P. P. , Kyser, T. K. , Sherry, T. W. and Ratcliffe, L. M. 2004. Tropical winter habitat limits reproductive success on the temperate breeding grounds in a migratory bird. Philosophical Transactions of the Royal Society of Londons B , 271: 59–64. Google Scholar

256.

Ogada, D. L. , Keesing, F. and Virani, M. Z. 2012. Dropping dead: causes and consequences of vulture population declines worldwide. Annals of the New York Academy of Sciences , 1249: 57–71. Google Scholar

257.

Ogada, D. , Shaw, P. , Beyers, R. L. , Buij, R. , Murn, C. , Thiollay, J. M. , Beale, C. M. , Holdo, R. M. , Pomeroy, D. , Baker, N. , Krüger, S. C. , Botha, A. , Virani, M. Z. , Monadjem, A. and Sinclair, A. R. E. 2016. Another continental vulture crisis: Africa's vultures collapsing toward extinction. Conservation Letters , 9: 89–97. Google Scholar

258.

Olsen, P. 2006. Wedge-tailed Eagle. CSIRO Publishing. Clayton. Google Scholar

259.

Padgett, D. and Glaser, R. 2003. How stress influences the immune response. Trends in Immunology , 24: 444–448. Google Scholar

260.

Pain, D. J. , Cunningham, A. A. , Donald, P. F. , Duckworth, J. W. , Houston, D. C. , Katzner, T. , Parry-Jones, J. , Poole, C. , Prakash, V. , Round, P. and Timmins, R. 2003. Causes and effects of temporal and spatial declines of Gyps vultures in Asia. Conservation Biology , 17: 661–671. Google Scholar

261.

Papp, S. 2011. Breeding of Eurasian sparrowhawks (Accipiter nisus) in two Hungarian towns. Aquila , 118: 49–54. Google Scholar

262.

Park, K. J. , Graham, K. E. , Calladine, J. and Wernham, C. W. 2008. Impacts of birds of prey on gamebirds in the UK: a review. Ibis , 150: 9–26. Google Scholar

263.

Paulos, C. M. 2006. Juntas provinciales de extinción de animales dañinos: de alimañas a especies protegidas. Jara y Sedal , 50: 72–78. Google Scholar

264.

Paz, A. , Jareño, D. , Arroyo, L. , Viñuela, J. , Arroyo, B. E. , Mougeot, F. , Luque-Larena, J. J. and Fargallo, J. A. 2013. Avian predators as a biological control system of common vole (Microtus arvalis) populations in NW Spain: experimental set-up and preliminary results. Pest Management Science , 69: 444–450. Google Scholar

265.

Penteriani, V. and Delgado, M. M. 2009. Thoughts on natal dispersal. Journal of Raptor Research , 43: 90–98. Google Scholar

266.

Penteriani, V. and Delgado, M. M. 2012. There is a limbo under the moon: what social interactions tell us about the floaters' underworld. Behavioral Ecology and Sociobiology , 66: 317–327. Google Scholar

267.

Penteriani, V. , Otalora, F. and Ferrer, M. 2008. Floater mortality within settlement areas can explain the Allee effect in breeding populations. Ecological Modelling , 213: 98–104. Google Scholar

268.

Penteriani, V. , Otalora, F. , Sergio, F. and Ferrer, M. 2005. Environmental stochasticity in dispersal areas can explain the ‘mysterious’ disappearance of breeding populations. Philosophical Transactions of the Royal Society of London B , 272: 1265–1269. Google Scholar

269.

Pereira, H. M. and Navarro, L. 2015. Rewilding European Landscapes. Springer, London. Google Scholar

270.

Pereira, H. M. , Leadley, P. W. , Proença, V. , Alkemade, R. , Scharlemann, J. P. W. , Fernández-Manjarres, J. R , Araújo, M. B. , Balvanera, P. , Biggs, R. , Cheung, W. W. L. , Chini, L. , Cooper, H. D. , Gilman, E. L. , Guenette, S. , Hurtt, G. C. , Huntington, H. P. , Mace, G. M. , Oberdorff, T. , Revenga, C. , Rodrigues, P. , Scholes, R. J. , Sumaila, U. R. and Walpole, M. 2010. Scenarios for global biodiversity in the 21st century. Science , 330: 1496–1501. Google Scholar

271.

Pérez-García, J. M. , Botella, F. , Sánchez-Zapata, J. A. and Moleón, M. 2011. Conserving outside protected areas: edge effects and avian electrocutions in the periphery of Special Protected Areas. Bird Conservation International , 21: 296–302. Google Scholar

272.

Pérez-García, J. M. , Sebastián-González, E. , Botella, F. and Sánchez-Zapata, J. A. 2016. Selecting indicator species of infrastructure impacts using network analysis and biological traits: bird electrocution and power lines. Ecological Indicators , 60: 428–433. Google Scholar

273.

Peterson, M. N. , Birckhead, J. L. , Leon, K. , Peterson, M. J. and Peterson, T. R. 2010. Rearticulating the myth of human-wildlife conflict. Conservation Letters , 3: 74–82. Google Scholar

274.

Piana, R. P. and Stuart, J. 2012. Diversity, community structure, and niche characteristics within a diurnal raptor assemblage of north-western Peru. Condor , 114: 279–289. Google Scholar

275.

Pirovano, A. , Rubolimi, D. , Brambilla, S. and Ferrari, N. 2000. Winter diet of urban roosting long-eared owls (Asio otus) in northern Italy: the importance of the brown rat (Rattus norvegicus). Bird Study , 47: 242–244. Google Scholar

276.

Ranazzi, L. , Manganaro, A. , Ranazzi, R. and Salvati, L. 2000. Woodland cover and tawny owl Strix aluco density in a Mediterranean urban area. Biota , 1: 27–34. Google Scholar

277.

Real, J. , Grande, J. M. , Mañosa, S. and Sánchez-Zapata, J. A. 2001. Causes of death in different areas for Bonelli's eagle (Hieraaetus fasciatus) in Spain. Bird Study , 48: 221–228. Google Scholar

278.

Rebolo-Ifrán, N. , Carrete, M. , Sanz-Aguilar, A. , Rodríguez-Martínez, S. , Cabezas, S. , Marchant, T. A. , Bortolotti, G. R. and Tella, J. L. 2015. Links between fear of humans, stress and survival support a non-random distribution of birds among urban and rural habitats. Scientific Reports , 5: 13723. Google Scholar

279.

Redpath, S. M. , Bhatia, S. and Young, J. 2015. Tilting at wildlife: reconsidering human-wildlife conflict. Oryx , 49: 222–225. Google Scholar

280.

Redpath, S. M. and Thirgood, S. J. 1997. Birds of Prey and Red Grouse. Her Majesty's Stationery Office. London. Google Scholar

281.

Redpath, S. M. and Thirgood, S. J. 1999. Numerical and functional responses of generalist predators: hen harriers and peregrine falcons on Scottish grouse moors. Journal of Animal Ecology , 68: 879–892. Google Scholar

282.

Redpath, S. M. , Thirgood, S. J. and Leckie, F. M. 2001. Does supplementary feeding reduce harrier predation on red grouse? Journal of Applied Ecology , 38: 1157–1168. Google Scholar

283.

Redpath, S. M. , Young, J. , Evely, A. , Adams, W. M. , Sutherland, W. J. , Whitehouse, A. , Amar, A. , Lambert, R. , Linnell, J. , Watt, A. and Gutiérrez, R. J. 2013. Understanding and managing conservation conflicts. Trends in Ecology & Evolution , 28: 100–109. Google Scholar

284.

Reif, J. 2013. Long-term trends in bird populations: a review of patterns and potential drivers in North America and Europe. Acta Ornithologica , 48: 1–16. Google Scholar

285.

Remes, V. 2003. Effects of exotic habitat on nesting success, territory density, and settlement patterns in the blackcap (Sylvia atricapilla). Conservation Biology , 17: 1127–1133. Google Scholar

286.

Resano-Mayor, J. , Real, J. , Moleón, M. , Sánchez-Zapata, J. A. , Palma, L. and Hernández-Matías, A. 2016. Diet-demography relationships in a long-lived predator: from territories to populations. Oikos , 125: 262–270. Google Scholar

287.

Riegert, J. , Fainová, D. , Mikes, V. and Fuchs, R. 2007. How urban kestrels Falco tinunculus divide their hunting grounds: partitioning or cohabitation? Acta Ornithologica , 42: 69–76. Google Scholar

288.

Riegert, J. , Fainová, D. and Bystrická, D. 2010. Genetic variability, body characteristics and reproductive parameters of neighbouring rural and urban common kestrel (Falco tinnunculus) populations. Population Ecology , 52: 73–79. Google Scholar

289.

Rivers, J. W. , Johnson, J. M. , Haig, S. M. , Schwarz, C. J. , Burnett, I. J. , Brandt, J. , George, D. and Grantham, J. 2014. An analysis of monthly home range size in the critically endangered California condor Gymnogyps californianus. Bird Conservation International , 24: 492–504. Google Scholar

290.

Rodríguez, A. , Negro, J. J. , Bustamante, J. , Fox, J. V. and Afanasyev, V. 2009. Geolocators map the wintering grounds of threatened lesser kestrels in Africa. Diversity and Distributions , 15: 1010–1016. Google Scholar

291.

Rodríguez, A. , Negro, J. J. and Bustamante, J. 2013. Establishing a lesser kestrel colony in an urban environment for research purposes. Journal of Raptor Research , 47: 214–218. Google Scholar

292.

Rodríguez-Estrella, R. , Donázar, J. A. and Hiraldo, F. 1998. Raptors as indicators of environmental change in scrub habitat of Baja California, Mexico. Conservation Biology , 12: 921–925. Google Scholar

293.

Rohner, C. 1997. Non-territorial floaters in great horned owls (Bubo virginianus). In, J. R. Duncan , D. H. Johnson and T. H. Nicholls (Eds): Biology and Conservation of Owls of the Northern Hemisphere . 2nd International Symposium , pp. 347–362. Gen. Tech. Rep. NC-190. US Dept, of Agriculture, Forest Service. North Central Forest Experiment Station. St. Paul. Google Scholar

294.

Rubolini, D. , Gustin, M. , Bogliani, G. and Garavaglia, R. 2005. Birds and powerlines in Italy: an assessment. Bird Conservation International , 15: 131–145. Google Scholar

295.

Rutz, C. 2003. Post-fledging dispersal of northern goshawks Accipiter gentilis in an urban environment. Vogelwelt , 124: 93–101. Google Scholar

296.

Rutz, C. 2006. Home range size, habitat use, activity patterns and hunting behaviour of urban-breeding northern goshawks Accipiter gentilis. Ardea , 94: 185–202. Google Scholar

297.

Rutz, C. 2008. The establishment of an urban bird population. Journal of Animal Ecology , 11: 1008–1019. Google Scholar

298.

Salafsky, N. , Margoluis, R. and Redford, K. 2001. Adaptive Management: A Tool for Conservation Practitioners . Biodiversity Support Program. Washington D.C. Google Scholar

299.

Salamolard, M. , Butet, A. , Leroux, A. and Bretagnolle, V. 2000. Responses of an avian predator to variations in prey density at a temperate latitude. Eco logy , 81: 2428–2441. Google Scholar

300.

Salo, P. , Banks, P. B. , Dickman, C. R. and Korpimäki, E. 2010. Predator manipulation experiments: impacts on populations of terrestrial vertebrate prey. Ecological Monographs , 80: 531–546. Google Scholar

301.

Sánchez-Zapata, J. A. 2012. Overview of raptor monitoring in Spain. Acrocephalus , 33: 321–323. Google Scholar

302.

Sánchez-Zapata, J. A. , Clavero, M. , Carrete, M. , De Vault, T. , Hermoso, V. , Losada, M. A. , Polo, M. J. , Sánchez-Navarro, S. , Pérez-García, J. M. , Botella, F. , Ibáñez, C. and Donázar, J. A. 2016. Effects of renewable energy production and infrastructure on wildlife. In, R. Mateo , B. Arroyo and J. T. García (Eds.). Current Trends in Wildlife Research . Springer. London. Google Scholar

303.

Sanz-Aguilar, A. , Sánchez-Zapata, J. A. , Carrete, M. , Benítez, J. R. , Ávila, E. , Arenas, R. and Donázar, J. A. 2015. Action on multiple fronts, illegal poisoning and wind farm planning, is required to reverse the decline of the Egyptian vulture in southern Spain. Biological Conservation , 187: 10–18. Google Scholar

304.

Schaub, M. , Ullrich, B. , Knoötzsch, G. , Albrecht, P. and Meisser, C. 2006. Local population dynamics and the impact of scale and isolation: a study on different little owl populations. Oikos , 115: 389–400. Google Scholar

305.

Schlaepfer, M. A. , Runge, M. C. and Sherman, P. W. 2002. Ecological and evolutionary traps. Trends in Ecology & Evolution , 17: 474–480. Google Scholar

306.

Schumaker, N. H. , Brookes, A. , Dunk, J. R. , Woodbridge, B. , Heinrichs, J. A. , Lawler, J. J. , Carroll, C. and LaPlante, D. 2014. Mapping sources, sinks, and connectivity using a simulation model of northern spotted owls. Landscape Ecology , 29: 579–592. Google Scholar

307.

Şekercioğu, Ç. H. 2002. Impacts of birdwatching on human and avian communities. Environmental Conservation , 29: 282–289. Google Scholar

308.

Şekercioğlu, Ç. H. 2006. Increasing awareness of avian ecological function. Trends in Ecology & Evolution , 21: 464–471. Google Scholar

309.

Şekercioğlu, Ç. H. , Daily, G. C. and Ehrlich, P. R. 2004. Ecosystem consequences of bird declines. Proceedings of the National Academy of Sciences of the United States of America , 101: 18042–18047. Google Scholar

310.

Selva, N. and Cortés-Avizanda, A. 2009. Efectos de las carroñas y muladares sobre comunidades y ecosistemas. In, J. A. Donázar , A. Margalida and D. Campion (Eds.): Buitres, muladares y legislación sanitaria: perspectivas de un conflicto y sus consecuencias desde la biología de la conservación , pp. 452–473. Munibe 29 (Suplemento). Sociedad de Ciencias Aranzadi. San Sebastián. Google Scholar

311.

Selva, N. , Jedrzejewska, B. , Jedrzejewski, W. and Wajrak, A. 2003. Scavenging on European bison carcasses in BiałowiezaPrimeval Forest (eastern Poland). Ecoscience , 10: 303–311. Google Scholar

312.

Senzaki, M. and Yamaura, Y. 2015. Surrogate species versus landscape metric: does presence of a raptor species explains diversity of multiple taxa more than patch area? Wetlands Ecology and Management . ( https://doi.org/10.1007/s11273015-9469-4). Google Scholar

313.

Senzaki, M. , Yamaura, Y. and Nakamura, F. 2015. The usefulness of top predators as biodiversity surrogates indicated by the relationship between the reproductive outputs of raptors and other bird species. Biological Conservation , 191: 460–468. Google Scholar

314.

Sergio, F. , Blas, J. , Forero, M. G. , Fernández, N. , Donázar, J. A. and Hiraldo, F. 2005. Preservation of wide-ranging top predators by site-protection: black and red kites in Doñana National Park. Biological Conservation , 125: 11–21. Google Scholar

315.

Sergio, F. , Caro, T. , Brown, D. , Clucas, B. , Hunter, J. , Ketchum, J. , McHugh, K. and Hiraldo, F. 2008. Top predators as conservation tools: ecological rationale, assumptions, and efficacy. Annual review of Ecology, Evolution and Systematics , 39: 1–19. Google Scholar

316.

Sergio, F. and Hiraldo, F. 2008. Intraguild predation in raptor assemblages: a review. Ibis , 150: 132–145. Google Scholar

317.

Sergio, S. , Marchesi, L. , Pedrini, P. , Ferrer, M. and Penteriani, V. 2004. Electrocution alters the distribution and density of a top predator, the eagle owl Bubo bubo. Journal of Applied Ecology , 41: 836–845. Google Scholar

318.

Sergio, F. , Newton, I. A. N. , Marchesi, L. and Pedrini, P. 2006. Ecologically justified charisma: preservation of top predators delivers biodiversity conservation. Journal of Applied Ecology , 43: 1049–1055. Google Scholar

319.

Sergio, F. and Penteriani, V. 2005. Public information and territory establishment in a loosely colonial raptor. Ecology , 86: 340–346. Google Scholar

320.

Sergio, F. , Tanferna, A. , De Stephanis, R. , López Jiménez, L. , Blas, J. , Tavecchia, G. , Preatoni, D. and Hiraldo, F. 2014. Individual improvements and selective mortality shape lifelong migratory performance. Nature , 515: 410–413. Google Scholar

321.

Serrano, D. and Tella, J. L. 2007. The role of despotism and heritability in determining settlement patterns in the colonial lesser kestrel. American Naturalist , 169: E53–67. Google Scholar

322.

Serrano, D. and Tella, J. L. 2012. Lifetime fitness correlates of natal dispersal in a colonial bird. Journal of Animal Ecology , 81: 97–107. Google Scholar

323.

Serrano, D. , Tella, J. L. , Donázar, J. A. and Pomarol, M. 2003. Social and individual features affecting natal dispersal in the colonial lesser kestrel. Ecology , 84: 3044–3054. Google Scholar

324.

Serrano, D. , Tella, J. L. , Forero, M. G. and Donázar, J. A. 2001. Factors affecting breeding dispersal in the facultatively colonial lesser kestrel: individual experience vs. conspecific cues. Journal of Animal Ecology , 70: 568–578. Google Scholar

325.

Shivik, J. A. 2006. Are vultures birds, and do snakes have venom, because of macro- and microscavenger conflict? Bioscience , 56: 819–823. Google Scholar

326.

Sinclair, A. R. E. , Olsen, P. D. and Redhead, T. D. 1990. Can predators regulate small mammal populations? Evidence from house mouse outbreaks in Australia. Oikos , 59: 382–392. Google Scholar

327.

Singleton, G. R. , Belmain, S. R. and Brow, P. R. 2010. Rodent outbreaks: an age-old issue with a modern appraisal. In, G. R. Singleton , S. R. Belmain , P. R. Brown and B. Hardy (Eds.). Rodent Outbreaks: Ecology and Impacts. p. 289. International Rice Research Institute. Los Baños. Google Scholar

328.

Singleton, G. R. , Hinds, L. A. , Leirs, H. and Zhang, Z. 1999. Ecologically-based Rodent Management. ACIAR Monograph 59, Australian Centre for International Agricultural Research. Canberra. Google Scholar

329.

Smith, A. T. and Boyer, A. F. 2008. Oryctolagus cuniculus. The IUCN Red List of Threatened Species 2008: e.T41291A10415170. Google Scholar

330.

Sodhi, N. S. and Oliphant, L. W. 1992. Hunting ranges and habitat use and selection of urbanbreeding merlins. Condor , 94: 743–749. Google Scholar

331.

Sol, D. , Santos, D. M. and GarcÍa, J. 1998. Competition for food in urban pigeons: the cost of being juvenile. Condor , 100: 298–304. Google Scholar

332.

Soldatini, C. , Georgalas, V. , Torricelli, P. and Albores-Barajas, Y. V. 2010. An ecological approach to bird strike analysis. European Journal of Wildlife Research , 56: 623–632. Google Scholar

333.

Solonen, T. 2014a. Urbanization of the northern goshawk Accipiter gentilis in Finland. Linnut , 2014: 126–131. Google Scholar

334.

Solonen, T. 2014b: Timing of breeding in rural and urban tawny owls Strix aluco in Southern Finland: effects of vole abundance and winter weather. Journal of Ornithology , 155: 27–36. Google Scholar

335.

Solonen, T. and Ursin, K. A. 2008. Breeding of tawny owls Strix aluco in rural and urban habitats in southern Finland. Bird Study , 55: 216–221. Google Scholar

336.

Southern, W. E. 1964. Additional observations of winter bald eagle populations: including remarks on biotelemetry techniques and immature plumages. Wilson Bulletin , 76: 222–237. Google Scholar

337.

Steadman, D. W. 1995. Prehistoric extinctions of Pacific island birds: biodiversity meets zooarchaeology. Science , 267: 1123–1131. Google Scholar

338.

Sulkava, S. , Lokki, H. and Solonen, T. 2014. Diet of the northern goshawk during the breeding season in urban and rural environments in Southern Finland. Suomen Riista , 60: 20–30. Google Scholar

339.

Sullivan, A. R. , Flaspohler, D. J. , Froese, R. E. and Ford, D. 2016. Climate variability and the timing of spring raptor migration in eastern North America. Journal of Avian Biology , 47: 208–2018. Google Scholar

340.

Sunde, P. and Redpath, S. M. 2006. Combining information from range use and habitat selection: sex-specific spatial responses to habitat fragmentation in tawny owls Strix aluco. Ecography , 29: 152–158. Google Scholar

341.

Tanferna, A. , López-Jiménez, L. , Blas, J. , Hiraldo, F. and Sergio, F. 2013. Habitat selection by black kite breeders and floaters: implications for conservation management of raptor floaters. Biological Conservation , 160: 1–9. Google Scholar

342.

Tella, J. L. 2001. Action is needed now, or BSE crisis could wipe out endangered birds of prey. Nature , 410: 408. Google Scholar

343.

Tella, J. L. , Forero, M. G. , Hiraldo, F. and Donázar, J. A. 1998. Conflicts between lesser kestrel conservation and European agricultural policies as identified by habitat use analysis. Conservation Biology , 12: 593–604. Google Scholar

344.

Tella, J. L. , Hiraldo, F. , Donázar-Sancho, J. A. and Negro, J. J. 1996. Costs and benefits of urban nesting in the lesser kestrel. In, D. M. Bird , D. E. Varland and J. J. Negro . 1996. Raptors in human landscapes: Adaptations to built and cultivated environments , pp. 53–60. Academic Press. London. Google Scholar

345.

Tempel, D. J. , Peery, M. Z. and Gutiérrez, R. J. 2014. Using integrated population models to improve conservation monitoring: California spotted owls as a case study. Ecological Modelling , 289: 86–95. Google Scholar

346.

Temple, H. J. and Cox, N. A. 2009. European Red List of Amphibians . Office for Official Publications of the European Communities. Luxembourg. Google Scholar

347.

Temple, H. J. and Terry, A. 2007. The Status and Distribution of European Mammals . Office for Official Publications of the European Communities. Luxembourg. Google Scholar

348.

Terraube, J. , Guixé, D. and Arroyo, B. 2014. Diet composition and foraging success in generalist predators: Are specialist individuals better foragers? Basic and Applied Ecology , 15: 616–624. Google Scholar

349.

Therrien, J. -F. , Gauthier, G. , Pinaud, D. and Bety, J. 2014. Irruptive movements and breeding dispersal of snowy owls: a specialized predator exploiting a pulsed resource. Journal of Avian biology , 45: 536–544. Google Scholar

350.

Therrien, J. -F. , Pinaud, D. , Gauthier, G. , Lecomte, N. , Bildstein, K. L. and Bety, J. 2015. Is pre-breeding prospecting behaviour affected by snow cover in the irruptive snowy owl? A test using state-space modelling and environmental data annotated via Movebank. Movement Ecology , 3: 1. Google Scholar

351.

Thirgood, S. J. and Redpath, S. M. 2008. Hen harriers and red grouse: science, politics and human-wildlife conflict. Journal of Applied Ecology , 45: 1550–1554. Google Scholar

352.

Thirgood, S. , Redpath, S. , Haydon, D. , Rothery, P. , Newton, I. and Hudson, P. J. 2000c. Habitat loss and raptor predation: disentangling long term and short term causes of red grouse declines. Proceedings of the Royal Society of London B , 267: 651–656. Google Scholar

353.

Thirgood, S. J. , Redpath, S. M. , Newton, I. and Hudson, P. 2000a. Raptors and red grouse: conservation conflicts and management solutions. Conservation Biology , 14: 95–104. Google Scholar

354.

Thirgood, S. J. , Redpath, S. M. , Rothery, P. and Aebischer, N. 2000b. Raptor predation and population limitation in red grouse. Journal of Animal Ecology , 69: 504–516. Google Scholar

355.

Thompson, C. M. and McGarigal, K. 2002. The influence of research scale on bald eagle habitat selection along the lower Hudson River. New York (USA). Landscape Ecology , 17: 569–587. Google Scholar

356.

Thorpe, J. 2005. Fatalities and destroyed civil aircraft due to birdstrikes 2002–2004 (with an addendum of animal strikes). Proceedings of 27th International Birdstrike Committee: 17–24. Google Scholar

357.

Tilman, D. 1999. The ecological consequences of changes in biodiversity: a search for general principles. Ecology , 80: 1455–1474. Google Scholar

358.

Torre, I. , Díaz, M. , Martínez-Padilla, J. , Bonal, R. , Viňuela, J. and Fargallo, J. A. 2007. Cattle grazing, raptor abundance and small mammal communities in Mediterranean grasslands. Basic and Applied Ecology , 8: 565–575. Google Scholar

359.

Treynis, R. , Dementavičius, D. , Mozgeris, G. , Skuja, S , Rumbutis, S. and Stončius, D. 2011. Coexistence of protected avian predators: Does a recovering population of white-tailed eagle threaten to exclude other avian predators? European Journal of Wildlife Research , 57: 1165–1174. Google Scholar

360.

Trombulak, S. C. and Frissell, C. A. 2000. Review of effects of roads on terrestrial and aquatic communities. Biological Conservation , 14: 18–30. Google Scholar

361.

UIC. The Worldwide Railway Organisation. 2014. High Speed lines in the World . UIC High Speed Department.  http://old.uic.org/IMG/pdf/20140901_high_speed_lines_in_the_world.pdf (Accessed on 10/05/2016) Google Scholar

362.

Ursúa, E. , Serrano, D. and Tella, J. L. 2005. Does land irrigation actually reduce foraging habitat for breeding lesser kestrels? The role of crop types. Biological Conservation , 122: 643–648. Google Scholar

363.

Valkama, J. , Korpimäki, E. , Arroyo, B. , Beja, P. , Bretagnolle, V. , Bro, E. , Kenward, R. , Mañosa, S. , Redpath, S.M. , Thirgood, S. and Viňuela, J. 2005. Birds of prey as limiting factors of gamebird populations in Europe: a review. Biological Reviews , 80: 171–203. Google Scholar

364.

Van Moorter, B. , Rolandsen, C. M. , Basille, M. and Gaillard, J.-M. 2015. Movement is the glue connecting home ranges and habitat selection. Journal of Animal Ecology , 85: 21–31. Google Scholar

365.

Van Vuuren, D. P. , Sala, O. E. and Pereira, H. M. 2006. The future of vascular plant diversity under four global scenarios. Ecology and Society , 1: 25. Google Scholar

366.

Vasilakis, D. P. , Whitfield, P. , Schlinder, S. , Poirazidis, K. S. and Kati, V. 2016. Reconciling endangered species conservation with wind farm development: cinereous vultures (Aegypius monachus) in south-eastern Europe. Biological Conservation , 196: 10–17. Google Scholar

367.

Verbürg, P. and Overmars, K. 2009. Combining top-down and bottom-up dynamics in land use modeling: exploring the future of abandoned farmlands in Europe with the Dyna-CLUE model. Landscape Ecology , 24: 1167–1181. Google Scholar

368.

Villafuerte, R. , Viňuela, J. and Blanco, J. C. 1998. Extensive predator persecution caused by population crash in a game species: the case of red kites and rabbits in Spain. Biological Conservation , 84: 181–188. Google Scholar

369.

Virgós, E. and Travaini, A. 2005. Relationship between small-game hunting and carnivore diversity in central Spain. Biodiversity and Conservation , 14: 3475–3486. Google Scholar

370.

Vitousek, P. M. , Mooney, H. A. , Lubchenco, J. and Melillo, J. M. 1997. Human domination of earth's ecosystems. Science , 277: 494–499. Google Scholar

371.

Warkentin, I. G. and Oliphant, L. W. 1990. Habitat use and foraging behaviour of urban merlins (Falco columbarius) in winter. Journal of Zoology , 221: 539–563. Google Scholar

372.

Watson, J. 1992. Status of the golden eagle Aquila chrysaetos in Europe. Bird Conservation International , 2: 175–183. Google Scholar

373.

Wenny, D. G. , Devault, T. L. , Johnson, M. D. , Kelly, D. , ŞekercioĞlu, Ç. H. , Tomback, D. F. and Whelan, C. J. 2011. The need to quantify ecosystem services provided by birds. Auk , 128: 1–14. Google Scholar

374.

Whelan, C. J. , ŞekercioĞlu, Ç. H. and Wenny, D. G. 2015. Why birds matter: from economic ornithology to ecosystem services. Journal of Ornithology , 156: 227–238. Google Scholar

375.

Whelan, C. J. , Wenny, D. G. and Marquis, R. J. 2008. Ecosystem services provided by birds. Annals of the New York Academy of Sciences , 1134: 25–60. Google Scholar

376.

Whitfield, D. P. , Fielding, A. H. , McLeod, D. R. A. and Haworth, P. F. 2004. The effects of persecution on age of breeding and territory occupation in golden eagles in Scotland. Biological Conservation , 118: 249–259. Google Scholar

377.

Whitfield, D. P. , McLeod, D. R. A. , Watson, J. , Fielding, A. H. and Haworth, P. F. 2003. The association of grouse moor in Scotland with the illegal use of poisons to control predators. Biological conservation , 114: 157–163. Google Scholar

378.

Wieringa, K. (Ed.) 1995. Environment in the European Union 1995: Report for the Review of the Fifth Environmental Action Programme. European Environment Agency- EUROSTAT. Google Scholar

379.

Wiklund, C. G. 1996. Determinants of dispersal in breeding merlins (Falco columbarius). Ecology , 77: 1920–1927. Google Scholar

380.

Williams, D. R. , Pople, R. G. , Showler, D. A. , Dicks, L. V. , Child, M. F. , Erasmus K. H. J ., Ermgassen, Z. and Sutherland, W. J. 2012. Bird Conservation: Global Evidence for the Effects of Interventions. Pelagic Publishing, Exeter. Google Scholar

381.

Wilmers, C. C. , Nickel, B. , Bryce, C. M. , Smith, J. A. , Wheat, R. E. and Yovovich, V. 2015. The golden age of bio-logging: how animalborne sensors are advancing the frontiers of ecology. Ecology , 96: 1741–1753. Google Scholar

382.

Wilmers, C. C. and Post, E. 2006. Predicting the influence of wolf-provided carrion on community dynamics under climate change scenarios. Global Change Biolology , 12: 403–409. Google Scholar

383.

Wilson, E. E. and Wolkovich, E. M. 2011. Scavenging: how carnivores and carrion structure communities. Trends in Ecology & Evolution , 26: 129–135. Google Scholar

384.

Wood, B. J. and Fee, C. G. 2003. A critical review of the development of rat control in Malaysian agriculture since the 1960s. Crop Protection , 22: 445–461. Google Scholar

385.

Zuberogoitia, I. 2011. Weather influence on breeding success of the tawny owl on the southwest limit of Eurosiberian Region. In, I. Zuberogoitia and J. E. Martínez (Eds): Ecology and Conservation of European Forest-Dwelling Raptors , pp. 184–189. Diputación Foral de Bizkaia. Bilbao. Google Scholar

386.

Zuberogoitia, I. , del Real, J. , Torres, J. J. , Rodríguez, L. , Alonso, M. , de Alba, V , Azahara, C. and Zabala, J. 2015. Testing pole barrier as feasible mitigation measure to avoid bird vehicle collosions (BVC). Ecological Engineering , 83: 144–151. Google Scholar

387.

Zuberogoitia, I. , Martínez, J. E. , González-Oreja, J. A. , Calvo, J. F. and Zabala, J. 2013. The relationship between brood size and prey selection in a peregrine falcon population located in a strategic region on the western European fly way. Journal of Ornithology , 154: 73–82. Google Scholar

388.

Zuberogoitia, I. , Martínez, J. E. , Margalida, A. , Gómez, I. , Azkona, A. and Martínez, J. A. 2010. Reduced food availability induces behavioural changes in griffon vulture. Ornis Fennica , 87: 52–60. Google Scholar

389.

Zuberogoitia, I. , Ruiz Moneo, F. and Torres, J. J. (Eds). 2002. El Halcón Peregrino. Diputación Foral de Bizkaia. Bilbao. Google Scholar
José A. Donázar, Ainara Cortés-Avizanda, Juan A. Fargallo, Antoni Margalida, Marcos Moleón, Zebensui Morales-Reyes, Rubén Moreno-Opo, Juan M. Pérez-García, José A. Sánchez-Zapata, Iñigo Zuberogoitia, and David Serrano "Roles of Raptors in a Changing World: From Flagships to Providers of Key Ecosystem Services," Ardeola 63(1), 181-234, (1 June 2016). https://doi.org/10.13157/arla.63.1.2016.rp8
Published: 1 June 2016
KEYWORDS
animal ecology
aves de presa
birds of prey
cambio global
conflictos con fauna salvaje
conservación
conservation
Back to Top