Next Article in Journal
Sectoral Decomposition of Korea’s Energy Consumption by Global Value Chain Dimensions
Next Article in Special Issue
Drainage Ditch Berm Delineation Using Lidar Data: A Case Study of Waseca County, Minnesota
Previous Article in Journal
Successful Implementation of Climate-Friendly, Nutritious, and Acceptable School Meals in Practice: The OPTIMAT Intervention Study
Previous Article in Special Issue
Hydrological Effects of Urban Green Space on Stormwater Runoff Reduction in Luohe, China
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Assessment of Geomorphosites for Geotourism in the Northern Part of the “Ruta Escondida” (Quito, Ecuador)

by
Paúl Carrión-Mero
1,2,*,
Alicia Ayala-Granda
3,
Sthefano Serrano-Ayala
3,
Fernando Morante-Carballo
1,4,5,
Maribel Aguilar-Aguilar
1,
Miguel Gurumendi-Noriega
6,
Nataly Paz-Salas
6,
Gricelda Herrera-Franco
5,7 and
Edgar Berrezueta
8,*
1
Centro de Investigaciones y Proyectos Aplicados a las Ciencias de la Tierra (CIPAT), ESPOL Polytechnic University, Guayaquil P.O. Box 09-01-5863, Ecuador
2
Facultad de Ingeniería en Ciencias de la Tierra, ESPOL Polytechnic University, Guayaquil P.O. Box 09-01-5863, Ecuador
3
Sacharxeos Cía. Ltda. Tulipanes 153 y Arupos Capelo Valle de los Chillos, Cantón Rumiñahui 170501, Ecuador
4
Facultad de Ciencias Naturales y Matemáticas, ESPOL Polytechnic University, Guayaquil P.O. Box 09-01-5863, Ecuador
5
Geo-Recursos y Aplicaciones GIGA, ESPOL Polytechnic University, Guayaquil P.O. Box 09-01-5863, Ecuador
6
Bira Bienes Raíces S.A. (BIRA S.A.), Av. Alfonso de Mercadillo, Zaruma 071350, Ecuador
7
Facultad de Ciencias de la Ingeniería, Universidad Estatal Península de Santa Elena (UPSE), La Libertad 240204, Ecuador
8
Departamento de Infraestructura Geocientífica y Servicios, Instituto Geológico y Minero de España (IGME), 33005 Oviedo, Spain
*
Authors to whom correspondence should be addressed.
Sustainability 2020, 12(20), 8468; https://doi.org/10.3390/su12208468
Submission received: 17 September 2020 / Revised: 5 October 2020 / Accepted: 7 October 2020 / Published: 14 October 2020
(This article belongs to the Special Issue Landscape Planning for Sustainability)

Abstract

:
The relevant geomorphological characteristics of territory represent an essential part of its natural heritage. They are also an asset to be exploited for stimulating socio-economic development. The “Ruta Escondida” in Ecuador constitutes a historical place full of culture and landscapes that have been shaped over time by geological and geomorphological processes. Among the geomorphological features of the study area, volcanic cones, hilltops, terraces, foothills and glacial valleys stand out. The aims of this work were: (1) to characterize 18 places of geomorphological interest, located in the northern part of the Ruta Escondida and (2) to propose alternatives (geotourism) to contribute to the local development of the area. The applied methodology included: (1) the compilation of geomorphological elements; (2) the assessment of geomorphosites using the Inventario Español de Lugares de Interés Geológico (IELIG) method and (3) a strengths–opportunities–weaknesses–threats analysis of the contribution and influence of geomorphosites in the development of the study area. With this work, it was possible to determine that all the analyzed geomorphological sites have a high and very high interest. The strengths, weaknesses, opportunities, and threats (SWOT) analysis revealed that the geomorphosites could provide significant added value to the development of geotourism on the route, complementing the already known cultural and historical attractions.

1. Introduction

Geodiversity, as a concept analogous to biodiversity, was introduced in the early 1990s [1,2,3,4]. In recent years, this term became of everyday use in inventory studies, valuation and conservation of geological heritage [5,6,7,8]. According to Gray [9], geodiversity can be defined as “the natural range (diversity) of geological (rocks, minerals, fossils), geomorphological (landform, processes) and soil features”. It also refers to the assemblage of these features, their relationships, properties, interpretations, and systems. Geodiversity can also be considered as the basis of the increasing biological diversity existing in a territory [10,11]. Elements of geodiversity depend on and are directly exposed to human activities, with the consequent danger of their degradation. Thus, geodiversity assessment actions of recent years have been aiming to accumulate data that allow estimating its evolution. The assessment of geodiversity is carried out through both qualitative [9,12,13,14,15] and quantitative [10,16,17,18,19,20,21,22,23,24] assessment methods. These procedures make it possible to propose conservation plans (e.g., geoparks), to identify areas of interest (e.g., geological heritage) and to regulate their sustainable use (e.g., geotourism) [8,25,26].
According to Brilha [15], geological heritage is defined as geodiversity elements that have a unique scientific value. Geoheritage includes landforms, geomorphological, mining and glacial features [27]. The geomorphological heritage is part of the natural heritage of territorial and landscape relevance [28]. Figure 1 explains and correlates geodiversity and related concepts in a simplified framework [15].
Different methods have been established that allow the assessment of geomorphological heritage. These methods also enable the promotion of geomorphosites and their protection against the effect of anthropic activity [29]. Geomorphological heritage is a territorial resource that forms part of the geological heritage of a territory, therefore, its study supports sustainable development. It also has significant value regarding the landscape [30].
Geosites are sites with geological and geomorphological interest that determine geological heritage and promote its conservation. These places are mainly located in rural areas, outdoors, although they are also found in urban environments [31]. Geomorphosites are places with a specific relief or geomorphological resource, considered as vestiges of dynamic processes in the history of the Earth, to which landscape, scientific, cultural and socio-economic value can be attributed, and are useful for the community [32,33,34,35]. Geosites are considered unique elements or vast, vulnerable panoramas that must be protected by a legal framework [36], focused on conservation, education and sustainable development [37].
According to Reynard and Paniza [36], geomorphosites are particularly susceptible to the lack of protection measures, and four types of actions can be taken to reduce their vulnerability: (1) to improve assessment methods that allow the objective selection of sites of high interest; (2) to enhance public awareness of the geomorphological value of the territory through education; (3) to promote management structures, such as geoparks; (4) to improve the legal basis for protection, which can be enforced either through property rights or using a public policy [36].
The conservation of geosites must be based on their audit, assessment and the definition of the need for specific actions [38]. More globally, geoconservation focuses on the management of geological components of scientific, touristic and cultural importance within a responsible social context [39,40]. Adequate management helps to protect the geological heritage adequately [41]. Furthermore, scientific understanding by the community is essential to implement conservation strategies [42].
Geotourism is a form of natural tourism that focuses explicitly on geology and landscape. It promotes tourism to geosites and the conservation of geodiversity and an understanding of earth sciences through appreciation and learning [43]. This is achieved through independent visits to geological features, use of geo-trails and viewpoints, guided tours, geo-activities and patronage of geosite visitor centers [44]. Geotourism is an activity that focuses on the proper organization and management of geosites and on the generation of environmental, community and economic benefits [44].
The study region is the northern part of the Ruta Escondida, located on the north of the province of Pichincha (Ecuador). It is situated in an area with an outstanding geomorphological character. The geomorphological features of the territory are derived mainly from volcanic activity (e.g., lahars, lava flows, volcanic craters, columnar disjunction) and, also, from processes related to glaciation (e.g., cirques, moraines, glacial valleys). All of these features are typical results of the geological and geomorphological evolutions of the Ecuadorian Andes [45], which is a country with a rich and recognized geodiversity [46].
Despite the singularities of the landforms and the efforts of regional and local authorities, the geotourism impact of these places is limited. The main reason could be the absence of a strategic plan that includes the inventory, characterization and promotion of the morphological peculiarities of the territory.
The main aims of this study are (1) to characterize the geomorphological places of interest in the area through a semi-quantitative assessment by the Inventario Español de Lugares de Interés Geológico (IELIG) method and (2) to apply a strengths, weaknesses, opportunities, and threats (SWOT) matrix analysis that allows defining the influence of the inclusion of geomorphosites as a resource of the Ruta Escondida. Based on the characterization and assessment, strategies are proposed to facilitate the implementation of geotouristic activities as a basis for local development.

2. Geographical and Geological Setting

The study area is located within the Ruta Escondida, in the province of Pichincha, in the north of Ecuador (Figure 2a). The zone is bordered by the Imbabura province to the north, by Quito city to the south and west, and by Pedro Moncayo city to the east [47] (Figure 2b). The Ruta Escondida consists of five parishes: Atahualpa, Chavezpamba, Perucho, Puéllaro and San José de Minas that belong to the Quito canton (Figure 2b) and occupies an area of approximately 467 km2 [48]. The altitude of the site is between 1537 and 3500 meters above sea level and temperature ranges between 16 and 30 °C [49,50]. This work covers the Atahualpa and San Jose de Minas parishes (Figure 2) explicitly.
At a regional level, the Ecuadorian Andes is part of a volcanic arc generated by the subduction that occurs between the Nazca plate and the continental lithosphere of South America in Miocene. In general, three main groups are distinguished in the Ecuadorian Andes: (i) frontal arc (western Cordillera), (ii) main arc (eastern Cordillera) and (iii) back arc (Oriente) [51,52]. Approximately 50 volcanoes, located in an area of an approximate length of 300 km from N–S and a width of 100–120 km from E–W, constitute the volcanic arc in front of Carnegie Ridge [53].
The volcanic complex consisting of the Mojanda and Fuya Fuya volcanoes is located 50 km NNE of Quito, in the Interandean valley (depression between the eastern and western Cordilleras that originates from the Miocene). The distance between the volcanoes is 3 km [54]. The formation of the Mojanda volcano began approximately 0.6 Ma ago and ended 0.18–0.20 Ma [55] or 0.165 Ma [56] ago. The formation of the Fuya Fuya volcano occurred less than 0.5 Ma ago, and continued until the Maximum Glaciation [57], and even until the Holocene [58].
Mojanda consists of two cones (Lower Mojanda and Upper Mojanda), located southwest of Cushnirumi Peak (the oldest dissected volcano). The Lower Mojanda is a basal cone of 16 × 18 km, in which an ancient caldera (Caldera 1) of 5 km wide is evidenced, which separates this structure from Upper Mojanda. Lower Mojanda consists mainly of andesitic lavas with a high content of silica and two-pyroxene andesites (Unit M-I-1). Lava flows of amphibole-bearing dacite (Unit M-I-2) can also be found [53,58] (Figure 3).
Upper Mojanda began to form with basaltic to high-silica andesite lava flows (M-II-1), followed by pyroclastic flows of basaltic andesite (ash-and-slag flows, M-II-2) and andesite block-and-ash flows (Unit M-II-3). After the phase of volcanic activity, the slopes were covered by a stratified sequence of breccias of approximately 300 m thickness, with a basaltic andesite composition, interlayered with lavas and fallout deposits (Unit M-II-4). Finally, the slopes of Mojanda are covered by a sequence of ash and lapilli layers (Unit M-II-5), which underlie a Pifo ash-fall layer, product of the partial destruction of the cone and the formation of the caldera (Caldera 2) [53,58] (Figure 3).
Fuya Fuya consists of three volcanic cones: Lower Fuya Fuya, San Bartolo and Upper Fuya Fuya. The Lower Fuya Fuya is constituted of high-silica andesitic to dacitic domes, with high silica content and coarse lavas (FF-I Unit). Based on petrologic data, two eroded eccentric domes—the Puellaro and San Jorge–are associated with the FF-I unit. Ash deposits and blocks are found to the SSW of the volcano; these form the FF-II unit [53,58] (Figure 3).
The San Bartolo cone is the second construction phase of the volcano and is composed of andesitic lavas (FF-III Unit). A major collapse in the sector caused a debris avalanche (FF-IV-1 Unit) from the San Bartolo cone, which flowed westward reaching a depth of approximately 100 m. The debris avalanche was accompanied by dacitic ash flows with pumice blocks (FF-IV-2 Unit), which expanded to the west, with a thickness of 50 m. Finally, the collapse of the San Bartolo cone ended with the emission of a sequence of pumice-lapilli fall layers interspersed with pyroclastic deposits (FF-IV-3 Unit), with thicknesses of approximately 20 m [53,58] (Figure 3).
Upper Fuya Fuya consists of two series of thick, glacially eroded, lavas and domes (FF-V-1 Unit). The lava series consists of hornblende dacite, while the domes contain andesite and dacite. Then, block-and-ash flows and pyroclastic flows were directed towards the west and south, while numerous fall deposits were deposited in the upper part of the volcanic complex (FF-V-2). Finally, three central lavas and the Colangal and Panecillo domes represent the last extrusions of the Fuya Fuya (FF-V-3). In these lavas and domes, no glacial erosion is evidenced, based on which Holocene age was established for their formation [53,58] (Figure 3).
In Ecuador, according to Clapperton [59] and Schubert and Clapperton [60], the limit of glaciation is 3000 to 3600 m. The moraines have been dated between 33,000 and 43,000 BP, associated with the last ice age. However, volcanic activity makes observations on glacial deposits difficult [61]. Some geoforms present in the study area are considered products of the mini-glaciation or “little ice age”, which occurred in the 14th and 19th centuries [62,63,64], the same one that could have shaped the terrain on the volcanic formations.

3. Materials and Methods

The procedure described in the present study comprised three phases: (i) compilation of the geomorphological element register, (ii) assessment of geomorphosites and (iii) preparation of the SWOT analysis matrix (Figure 4).

3.1. Geomorphological Element Register

In this first phase, existing information was reviewed and processed. Specifically, we compiled thematic cartography and preliminary studies developed in the area [48,49,50,65,66], together with the existing geological data, to select a potential site list to be considered as geomorphosites. In addition, base thematic mapping (geographic and geological) was carried out for the selected geomorphosite candidates.

3.2. Geomorphosite Assessment

This phase focused on the assessment of the scientific, didactic and tourist interest of the geomorphosites, through the application of the IELIG methodology (acronym in Spanish for “Inventario Español de Lugares de Interés Geológico”) [14]. This method identifies sites of geological interest based on the assessment of a wide range of parameters, giving criteria scores between 0 and 4, where 0 is the lowest, and four is the highest [14]. These parameters are grouped into four sections: (i) values of interest (scientific (Sc), academic (Ac) and tourist or recreational (To)) (Table 1), (ii) fragility (Fr), (iii) vulnerability (Vul) and (iv) protection priority (Pp). Pp is determined from the degradation susceptibility (DS) calculation based on Fr and Vul produced by external threats (Table 2). This parameter complements the total interest values [14]. According to García et al. [14], the value of Pp on the different axes (i.e., scientific Pp (Sc), academic Pp (Ac) and tourist Pp (To)) is obtained by multiplying the square of the value of their interest with the area susceptibility degree (DS) and divided by 4002 to keep the same scale. Finally, the global priority Pp value is based on the average of the scores of interest multiplied by the degree of susceptibility (DS) and by (1/(400)2) as detailed in Table 3.

3.3. SWOT Analysis Matrix and Proposal for Geotourism

A strengths, weaknesses, opportunities, and threats (SWOT) analysis [67] was conducted taking into account the list of assessed sites of geological interest. This analysis made it possible to determine the potential that geomorphosites present for the development of geotourism. The analysis was carried out with the participation of members of the academy, the public sector, and the private sector, all of them with interest and knowledge of the topic. The aim was to define strategies that optimize geotourism in the Ruta Escondida.

4. Results

4.1. Geomorphological Element Register

Based on the information collected from 36 places of geomorphological interest [65], the north of the Ruta Escondida was decided upon as the target area. Specifically, 18 sites with great geomorphological interest were studied (Figure 5). The selected geomorphosites have unique geomorphological characteristics that were determined from a general description. Two of the evaluated sites (15 and 16) are located outside the province of Pichincha. Table 4 shows the selected places and a brief description of their main characteristics. In Figure 6, six of the identified geomorphosites are presented outlining their outstanding morphological features. This register is the starting point of the evaluation of scientific, academic and tourist potential.

4.2. Geomorphosite Assessment

The global results obtained from the average values of the interest types (Sc, Ac and To) of the 18 points evaluated by the IELIG method are presented in Table 5. The average interest (Sc + Ac + To/3) calculated for each one of the geomorphosites shows that the maximum value (267/400, Table 5) corresponds to the Mojanda lagoon (Figure 6a) and the Panecillo dome. The minimum value (97/400, Table 5) corresponds to Atahualpa ancient colluviums. In general, 5.56% of geomorphosites have very high interest, 33% high interest and 61.11% medium interest (Figure 7).
The obtained susceptibility results of the 18 sites, based on their fragility and vulnerability, are presented in Table 6. The results reflect that 72.22% of the sites have low susceptibility to degradation with a minimum value of 0/400. The remaining 27.78% have medium susceptibility with maximum values of 36.25/400 (Table 6 and Figure 8).
According to the results of the degree of interest and susceptibility to degradation, 72% of the sites have a low protection priority value (0–0.64/400), while the remaining 28% present medium PP (1.46–4.69/400) (Table 7 and Figure 9).

4.3. SWOT Analysis Matrix

The SWOT analysis of the geomorphosites made it possible to identify the main strengths, opportunities, weaknesses and threats in the study area, in order to generate strategies by combining internal (strengths and weaknesses) and external (opportunities and threats) characteristics, as summarized in Table 8.

4.4. Proposed Itinerary to Visit Geomorphosites

Based on the described data, we propose the development of a specific itinerary of the geomorphosites in the Ruta Escondida called “Rocks & Water in Ruta Escondida” (Figure 10).
The proposed itinerary is one example among several potential alternatives considering the sites inventoried in this study. The route matches the following criteria: (i) accessibility to every selected geomorphosite with a motor vehicle/walking; (ii) pleasant and attractive tour with reasonably short distances between sites of interest. Four possible accesses are proposed to the route (Figure 9) that allow visiting the geomorphosites and also integrate cultural attractions and biodiversity features.
Access A is an example of a touristic tour. The tour begins with a visit to Quito city, then continues to the Perucho parish, followed by the Chavezpamba parish, to end with the Atahualpa parish and part of Pedro Moncayo city. Through these places, the tourist can visit different geomorphosites, including specific sites with natural or cultural wealth (Table 9).
The circuit can be completed in one or two days, with the possibility of visiting other sites nearby. A general assessment of the proposed route from the average values of every suggested site is presented in Table 10. The results reveal the significance of this geotourism route and its potential contribution to the regional tourism offer. A complete visit to all the inventoried geomorphosites would take approximately four days.

5. Interpretation of Results and Discussion

The Ruta Escondida, located in a singular volcanic complex, holds unique geomorphological landscapes with geotouristic potential. This potential is reflected in the assessment of 18 geomorphosites (Table 4), according to the IELIG method [14]. The obtained results show that 5.56% of sites present a very high average interest (AI) value, highlighting the Laguna del Mojanda, which, due to its geological nature, is the most outstanding geomorphosite in the area (Table 5 and Figure 6). The “knowledge of the site” parameter is the one that generally obtains the lowest score due to the lack of geomorphosite studies that could promote their potential at national and international level.
The SD results of the sites are classified in the low to medium categories (Table 6 and Figure 7), which translates into low to medium PP values (Table 7 and Figure 8). This is due to the large extension of the sites, thanks to which the threat of anthropic activity is limited despite the relative proximity to communities. This also means that minimal protection measures are required.
One of the advantages of applying the method to a set of geomorphosites is the identification of weak points in the academic, scientific, tourist and SD aspects. The applied methodology allowed a detailed analysis of the studied geomorphosites. According to Štrba et al. [68,69], the application of several methods generates different results, rendering it essential to assess the sites of interest by a combination of various methodologies to obtain a more complete assessment. To improve the results obtained, the authors recommend the inclusion of one or more geomorphosite assessment methodologies (e.g., [18,34,70,71,72,73,74,75]). In this work, only one assessment method (IELIG) was used, because it is the base methodology recommended by the ASGMI (Ibero–American Association of Geological and Mining Surveys) for assessing geological interest sites [76]. Moreover, this methodology has already been employed in the other points assessment of geological interest in Ecuador (e.g., [8,26,77,78,79,80,81,82]). In general, the work carried out in the study area regarding the inventory, valuation, promotion, protection and use of geosites is similar to the approach adopted in other countries (e.g., [71,83,84,85,86,87,88,89,90]).
The semi-quantitative assessment of the geomorphosites formed the basis for the SWOT matrix analysis. Through this analysis the potential of sites of geomorphological interest of the Ruta Escondida was determined, as an alternative for the geotourism development in the area. In addition, guidelines or action strategies have been established in order to broaden the tourism attraction of the route through the adaptation of geomorphosites that highlight the importance and geological beauty of the area and its relationship with the biodiversity and culture. For better use of the sites in tourist activities, the specific typology (sizes) of the geomorphosites could be considered. However, in this first study, they were all grouped as a complex typology.
According to Kubalíková [91], the SWOT analysis is a qualitative assessment tool that integrates the knowledge of experts with the perception of the local population, and which provides meaningful information for authorities interested in new development alternatives. This analysis has already been used in the qualitative evaluation of places of geological interest in different zones [71,91,92,93,94,95].
Based on the inventory and assessment of the geomorphosites present in the study area and the SWOT analysis, a series of initiatives are proposed with the aim of facilitating the implementation of geotourism activities as a basis for local development. These proposals could be summarized as follows:
  • To correlate the inventories of this work with others obtained in nearby areas, such as the one by Ayala [50], and try to unify assessment criteria [14].
  • To promote the development of local biological, geological, archaeological and paleontological research. This initiative will allow us to recognize the scientific importance of the area at a national and international level, as well as its direct link with tourism and education.
  • To promote geomorphosites as a geotourism alternative that allows the sustainable development of the area. For this, the provincial, cantonal and parochial authorities need to work with academic and business support, on the creation and adaptation of sites as tourist destinations.
  • To strengthen the quality of services offered to the visitor, through the conditioning of the road infrastructure and facilities that ensure access to each geomorphosite. Additionally, to offer routes that allow visiting various geomorphosites with general and particular information (difficulty, distance, type of access).
  • To reinforce territorial development plans through the implementation of regulation that ensure the conservation of geomorphosites.
  • To determine the rate of erosion caused by the new activities (geotourism) and to propose initiatives that minimize the impact.
  • To implement other tourist activities (trail travel, cycling, sale of handicrafts, typical food and culture) at the geomorphosites to broaden the touristic offer of the Ruta Escondida.

6. Conclusions

From the methodological point of view, the performed assessment fosters the scientific dissemination of the geological potential of the Ruta Escondida as an alternative for economic development and as one of the essential geotourism destinations of the country.
The obtained results reveal that the northern part of the route is characterized by considerable scientific, educational and tourist interest, reflected by the average AI values that range between 97 and 267 (medium to very high interest). The SD and PP values also guarantee that it is feasible to add the studied area to the so-called Ruta Escondida, offering more and more varied alternatives for tourists through the integration of geological, biological, and cultural wealth. The semi-quantitative assessment by the IELIG method and the SWOT matrix analysis carried out in the north of the Ruta Escondida reflect a high potential as an alternative in geotourism development (e.g., suggested itinerary “Rocks & Water in Ruta Escondida” proposed in this article). The researchers recommend enhancing the assessed sites through (i) development of scientific research, (ii) reformulation of territorial development plans that ensure the protection of the area, and (iii) improvement of site conditions and complementary activities for tourists. These actions would contribute to the improvement of the quality of life of the people in the area.

Author Contributions

Conceptualization, P.C.-M., G.H.-F., F.M.-C., and M.A.-A.; methodology, P.C.-M., G.H.-F., F.M.-C., M.A.-A., and E.B.; investigation, A.A.-G., S.S.-A., M.G.-N., N.P.-S.; writing—original draft preparation, M.A.-A., and E.B.; writing—review and editing, P.C.-M., G.H.-F., F.M.-C., M.A.-A., M.G.-N., N.P.-S., G.H.-F. and E.B.; supervision, P.C.-M. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by ESPOL Polytechnic University research project called “Registry of geological and mining heritage and its impact on the defense and preservation of geodiversity in Ecuador, CIPAT–01–2018”.

Acknowledgments

This work has been made possible thanks to support from “Museo de Perucho: Arqueología e Historia, Quito-Ecuador”, Sacharxeos CIA.LTDA, BIRA Bienes Raices S.A. Company and Archeology of the Neotropics Master of ESPOL Polytechnic University. Furthermore, we thank the reviewers and editorial committee for their observations and suggestions to improve the manuscript.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Sharples, C. A Methodology for the Identification of Significant Landforms and Geological Sites for Geoconservation Purposes; Forestry Commission Tasmania: Burnie, Australia, 1993. [Google Scholar]
  2. Wiedenbein, F.W. Origin and use of the term ‘geotope’ in German-speaking countries. In Geological and Landscape Conservation; O’Halloran, D., Green, C., Harley, M., Knill, J., Eds.; Geological Society: London, UK, 1994; pp. 117–120. [Google Scholar]
  3. Semeniuk, V. The linkage between biodiversity and geodiversity. In Pattern and Processes: Towards a Regional Approach to National Estate Assessment of Geodiversity; Eberhard, R., Ed.; Environment Australia: Canberra, Australia, 1997; pp. 51–58. [Google Scholar]
  4. Joyce, E. Assessing geological heritage. In Pattern and Process: Towards a Regional Approach to National Estate Assessment of Geodiversity; Eberhard, R., Ed.; Environment Australia: Canberra, Australia, 1997; Volume 2, pp. 35–40. [Google Scholar]
  5. Erikstad, L. Geoheritage and geodiversity management—The questions for tomorrow. Proc. Geol. Assoc. 2013, 124, 713–719. [Google Scholar] [CrossRef]
  6. Carcavilla, L.; Durán, J.J.; López-Martínez, J. Geodiversity: Concept and relationship with geological heritage. Geo-Temas 2008, 10, 1299–1303. [Google Scholar]
  7. Carcavilla, L.; Durán, J.J.; García-Cortés, Á.; López-Martínez, J. Geological Heritage and Geoconservation in Spain: Past, Present, and Future. Geoheritage 2009, 1, 75–91. [Google Scholar] [CrossRef]
  8. Carrión Mero, P.; Herrera Franco, G.; Briones, J.; Caldevilla, P.; Domínguez-Cuesta, M.J.; Berrezueta, E. Geotourism and local development based on geological and mining sites utilization, Zaruma-Portovelo, Ecuador. Geosciences 2018, 8, 205. [Google Scholar] [CrossRef]
  9. Gray, M. Defining Geodiversity. In Geodiversity: Valuing and Conserving Abiotic Nature; John Wiley & Sons: Chichester, UK, 2004; pp. 1–9. [Google Scholar]
  10. Kozowski, S. Geodiversity. The concept and scope of geodiversity. Prz. Geol. 2004, 52, 833–837. [Google Scholar]
  11. Quesada-Román, A.; Pérez-Umaña, D. State of the art of geodiversity, geoconservation, and geotourism in Costa Rica. Geosciences 2020, 10, 211. [Google Scholar] [CrossRef]
  12. Brilha, J.; Gray, M.; Pereira, D.I.; Pereira, P. Geodiversity: An integrative review as a contribution to the sustainable management of the whole of nature. Environ. Sci. Policy 2018, 86, 19–28. [Google Scholar] [CrossRef] [Green Version]
  13. Miljković, Ð.; Božić, S.; Miljković, L.; Marković, S.B.; Lukić, T.; Jovanović, M.; Bjelajac, D.; Vasiljević, Đ.A.; Vujičić, M.D.; Ristanović, B. Geosite Assessment Using Three Different Methods; a Comparative Study of the Krupaja and the Žagubica Springs—Hydrological Heritage of Serbia. Open Geosci. 2018, 10, 192–208. [Google Scholar] [CrossRef]
  14. García-Cortés, Á.; Carcavilla Urquií, L.; Apoita Mugarza, B.; Arribas, A.; Bellido, F.; Barrón, E.; Delvene, G.; Díaz-Martínez, E.; Díez, A.; Durán, J.J.; et al. Documento Metodológico para la Elaboración del Inventario Español de Lugares de Interés Geológico (IELIG); Propuesta Para la Actualización Metodológica; Instituto Geológico y Minero de España: Madrid, Spain, 2013; pp. 1–64. [Google Scholar]
  15. Brilha, J. Inventory and Quantitative Assessment of Geosites and Geodiversity Sites: A Review. Geoheritage 2016, 8, 119–134. [Google Scholar] [CrossRef] [Green Version]
  16. Kot, R. The point bonitation method for evaluating geodiversity: A guide with examples (polish lowland). Geogr. Ann. Ser. A Phys. Geogr. 2015, 97, 375–393. [Google Scholar] [CrossRef] [Green Version]
  17. Ruban, D.A. Quantification of geodiversity and its loss. Proc. Geol. Assoc. 2010, 121, 326–333. [Google Scholar] [CrossRef]
  18. Reynard, E.; Fontana, G.; Kozlik, L.; Scapozza, C. A method for assessing scientific and additional values of geomorphosites. Geogr. Helv. 2007, 62, 148–158. [Google Scholar] [CrossRef]
  19. Serrano, E.; Ruiz-Flaño, P. Geodiversity: A theoretical and applied concept. Geogr. Helv. 2007, 62, 140–147. [Google Scholar] [CrossRef]
  20. Pereira, D.I.; Pereira, P.; Brilha, J.; Santos, L. Geodiversity Assessment of Paraná State (Brazil): An Innovative Approach. Environ. Manag. 2013, 52, 541–552. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  21. Melelli, L. Geodiversity: A new quantitative index for natural protected areas enhancement. Geoj. Tour. Geosites 2014, 13, 27–37. [Google Scholar]
  22. Jačková, K.; Romportl, D. The Relationship Between Geodiversity and Habitat Richness in Šumava National Park and Křivoklátsko PLA (Czech Republic): A Quantitative Analysis Approach. J. Landsc. Ecol. 2008, 1, 23–38. [Google Scholar] [CrossRef] [Green Version]
  23. Zwoliński, Z. The routine of landform geodiversity map design for the Polish Carpathian Mts. Landf. Anal. 2009, 11, 77–85. [Google Scholar]
  24. Forte, J.P.; Brilha, J.; Pereira, D.I.; Nolasco, M. Kernel Density Applied to the Quantitative Assessment of Geodiversity. Geoheritage 2018, 10, 205–217. [Google Scholar] [CrossRef] [Green Version]
  25. Nobre da Silva, M.L.; Leite do Nascimento, M.A.; Leite Mansur, K. Quantitative Assessments of Geodiversity in the Area of the Seridó Geopark Project, Northeast Brazil: Grid and Centroid Analysis. Geoheritage 2019, 11, 1177–1186. [Google Scholar] [CrossRef]
  26. Herrera-Franco, G.; Carrión-Mero, P.; Alvarado, N.; Morante-Carballo, F.; Maldonado, A.; Caldevilla, P.; Briones-Bitar, J.; Berrezueta, E. Geosites and georesources to foster geotourism in communities: Case study of the Santa Elena peninsula geopark project in Ecuador. Sustainability 2020, 12, 4484. [Google Scholar] [CrossRef]
  27. Brilha, J. Geoheritage and Geoparks. In Geoheritage: Assessment, Protection, and Management; Reynard, E., José, B., Eds.; Elsevier: Amsterdam, The Netherlands, 2018; pp. 323–335. [Google Scholar]
  28. González Amuchastegui, M.J.; Serrano Cañadas, E.; González García, M. Lugares de interés geomorfológico, geopatrimonio y gestión de espacios naturales protegidos: El Parque Natural de Valderejo (Álava, España). Rev. Geogr. Norte Gd. 2014, 59, 45–64. [Google Scholar] [CrossRef] [Green Version]
  29. Feuillet, T.; Sourp, E. Geomorphological Heritage of the Pyrenees National Park (France): Assessment, Clustering, and Promotion of Geomorphosites. Geoheritage 2011, 3, 151–162. [Google Scholar] [CrossRef]
  30. Coratza, P.; Hobléa, F. The Specificities of Geomorphological Heritage. In Geoheritage: Assessment, Protection, and Management; Reynard, E., Brilha, J., Eds.; Elsevier: Amsterdam, The Netherlands, 2018; pp. 87–106. [Google Scholar]
  31. Palacio-Prieto, J.L. Geoheritage Within Cities: Urban Geosites in Mexico City. Geoheritage 2014, 7, 365–373. [Google Scholar] [CrossRef]
  32. Panizza, M. Geomorphosites: Concepts, methods and examples of geomorphological survey. Chin. Sci. Bull. 2001, 46, 4–5. [Google Scholar] [CrossRef]
  33. Reynard, E.; Brilha, J. Geomorphological Heritage and Geomorphosites: Definitions. In Geoheritage: Assessment, Protection, and Management; Reynard, E., Brilha, J., Eds.; Elsevier: Amsterdam, The Netherlands, 2017; pp. 88–94. [Google Scholar]
  34. Serrano, E.; González-Trueba, J.J. Assessment of geomorphosites in natural protected areas: The Picos de Europa National Park (Spain). Géomorphol. Relief Process. Environ. 2005, 11, 197–208. [Google Scholar] [CrossRef] [Green Version]
  35. Albani, R.A.; Mansur, K.L.; de Souza Carvalho, I.; dos Santos, W.F.S. Quantitative evaluation of the geosites and geodiversity sites of João Dourado Municipality (Bahia—Brazil). Geoheritage 2020, 12, 1–15. [Google Scholar] [CrossRef]
  36. Reynard, E.; Paniza, M. Geomorphosites: Definition, assessment and mapping. An introduction. Géomorphol. Relief Process. Environ. 2005, 11, 177–180. [Google Scholar] [CrossRef] [Green Version]
  37. Palacio Prieto, J. Geosites, geomorphosites and geoparks: Importance, actual situation and perspectives in Mexico. Investig. Geogr. 2013, 82, 24–37. [Google Scholar]
  38. Prosser, C.D.; Díaz-Martínez, E.; Larwood, J.G. The Conservation of Geosites. In Geoheritage: Assessment, Protection, and Management; Reynard, E., Brilha, J., Eds.; Elsevier: Amsterdam, The Netherlands, 2018; pp. 193–212. [Google Scholar]
  39. Henriques, M.H.; Pena dos Reis, R.; Brilha, J.; Mota, T.S. Geoconservation as an emerging geoscience. Geoheritage 2011, 3, 17–128. [Google Scholar] [CrossRef] [Green Version]
  40. Pereira, R. Geoconservação e Desenvolvimento Sustentável na Chapada Diamantina (Bahia-Brasil). Ph.D. Thesis, Minho University, Braga, Portugal, 2010. [Google Scholar]
  41. Brocx, M.; Semeniuk, V. Geoheritage and geoconservation—History, definition, scope and scale. J. R. Soc. West. Aust. 2007, 90, 53–87. [Google Scholar]
  42. Brilha, J. Geoconservation and protected areas. Environ. Conserv. 2002, 29, 273–276. [Google Scholar] [CrossRef] [Green Version]
  43. Herrera-Franco, G.; Montalván-Burbano, N.; Carrión-Mero, P.; Apolo-Masache, B.; Jaya-Montalvo, M. Research Trends in Geotourism: A Bibliometric Analysis Using the Scopus Database. Geosciences 2020, 10, 379. [Google Scholar] [CrossRef]
  44. Newsome, D.; Dowling, R. Geoheritage and Geotourism. In Geoheritage: Assessment, Protection, and Management; Reynard, E., Brilha, J., Eds.; Elsevier: Amsterdam, The Netherlands, 2018; pp. 305–321. [Google Scholar]
  45. Lonsdale, P. Ecuadorian subduction system. Am. Assoc. Pet. Geol. Bull. 1978, 62, 2454–2477. [Google Scholar]
  46. Carrión-Mero, P.C.; Morante-Carballo, F.E.; Herrera-Franco, G.A.; Maldonado-Zamora, A.; Paz-Salas, N. The context of Ecuador’s world heritage, for sustainable development strategies. Int. J. Des. Nat. Ecodyn. 2020, 15, 39–46. [Google Scholar] [CrossRef]
  47. Castillo Benalcázar, C.E. Análisis Situacional de “La Ruta Escondida” Quito-San José de Minas en la Provincia de Pichincha y Elaboración de una Guía Turística. Bachelor’s Thesis, Universidad Central del Ecuador, Quito, Ecuador, 2015. [Google Scholar]
  48. Villacís, M.; Cantos, A.; Pons, G.; Ludeña, V. Cultural Eco-Route Mojanda-Cochasqui: A sustainable tourism development proposal for the rural area of Pichincha, Ecuador. Turismo y Sociedad 2016, 19, 121–135. [Google Scholar]
  49. Gaibor Chiriboga, C.V. Diseño de una Ruta Agroecoturística en las Parroquias Nor-Centrales de la Provincia de Pichincha. Bachelor’s Thesis, Universidad Central del Ecuador, Quito, Ecuador, 2018. [Google Scholar]
  50. Ayala-Granda, A.; Carrión-Mero, P.; Paz-Salas, N.; Herrera-Franco, G.; Morante-Carballo, F.; Gurumendi-Noriega, M. Registro y valoración de geomorfositios de la zona sur de la Ruta Escondida, como alternativa de fomento a la geoconservación del paisaje en la región Caranqui-Ecuador. In Proceedings of the 18th LACCEI International Multi-Conference for Engineering, Education, and Technology, Virtual Edition, Buenos Aires, Argentina, 29–31 July 2020; pp. 1–10. [Google Scholar]
  51. Monzier, M.; Hall, M.; Cotten, J.; Robin, C.; Mothes, P.; Eissen, J.; Samaniego, P. Les adakites d’Equateur: Modèle préliminaire. C. R. Acad. Sci. Paris 1997, 324, 545–552. [Google Scholar]
  52. Bourdon, E.; Eissen, J.-P.; Gutscher, M.-A.; Monzier, M.; Hall, M.L.; Cotten, J. Magmatic response to early aseismic ridge subduction: The Ecuadorian margin case (South America). Earth Planet. Sci. Lett. 2003, 205, 123–138. [Google Scholar] [CrossRef]
  53. Robin, C.; Eissen, J.-P.; Samaniego, P.; Martin, H.; Hall, M.; Cotten, J. Evolution of the late Pleistocene Mojanda–Fuya Fuya volcanic complex (Ecuador), by progressive adakitic involvement in mantle magma sources. Bull. Volcanol. 2009, 71, 233–258. [Google Scholar] [CrossRef] [Green Version]
  54. Barberi, F.; Coltelli, M.; Ferrara, G. Plio–Quaternary volcanism in Ecuador. Geol. Mag. 1988, 125, 1–14. [Google Scholar] [CrossRef]
  55. Bigazzi, G.; Colteli, M.; Haadler Neto, J.; Osorio, A.; Oddone, M.; Salazar, E. Obsidian-bearing flows and pre-Colombian artifacts from the Ecuadorian Andes: First new multidisciplinary data. J. S. Am. Earth Sci. 1992, 6, 21–32. [Google Scholar] [CrossRef]
  56. Hall, M.; Mothes, P. El Origen y Edad de la Cangahua Superior, Valle de Tumbaco, Ecuador, Quito, Ecuador. In Suelos Volcánicos Endurecidos (Quito, Diciembre 1996); Zebrowski, C., Quantin, P., Trujillo, G., Eds.; ORSTOM: Paris, France, 1997; pp. 19–28. [Google Scholar]
  57. Clapperton, C. Quaternary Geology and Geomorphology of South America; Elsevier: Amsterdam, The Netherlands, 1993. [Google Scholar]
  58. Robin, C.; Hall, M.; Jimenez, M.; Monzier, M.; Escobar, P. Mojanda volcanic complex (Ecuador): Development of two adjacent contemporaneous volcanoes with contrasting eruptive styles and magmatic suites. J. S. Am. Earth Sci. 1997, 10, 345–359. [Google Scholar] [CrossRef]
  59. Clapperton, C.M. Glacial geomorphology Quatemary glacial sequence and paleoclimatic inferences in the Ecuatorian Andes. Int. Geomorphol. 1987, 843–870. [Google Scholar]
  60. Schubert, C.; Clapperton, C.M. Quaternary Glaciations in the Northern Andes (Venezuela, Colombia and Ecuador). Quat. Sci. Rev. 1990, 9, 123–135. [Google Scholar] [CrossRef]
  61. Toro, G.E.; Hermelin, M. Estudio comparativo de los paleoclimas en Colombia, Ecuador y Venezuela. In Climas Cuaternarios en Améria del Sur; Argollo, J., Mourguiart, P., Eds.; ORSTOM: La Paz, Bolivia, 1995; p. 46. [Google Scholar]
  62. Cáceres, B. Actualización del Inventario de Tres Casquetes Glaciares del Ecuador; Informe de Pasantía de Investigación; Université Nice Sophia Antipolis: Nice, France, 2010; pp. 10–11. [Google Scholar]
  63. Francou, B.; Rabatel, A.; Soruco, A.; Sicart, J.E.; Silvestre, E.; Ginot, P.; Cáceres, B.; Condom, T.; Villacís, M.; Ceballos, J.L.; et al. El retroceso de los glaciares andinos durante los últimos siglos y la aceleración del proceso desde el año 1976. In Glaciares de los Andes Tropicales: Víctimas del Cambio Climático; Silvestre, E., Francou, B., Villacís, M., Eds.; Comunidad Andina: Lima, Peru, 2013; pp. 27–41. [Google Scholar]
  64. Jomelli, V.; Favier, V.; Rabalet, A.; Brunstein, D.; Hoffmann, D.; Francou, B. Fluctuations of glaciers in the tropical Andes over the last millennium and palaeoclimatic implications: A review. Palaeogeogr. Palaeoclimatol. Palaeoecol. 2009, 281, 269–282. [Google Scholar] [CrossRef]
  65. Ayala, A. Aspectos Geológicos-Geomorfológicos y su Implicación en la Ocupación Pre-Hispánica del Período de Integración (500 d.c.–1535 d.c.), en la Ruta Escondida de la Región Caranqui. Master’s Thesis, Escuela Superior Politécnica del Litoral, Guayaquil, Ecuador, 2018. [Google Scholar]
  66. Serrano, S. Etnoarqueología de Intercambio de Bienes y Productos en los Caminos Precolombinos de Pichincha y Napo, Ecuador. Bachelor’s Thesis, Escuela Superior Politécnica del Litoral, Guayaquil, Ecuador, 2017. [Google Scholar]
  67. Dyson, R.G. Strategic development and SWOT analysis at the University of Warwick. Eur. J. Oper. Res. 2004, 152, 631–640. [Google Scholar] [CrossRef]
  68. Štrba, L.; Rybár, P.; Baláž, B.; Moloká, M. Current Issues in Tourism Geosite assessments: Comparison of methods and results. Curr. Issues Tour. 2015, 18, 496–510. [Google Scholar] [CrossRef]
  69. Štrba, L.; Kršák, B.; Sidor, C. Some Comments to Geosite Assessment, Visitors, and Geotourism Sustainability. Sustainability 2018, 10, 2589. [Google Scholar] [CrossRef] [Green Version]
  70. Pereira, P.; Pereira, D.; Caetano Alves, M.I. Geomorphosite assessment in Montesinho Natural Park (Portugal). Geogr. Helv. 2007, 62, 159–168. [Google Scholar] [CrossRef] [Green Version]
  71. Kubalíková, L.; Kirchner, K. Geosite and Geomorphosite Assessment as a Tool for Geoconservation and Geotourism Purposes: A Case Study from Vizovická vrchovina Highland (Eastern Part of the Czech Republic). Geoheritage 2016, 8, 5–14. [Google Scholar] [CrossRef]
  72. Pellitero, R.; González-Amuchastegui, M.J.; Ruiz-Flaño, P.; Serrano, E. Geodiversity and geomorphosite assessment applied to a natural protected area: The Ebro and Rudron Gorges Natural Park (Spain). Geoheritage 2010, 3, 163–174. [Google Scholar] [CrossRef]
  73. Kubalíková, L. Geomorphosite assessment for geotourism purposes. Czech J. Tour. 2013, 2, 80–104. [Google Scholar] [CrossRef]
  74. Reynard, E.; Perret, A.; Bussard, J.; Grangier, L.; Martin, S. Integrated Approach for the Inventory and Management of Geomorphological Heritage at the Regional Scale. Geoheritage 2016, 8, 43–60. [Google Scholar] [CrossRef]
  75. Pralong, J.P. A method for assessing tourist potential and use of geomorphological sites. Géomorphol. Relief Process. Environ. 2005, 11, 189–196. [Google Scholar] [CrossRef]
  76. Asociación de Servicios de Geología y Minería de Iberoamérica (ASGMI). Bases para el Desarrollo Común del Patrimonio Geológico en los Servicios Geológicos de Iberoamérica; ASGMI: Salta, Argentina, 2018. [Google Scholar]
  77. Berrezueta, E.; Domínguez Cuesta, M.J.; Carrión, P.; Berrezueta, T.; Herrero, G. Propuesta metodológica para el aprovechamiento del patrimonio geológico minero de la zona Zaruma-Portovelo (Ecuador). Trab. Geol. 2006, 26, 103–109. [Google Scholar]
  78. Carrión-Mero, P.; Loor-Oporto, O.; Andrade-Ríos, H.; Herrera-Franco, G.; Morante-Carballo, F.; Jaya-Montalvo, M.; Aguilar-Aguilar, M.; Torres-Peña, K.; Berrezueta, E. Quantitative and Qualitative Assessment of the “El Sexmo” Tourist Gold Mine (Zaruma, Ecuador) as A Geosite and Mining Site. Resources 2020, 9, 28. [Google Scholar] [CrossRef] [Green Version]
  79. Herrera-Franco, G.A.; Carrión-Mero, P.C.; Mora-Frank, C.V.; Caicedo-Potosí, J.K. Comparative analysis of methodologies for the evaluation of geosites in the context of the Santa Elena-Ancón geopark project. Int. J. Des. Nat. Ecodyn. 2020, 15, 183–188. [Google Scholar] [CrossRef]
  80. Herrera Franco, G.; Carrion Mero, P.; Morante Carballo, F.; Herrera Narváez, G.; Briones Bitar, J.; Blanco Torrens, R. Strategies for the development of the value of the mining-industrial heritage of the Zaruma-Portovelo, Ecuador, in the context of a geopark project. Int. J. Energy Prod. Manag. 2020, 5, 48–59. [Google Scholar]
  81. Morante-Carballo, F.; Herrera-Narváez, G.; Jiménez-Orellana, N.; Carrión-Mero, P. Puyango, Ecuador Petrified Forest, a Geological Heritage of the Cretaceous Albian-Middle, and Its Relevance for the Sustainable Development of Geotourism. Sustainability 2020, 12, 6579. [Google Scholar] [CrossRef]
  82. Herrera, G.; Carrión, P.; Briones, J. Geotourism potential in the context of the Geopark project for the development of Santa Elena province, Ecuador. WIT Trans. Ecol. Environ. 2018, 217, 557–568. [Google Scholar]
  83. Mehdioui, S.; El Hadi, H.; Tahiri, A.; Brilha, J.; El Haibi, H.; Tahiri, M. Inventory and Quantitative Assessment of Geosites in Rabat-Tiflet Region (North Western Morocco): Preliminary Study to Evaluate the Potential of the Area to Become a Geopark. Geoheritage 2020, 12, 1–17. [Google Scholar] [CrossRef]
  84. Niculiţă, M.; Mărgărint, M.C. Landslides and Fortified Settlements as Valuable Cultural Geomorphosites and Geoheritage Sites in the Moldavian Plateau, North-Eastern Romania. Geoheritage 2018, 10, 613–634. [Google Scholar] [CrossRef]
  85. Clivaz, M.; Reynard, E. How to Integrate Invisible Geomorphosites in an Inventory: A Case Study in the Rhone River Valley (Switzerland). Geoheritage 2018, 10, 527–541. [Google Scholar] [CrossRef]
  86. Chingombe, W. Preliminary geomorphosites assessment along the panorama route of mpumalanga province, South Africa. Geoj. Tour. Geosites 2019, 27, 1261–1270. [Google Scholar] [CrossRef]
  87. Zgłobicki, W.; Poesen, J.; Cohen, M.; Del Monte, M.; García-Ruiz, J.M.; Ionita, I.; Niacsu, L.; Machová, Z.; Martín-Duque, J.F.; Nadal-Romero, E.; et al. The Potential of Permanent Gullies in Europe as Geomorphosites. Geoheritage 2019, 11, 217–239. [Google Scholar] [CrossRef]
  88. Sinnyovsky, D.; Sachkov, D.; Tsvetkova, I.; Atanasova, N. Geomorphosite Characterization Method for the Purpose of an Aspiring Geopark Application Dossier on the Example of Maritsa Cirque Complex in Geopark Rila, Rila Mountain, SW Bulgaria. Geoheritage 2020, 12, 1–17. [Google Scholar] [CrossRef]
  89. Chrobak, A.; Witkowski, K.; Szmańda, J. Assessment of the Educational Values of Geomorphosites Based on the Expert Method, Case Study: The Białka and Skawa Rivers, the Polish Carpathians. Quaest. Geogr. 2020, 39, 45–57. [Google Scholar] [CrossRef]
  90. Vukoičić, D.; Ivanović, R.; Radovanović, D.; Dragojlović, J.; Martić-Bursać, N.; Ivanović, M.; Ristić, D. Assessment of geotourism values and ecological status of mines in Kopaonik mountain (Serbia). Minerals 2020, 10, 269. [Google Scholar] [CrossRef] [Green Version]
  91. Kubalíková, L. Assessing geotourism resources on a local level: A case study from Southern Moravia (Czech Republic). Resources 2019, 8, 150. [Google Scholar] [CrossRef] [Green Version]
  92. Kirchner, K.; Kubalíková, L. Relief assessment methodology with respect to geoheritage based on example of the Deblínská vrchovina Highland. In Proceedings of the Public Recreation and Landscape Protection—With Man Hand in Hand, Brno, Czech Republic, 1–3 May 2013; pp. 131–141. [Google Scholar]
  93. Nazaruddin, D.A. Systematic Studies of Geoheritage in Jeli District, Kelantan, Malaysia. Geoheritage 2017, 9, 19–33. [Google Scholar] [CrossRef]
  94. Torabi Farsani, N.; Coelho, C.; Costa, C. Geotourism and Geoparks as Gateways to Socio-cultural Sustainability in Qeshm Rural Areas, Iran. Asia Pacific J. Tour. Res. 2012, 17, 30–48. [Google Scholar] [CrossRef]
  95. Cai, Y.; Wu, F.; Han, J.; Chu, H. Geoheritage and Sustainable Development in Yimengshan Geopark. Geoheritage 2019, 11, 991–1003. [Google Scholar] [CrossRef]
Figure 1. Conceptual framework of geodiversity, geological heritage and geoconservation. Modified from [15].
Figure 1. Conceptual framework of geodiversity, geological heritage and geoconservation. Modified from [15].
Sustainability 12 08468 g001
Figure 2. Location of the study area (a) general context; (b) local context.
Figure 2. Location of the study area (a) general context; (b) local context.
Sustainability 12 08468 g002
Figure 3. Geological map of the Mojanda–Fuya Fuya volcanic complex. Based on [53].
Figure 3. Geological map of the Mojanda–Fuya Fuya volcanic complex. Based on [53].
Sustainability 12 08468 g003
Figure 4. The general methodology of the study.
Figure 4. The general methodology of the study.
Sustainability 12 08468 g004
Figure 5. Location of potential geomorphosites: (1) Cimas San José de Minas (peaks), (2) Relieves de Chavezpamba y San José de Minas (mountainous relief), (3) Relieve de San José de Minas (volcanic relief), (4) Piedemonte San José de Minas (foothills), (5) Valle Atahualpa (valley), (6) Coluviones de Atahualpa (colluviums), (7) Domo Atahualpa (dome), (8) Cima de Atahualpa (peak), (9) Terrazas de Atahualpa (hanging terraces), (10) Domo El Panecillo (dome), (11) Pendiente Fuya Fuya (spreading slope), (12) Flujos de Atahualpa (lahar flows), (13) Relieve de Atahualpa (volcanic relief), (14) Valle de San Bartolo–Fuya Fuya (glacial valley), (15) Montículos Fuya Fuya (moraines), (16) Laguna de Mojanda (lagoon), (17) Macizo Mojanda (rock massif) y (18) Flujos Mojanda (flows).
Figure 5. Location of potential geomorphosites: (1) Cimas San José de Minas (peaks), (2) Relieves de Chavezpamba y San José de Minas (mountainous relief), (3) Relieve de San José de Minas (volcanic relief), (4) Piedemonte San José de Minas (foothills), (5) Valle Atahualpa (valley), (6) Coluviones de Atahualpa (colluviums), (7) Domo Atahualpa (dome), (8) Cima de Atahualpa (peak), (9) Terrazas de Atahualpa (hanging terraces), (10) Domo El Panecillo (dome), (11) Pendiente Fuya Fuya (spreading slope), (12) Flujos de Atahualpa (lahar flows), (13) Relieve de Atahualpa (volcanic relief), (14) Valle de San Bartolo–Fuya Fuya (glacial valley), (15) Montículos Fuya Fuya (moraines), (16) Laguna de Mojanda (lagoon), (17) Macizo Mojanda (rock massif) y (18) Flujos Mojanda (flows).
Sustainability 12 08468 g005
Figure 6. Examples of geomorphosites in the study area: (a) Laguna de Mojanda, (b) Valle de San Bartolo–Fuya Fuya, (c) Cima de Atahualpa, (d) Piedemonte San José de Minas, (e) Cimas de San José de Minas, (f) Domo Atahualpa [65].
Figure 6. Examples of geomorphosites in the study area: (a) Laguna de Mojanda, (b) Valle de San Bartolo–Fuya Fuya, (c) Cima de Atahualpa, (d) Piedemonte San José de Minas, (e) Cimas de San José de Minas, (f) Domo Atahualpa [65].
Sustainability 12 08468 g006
Figure 7. Classification of the results obtained from the AI of the geomorphosites.
Figure 7. Classification of the results obtained from the AI of the geomorphosites.
Sustainability 12 08468 g007
Figure 8. Classification of the obtained results of susceptibility to degradation (SD) of the geomorphosites.
Figure 8. Classification of the obtained results of susceptibility to degradation (SD) of the geomorphosites.
Sustainability 12 08468 g008
Figure 9. Classification of the results obtained from the protection priority (PP) assessment of the geomorphosites.
Figure 9. Classification of the results obtained from the protection priority (PP) assessment of the geomorphosites.
Sustainability 12 08468 g009
Figure 10. Suggested itinerary “Rocks & Water in Ruta Escondida”, selecting several geomorphosites. Geomorphosites: (5) Valle Atahualpa (valley), (6) Coluviones antiguos de Atahualpa (colluvium), (7) Domo Atahualpa (volcanic dome), (9) Terrazas de Atahualpa (hanging terrace), (10) Domo El Panecillo (Volcanic dome), (11) Pendiente Fuya Fuya (volcanic spreading slope), (12) Flujos de Atahualpa (lahar flow), (13) Relieve de Atahualpa (volcanic relief), (14) Valle de San Bartolo–Fuya Fuya (glacial valley), (15) Montículos Fuya Fuya (glacial moraines), (16) Laguna de Mojanda (lagoon), (17) Macizo Mojanda (rock massif) y (18) Flujos Mojanda (Mojanda flows). Access to the itinerary: (A) access from Quito city, (B) access from Puéllaro parish, (C) access from Imbabura UNESCO (United Nations Educational, Scientific and Cultural Organization) Global Geopark, (D) access from Pedro Moncayo city.
Figure 10. Suggested itinerary “Rocks & Water in Ruta Escondida”, selecting several geomorphosites. Geomorphosites: (5) Valle Atahualpa (valley), (6) Coluviones antiguos de Atahualpa (colluvium), (7) Domo Atahualpa (volcanic dome), (9) Terrazas de Atahualpa (hanging terrace), (10) Domo El Panecillo (Volcanic dome), (11) Pendiente Fuya Fuya (volcanic spreading slope), (12) Flujos de Atahualpa (lahar flow), (13) Relieve de Atahualpa (volcanic relief), (14) Valle de San Bartolo–Fuya Fuya (glacial valley), (15) Montículos Fuya Fuya (glacial moraines), (16) Laguna de Mojanda (lagoon), (17) Macizo Mojanda (rock massif) y (18) Flujos Mojanda (Mojanda flows). Access to the itinerary: (A) access from Quito city, (B) access from Puéllaro parish, (C) access from Imbabura UNESCO (United Nations Educational, Scientific and Cultural Organization) Global Geopark, (D) access from Pedro Moncayo city.
Sustainability 12 08468 g010
Table 1. Criteria and indicators used for the quantitative assessment of geosites using the IELIG methodology (acronym in Spanish for “Inventario Español de Lugares de Interés Geológico”). Weight (constant values in %) for the value of interest: scientific (Sc), academic (Ac) and Tourist (To). Interpretation: maximum (400), very high (267–400), high (134–266), medium (50–134), low (<50) [14].
Table 1. Criteria and indicators used for the quantitative assessment of geosites using the IELIG methodology (acronym in Spanish for “Inventario Español de Lugares de Interés Geológico”). Weight (constant values in %) for the value of interest: scientific (Sc), academic (Ac) and Tourist (To). Interpretation: maximum (400), very high (267–400), high (134–266), medium (50–134), low (<50) [14].
Geosite Assessment Model (IELIG)
IndicatorsValuesWeight
Parameters ScAcTo
Representativeness0–43050
Standard or reference site1050
Knowledge of the site1500
State of conservation1050
Conditions of observation1055
Scarcity, rarity1550
Geological diversity10100
Educational values0200
Logistics infrastructure0155
Population density055
Possibilities for public outreach (accessibility)01510
Size of site0015
Association with other natural elements055
Beauty0520
Informative value0015
Possibility of recreational and leisure activities005
Proximity to other places of interest005
Socio-economic situation0010
Total100100100
Table 2. Susceptibility assessment of the geosites and mining sites. Fragility (Fr), vulnerability (Vul), degradation susceptibility (DS) [14].
Table 2. Susceptibility assessment of the geosites and mining sites. Fragility (Fr), vulnerability (Vul), degradation susceptibility (DS) [14].
Susceptibility
Parameter/CharacteristicsFr
ValueWeight
Geosite size0–440
Vulnerability to looting30
Natural hazards30
Total Value100
Parameter/CharacteristicsVul
ValueWeight
Proximity to infrastructures0–420
Mining exploitation interest15
Protected area designation15
Indirect protection15
Accessibility15
Ownership status10
Population density5
Proximity to recreational areas5
Total Value100
DS
DS: (Fr × Vul)/400(>400) Maximum
(400–200) Very high
(199–68) High
(67–13) Medium
(<13) Low
Table 3. Protection priority (Pp) assessment of geosites and mining sites, based on the priority of protection of the scientific value Pp (Sc), academic Pp (Ac) and tourist Pp (To) [14].
Table 3. Protection priority (Pp) assessment of geosites and mining sites, based on the priority of protection of the scientific value Pp (Sc), academic Pp (Ac) and tourist Pp (To) [14].
PROTECTION
Pp   ( Sc )   =   ( ISc ) 2 )   ×   DS   × ( 1 / 400 2 )   Ec. 1
Pp   ( Ac )   =   ( IAc ) 2   ×   DS   × ( 1 / 400 2 )   Ec. 2
Pp   ( To )   =   ( ITo ) 2   ×   SD   × ( 1 / 400 2 )   Ec. 3
Pp   = ( ISc   +   IAc   +   ITo 3 ) 2   ×   DS   × ( 1 / ( 400 ) 2 )
Ec. 4
(>400) Maximum
(400–113) Very high
(113–17) High
(16–1) Medium
(0) Low
Table 4. List of potential geomorphosites in the study area, typological classification and general characteristics.
Table 4. List of potential geomorphosites in the study area, typological classification and general characteristics.
Geomorphosite TypeMain Characteristics
1Cimas de San José de MinasSharp peaksThese geomorphosites have slopes between 70 and 100%, located in the Atahualpa and San José de Minas sectors.
2Relieves de Chavezpamba y San José de MinasMountainous relief Most of the area presents geoforms with sharp and rounded peaks, slopes between the range 12–25% in the sectors of Puéllaro and Perucho and 70–100% in Atahualpa and San José de Minas.
3Relieve de San José de MinasVolcanic reliefReliefs, product of the volcanism of the area, are found on the slopes of Mojanda and Fuya Fuya.
4Piedemonte San José de MinasFoothillsPlain located next to the volcanic Cushnirumi, consisting of colluvial alluvial material and deposits of the Fuya Fuya.
5Valle AtahualpaValleyValley of sand, gravel and blocks of variable composition, located in the Atahualpa and San José de Minas sectors.
6Coluviones de AtahualpaColluviumGeoforms of sand with blocks of one meter in diameter are found throughout the study area.
7Domo AtahualpaVolcanic domeSmall dome located in the Atahualpa forest, composed of lava with a high content of silica.
8Cima de AtahualpaPeaks Peaks in the Atahualpa sector with slopes between 60 and 80% and some similarity to those of San José de Minas.
9Terrazas de AtahualpaHanging terraceMorphological unit of tectonic origin, with slopes between 2 and 12%, constituted of andesitic tuffs, covered with deposits from the Fuya Fuya.
10Domo El PanecilloVolcanic domeDome constituted of andesitic and dacitic lava flows, on its southwest side evidence of the collapse can be seen.
11Pendiente Fuya FuyaVolcanic spreading slopeA steep slope of tectonic origin, through which avalanche flows descended from the Fuya Fuya volcano.
12Flujos de AtahualpaLahar flowPyroclasts from the Fuya Fuya volcano are found throughout the route, except in Perucho parish.
13Relieve de AtahualpaVolcanic relief Volcanic reliefs similar to those present in San José de Minas. In the lower part they present rounded shapes, as a product of erosion.
14Valle de San Bartolo–Fuya FuyaGlacial valley U-shaped valley consists of Mojanda glacial deposits. The valley is constituted of the San Bartolo cone and the Fuya Fuya domes.
15Montículos Fuya FuyaGlacial morainesSmall rounded hill, constituted of heterogeneous and sub-rounded material, as a product of the advance of the ice, which is located next to the Laguna Mojanda.
16Laguna de MojandaLagoonThe lagoon is in the caldera of the Mojanda volcano, on top of basaltic andesites and sediments interspersed with volcanic ash.
17Macizo MojandaRock massifRocky massif constituted of andesitic lavas from the Mojanda volcanic complex.
18Flujos MojandaFlowsFlows situated under the rocky massifs, in areas with intense erosion.
Table 5. Scores of the scientific (Sc), didactic (Ac) and tourist (To) interests of the geomorphosites.
Table 5. Scores of the scientific (Sc), didactic (Ac) and tourist (To) interests of the geomorphosites.
GeomorphositeInterestsAverage Interest (AI)
ScAcTo
1Cimas de San José de Minas145115135132
2Relieves de Chavezpamba y San José de Minas10095135110
3Relieve de San José de Minas110115160128
4Piedemonte San José de Minas75110150112
5Valle Atahualpa240160125175
6Coluviones de Atahualpa1206510597
7Domo Atahualpa135105170137
8Cima de Atahualpa130130145135
9Terrazas de Atahualpa9080140103
10Domo El Panecillo240175190267
11Pendiente Fuya Fuya120115160132
12Flujos de Atahualpa11080105130
13Relieve de Atahualpa115110112113
14Valle de San Bartolo—Fuya Fuya130110150130
15Montículos Fuya Fuya10585120103
16Laguna de Mojanda325250225267
17Macizo Mojanda225145195188
18Flujos Mojanda175160190175
Table 6. Assessment results of the susceptibility to degradation (SD), fragility (F) and vulnerability (Vul.) due to anthropic threats.
Table 6. Assessment results of the susceptibility to degradation (SD), fragility (F) and vulnerability (Vul.) due to anthropic threats.
GeomorphositeSusceptibility
FVulSD
1Cimas de San José de Minas0250
2Relieves de Chavezpamba y San José de Minas30806
3Relieve de San José de Minas11010027.50
4Piedemonte San José de Minas10014536.25
5Valle Atahualpa7014024.50
6Coluviones de Atahualpa10010025
7Domo Atahualpa0250
8Cima de Atahualpa0250
9Terrazas de Atahualpa0700
10Domo El Panecillo01150
11Pendiente Fuya Fuya0800
12Flujos de Atahualpa8016032
13Relieve de Atahualpa40808
14Valle de San Bartolo—Fuya Fuya0550
15Montículos Fuya Fuya80255
16Laguna de Mojanda0700
17Macizo Mojanda0550
18Flujos Mojanda0850
Table 7. Results of the protection priority (PP) assessment.
Table 7. Results of the protection priority (PP) assessment.
GeomorphositeProtection Priority
PP (Sc)PP (Ac)PP (To)PP
1Cimas de San José de Minas0000
2Relieves de Chavezpamba y San José de Minas0.380.340.680.45
3Relieve de San José de Minas2.082.274.402.83
4Piedemonte San José de Minas1.272.745.12.83
5Valle Atahualpa8.823.922.394.69
6Coluviones de Atahualpa2.250.661.721.46
7Domo Atahualpa0000
8Cima de Atahualpa0000
9Terrazas de Atahualpa0000
10Domo El Panecillo0000
11Pendiente Fuya Fuya0000
12Flujos de Atahualpa2.421.282.211.93
13Relieve de Atahualpa0.660.610.660.64
14Valle de San Bartolo–Fuya Fuya0000
15Montículos Fuya Fuya0.340.230.450.33
16Laguna de Mojanda0000
17Macizo Mojanda0000
18Flujos Mojanda0000
Table 8. Strengths, weaknesses, opportunities, and threats (SWOT) matrix analysis of geomorphosites in the Ruta Escondida. The SWOT combining internal environment (strengths and weaknesses) identified by numbers 1 to 7 and the external environment (opportunities and threats) identified by letters (a) to (e).
Table 8. Strengths, weaknesses, opportunities, and threats (SWOT) matrix analysis of geomorphosites in the Ruta Escondida. The SWOT combining internal environment (strengths and weaknesses) identified by numbers 1 to 7 and the external environment (opportunities and threats) identified by letters (a) to (e).
StrengthsWeaknesses
Internal
Environment
  • Large number of geomorphosites.
  • Existence of scientific research on the Ruta Escondida (geological, archaeological and paleoecological).
  • Presence of natural and cultural landscape.
  • Variety of climatic levels.
  • Some erosive processes (natural erosion) reveal geological structures for observation, increasing educational and research interest.
  • Proximity to populated areas and the capital of the country.
  • Second-order road infrastructure in good condition.
  • Geomorphosite assessment scarcity.
  • Geoforms exposed to erosion (anthropic).
  • Lack of knowledge of natural potential.
  • Low development of geotourism in the country.
  • Lack of geomorphosite conservation.
  • Third-order roads with low maintenance.
External
Environment
OpportunitiesStrategies: Strengths + OpportunitiesStrategies: Weaknesses + Opportunities
  • Geomorphosite promotion to increase the local economy.
  • Conservation of culture and traditions.
  • Creation of training programs in the area.
  • Interest of the provincial, municipal and parochial government to enhance natural and cultural resources.
  • Academic research interest in the area.
1-a. Promote geomorphosites that illustrate the importance of the natural environment of the region.
2-e. Strengthen links between rural communities and academia for future research in different scientific areas.
3-d. Promote natural resources by each parish government to publicize geological resources.
5-c. Build awareness in the community about geomorphosites and their importance in the geotourism sector.
6-d. Access to places in good condition, which favors geotourism development.
1-a. Assess geological elements in the scientific, educational and tourist fields for their promotion to the community.
2-c. Train local community about protection measures that geomorphosites need due to anthropic activity.
3-c. Provide training on the preservation and importance of the resources that nature offers.
4-c. Promote geotourism development nationwide through documentaries.
6-d. Monitor access roads to geomorphosites to ensure geotourism.
ThreatsStrategies: Strengths-ThreatsStrategies: Weaknesses-Threats
  • Global economic crisis and health emergency, which will prevent the allocation of resources for further research and tourism promotion.
  • Aggressive climate changes.
  • Low financing in projects of scientific interest.
2-c. Scientific development of research institutions through funding from public and private organizations.
3-b. Implementation of protection measures for geological and cultural heritage.
1-c. Implementation of programs that cultivate funds aimed at the assessment of the geological elements.
3-a. Documentaries about geological resources.
Table 9. List of touristic places with natural and cultural wealth.
Table 9. List of touristic places with natural and cultural wealth.
SiteTypeMain Characteristics
1Río CubíNatural wealthWaterfalls, rich vegetation and several species of birds.
2Museo de Perucho: Arqueología e HistoriaCultural wealthMuseum of the Ruta Escondida pre-Columbian history.
3Iglesia de PeruchoCultural wealthWooden structure, built in 1700, with a sober style and austerity touch.
4Cerro ItaguaNatural wealthNatural viewpoint to observe the Chacezpamba parish, the Fuya Fuya hill and the Mojanda forest.
5Iglesia Parroquial de ChavezpambaCultural wealthBrick, clay and wood structure, built in 1950.
6Aguas termales de AtahualpaNatural wealthThermal water spa, rivers and vegetation.
7Camposanto de AtahualpaCultural wealthSite with cypress trees molded to different shapes, such as bears and birds.
8Iglesia Parroquial de AtahualpaCultural wealthChurch built in 1923 with community collaboration. Contains a legendary rock with the image of the Virgen del Quinche.
9Bosque protector MojandaNatural wealthA forest of approximately 1200 hectares, with several species of native Andean flora, of which two areas have been declared protective forest.
Table 10. Interest, Fr., Vul., DS, and Pp assessment in the context of the proposed route (Figure 10).
Table 10. Interest, Fr., Vul., DS, and Pp assessment in the context of the proposed route (Figure 10).
ItineraryInterestSusceptibilityProtection Priority
ScAcToAIVulFrDSScAcToPp
“Rocks & Water in Ruta Escondida”163.8113.8152.8143.5
(High)
81.528.55.80
(Low)
1.10.50.60.7
(Low)
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Carrión-Mero, P.; Ayala-Granda, A.; Serrano-Ayala, S.; Morante-Carballo, F.; Aguilar-Aguilar, M.; Gurumendi-Noriega, M.; Paz-Salas, N.; Herrera-Franco, G.; Berrezueta, E. Assessment of Geomorphosites for Geotourism in the Northern Part of the “Ruta Escondida” (Quito, Ecuador). Sustainability 2020, 12, 8468. https://doi.org/10.3390/su12208468

AMA Style

Carrión-Mero P, Ayala-Granda A, Serrano-Ayala S, Morante-Carballo F, Aguilar-Aguilar M, Gurumendi-Noriega M, Paz-Salas N, Herrera-Franco G, Berrezueta E. Assessment of Geomorphosites for Geotourism in the Northern Part of the “Ruta Escondida” (Quito, Ecuador). Sustainability. 2020; 12(20):8468. https://doi.org/10.3390/su12208468

Chicago/Turabian Style

Carrión-Mero, Paúl, Alicia Ayala-Granda, Sthefano Serrano-Ayala, Fernando Morante-Carballo, Maribel Aguilar-Aguilar, Miguel Gurumendi-Noriega, Nataly Paz-Salas, Gricelda Herrera-Franco, and Edgar Berrezueta. 2020. "Assessment of Geomorphosites for Geotourism in the Northern Part of the “Ruta Escondida” (Quito, Ecuador)" Sustainability 12, no. 20: 8468. https://doi.org/10.3390/su12208468

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop