Next Article in Journal
Investigating Company’s Technical Development Directions Based on Internal Knowledge Inheritance and Inventor Capabilities: The Case of Samsung Electronics
Next Article in Special Issue
Geoheritage Management in Areas with Multicultural Interest Contexts
Previous Article in Journal
Application of a Novel Optimized Fractional Grey Holt-Winters Model in Energy Forecasting
Previous Article in Special Issue
Evaluation of the Visitor Understanding of Coastal Geotourism and Geoheritage Potential Based on Sustainable Regional Development in Western Black Sea Region, Turkey
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Assessment of Geosites within a Natural Protected Area: A Case Study of Cajas National Park

by
Edison Navarrete
1,
Fernando Morante-Carballo
1,2,3,
Jairo Dueñas-Tovar
1,4,
Paúl Carrión-Mero
1,4,5,*,
María Jaya-Montalvo
1,4,* and
Edgar Berrezueta
6
1
Centro de Investigación y Proyectos Aplicados a las Ciencias de la Tierra (CIPAT), Campus Gustavo Galindo, ESPOL Polytechnic University, Km. 30.5 Vía Perimetral, Guayaquil P.O. Box 09-01-5863, Ecuador
2
Facultad de Ciencias Naturales y Matemáticas (FCNM), Campus Gustavo Galindo, ESPOL Polytechnic University, Km. 30.5 Vía Perimetral, Guayaquil P.O. Box 09-01-5863, Ecuador
3
Geo-Recursos y Aplicaciones GIGA, Campus Gustavo Galindo, ESPOL Polytechnic University, Km. 30.5 Vía Perimetral, Guayaquil P.O. Box 09-01-5863, Ecuador
4
Facultad de Ingeniería Ciencias de la Tierra (FICT), Campus Gustavo Galindo, ESPOL Polytechnic University, Km 30.5 Vía Perimetral, Guayaquil P.O. Box 09-01-5863, Ecuador
5
Instituto de Investigación Geológico y Energético (IIGE), Av. de la República E7-263 y Diego de Almagro, Quito 170518, Ecuador
6
Departamento de Recursos para la Transición Ecológica, Instituto Geológico y Minero de España (IGME, CSIC), 33005 Oviedo, Spain
*
Authors to whom correspondence should be addressed.
Sustainability 2022, 14(5), 3120; https://doi.org/10.3390/su14053120
Submission received: 29 January 2022 / Revised: 27 February 2022 / Accepted: 28 February 2022 / Published: 7 March 2022

Abstract

:
Cajas National Park (CNP), located in southern Ecuador, comprises an area of high natural, scientific and cultural value with wide recognition worldwide. This national park has a large number of elements that, as a whole, constitute a relevant geological heritage. However, this geological heritage requires an enhancement that complements the important contribution made by the Natural Park in terms of conservation and protection of the natural heritage. This study aims to evaluate sites of geological relevance present in CNP through international geosite assessment methodologies and thus provide knowledge favouring these resources’ sustainable use and geoconservation. The study phases comprise four stages: (i) a base information analysis of the study area; (ii) identification and selection of sites of geological interest; (iii) a geosite and geomorphosite assessment using the Inventario Español de Lugares de Interés Geological (IELIG) method and Brilha method; (iv) a qualitative assessment using a strengths, weaknesses, opportunities and threats (SWOT) analysis for the contribution and influence of geomorphosites in the development of the study area. This work made it possible to determine that all the analysed geosites and geomorphological sites (14) have a high and very high interest. The Llaviucu valley site stands out for its relevant scientific, academic and tourist value. The IELIG method revealed that 50% of the evaluated sites have a high protection priority, while the rest are in the “medium” category. In addition, the investigation through the SWOT analysis revealed that the geomorphosites could provide significant added value to the development of geotourism and of the NP itself, complementing the already known natural attractions; moreover, the study presented strategies for the use of these in the sustainable development of the area.

1. Introduction

“Geodiversity” is a key term used to understand the role of geological elements within the natural diversity of the territory [1]. More specifically, geoheritage refers to the occurrence of in situ or ex situ elements with a strong geoscientific content of said natural diversity (e.g., rocks, minerals, fossils, landforms and geological collections in museums) [2,3,4]. However, this term also includes the cultural, social, economic and environmental values presented by the various geological elements of a territory [5], which are important for understanding the geological history of the sector [6].
The conservation of the geological heritage of a place that includes its management, protection, promotion and periodic monitoring of its preservation to ensure that these sites remain accessible as reference sites for the future, is known as “geoconservation” [4,7]. It is important to note that many official programs (e.g., Global Geoparks from UNESCO, the Geoheritage Specialist Group of the International Union for Conservation of Nature and geoheritage–federal programs from the USA) aim to preserve the sites that are most valuable in terms of their geodiversity, especially if they are classified as geoheritage. Additionally, the geodiversity’s preservation and conservation can also be under other protection figures such as natural or national parks [8,9,10]. However, some natural parks have very relevant geological features that have led to their recognition as geoparks (e.g., Fforest Fawr Geopark [11] and Cilento Vallo di Diano Geopark [12]).
Geoparks are born as a model for promoting natural heritage (both biotic and abiotic) and cultural heritage (tangible and intangible) in an environment of sustainable development with benefits for local communities [13,14].
When a certain area is recognised by the predominance of geological elements of high scientific interest, portions of that territory can recreate the earth’s geological evolution [15]. This singularity also includes an example of geodiversity; these geological or geomorphological objects are called “geosites” [16].
Depending on the geological elements’ representation level, geosites can be of various types (e.g., stratigraphical, geomorphological, paleontological and mineralogical) [17]. More specifically, Panizza [18] suggests the term “geomorphosite” to define those sites with high geomorphological content, which present scenic, socio-economic, cultural and scientific values, which can be useful to society.
Unlike other sites (e.g., paleontological sites), geomorphosites become unique geoheritage sites because they have aesthetic dimension, dynamics dimension and imbrication scales [19]. In addition, they constitute an essential tool to promote geocultural diffusion [20] and geological education [21].
Protected areas host glacial features with outstanding universal value as geoheritage sites [22]. The sectors with geoheritage elements of glacial origin are not limited to a specific region; several examples of protected areas that present geoheritage elements of glacial origin that exist worldwide are Chirripó National Park (Costa Rica) [22], Mt. Kumgang (DPR Korea) [23] and the Ruiz–Tolima Volcano and Glacier Complex (Colombia) [24]. Within the UNESCO Global Geoparks (UGGp) Network, there are numerous examples of geoparks with glacier representation; among them are Chablais (France) [25], Estrela (Portugal) [26] and the Courel Mountains (Spain) [27].
The protection and promotion of geoheritage has led to the generation of various assessment methodologies with scientific, educational and recreational use and management approaches [28,29,30]. The assessment dimensions vary by method. In the literature, there are a large number of quantitative and qualitative methodologies for geological heritage inventories and site evaluations; among the most prominent are: Reynard et al. [31], Brilha [32], Pereira et al. [33], de Lima et al. [34], Kubalíková [35], Pralong [36,37], Cendrero and Bruschi [38], Coratza and Giusti [39], García-Cortés and Carcavilla [40], Vujičić et al. [41] and Comanescu and Nedelea [42]. The methodologies’ evaluation criteria help quantify the state of the sites of geological interest evaluated in each case study, allowing the determination of strategies for the sustainable use of the geological heritage.
The inclusion of communities in the disclosure of space with high geoscientific content has allowed the development of a new concept within studies related to these spaces [43]. This new term is “geotourism”, which is a concept that has a degree of uncertainty in the scientific literature [44]. One of the first accepted definitions was the one proposed by Hose [45], stating that its meaning includes the provision of facilities to promote geological knowledge beyond an aesthetic appreciation.
Ecuador is a megadiverse country [46] with an international level of natural and cultural diversity [47], having being recognised as one of the countries with the greatest biodiversity worldwide [48]. In contrast, Sánchez-Cortez [49] explains that the Ecuadorian state has not presented specific criteria for developing strategies focused on protecting geoheritage. In recent years, these have been promoted by the academic and private sectors, significantly impacting tourism development in local communities.
In Ecuador, the Cajas National Park (CNP) is part of the national system of protected areas of the Ministry of the Environment of Ecuador, which groups 56 protected areas and national parks in the country [50]. At the same time, CNP presents two UNESCO designations: world biosphere reserve (since 2003) [51] and site protected by the Ramsar Convention (since 2002) [52]. Within its limits, there is also part of the oldest network of roads in the American continent, called Qhapaq Ñan or Camino del Inca, which was also recognised by UNESCO as a World Heritage Site in 2014. In addition, the park has a great number of elements of glacial origin, which constitute a heritage of regional and international reference within glacial geomorphology [53].
As for the UNESCO geopark protection figure, in 2019 this denomination was granted to the Imbabura geopark initiative. This recognition opened a new trend in developing geological heritage in the country. Currently, Ecuador has two aspiring geoparks: Napo-Sumaco and Volcán Tungurahua, as well as another twelve in progress [54].
The Ecuadorian state has recognised CNP as a national park, and to favour and complement its sustainable development, the enhancement of its geological elements is appropriate. Therefore, the main aim of this research is to evaluate CNP geosites and geomorphosites using two international assessment methodologies, to highlight the geological qualities of CNP as a complement to the current figure of protection and sustainable development of the area.
To fulfil this aim, this research has three sections. The first section provides a vision of the geographical, geological and geomorphological environment of CNP, especially reviewing the scientific bases and concepts related to glacier geomorphology. Section 2 presents the methodological approach of the study and details the selection of the 14 representative geosites–geomorphosites and their evaluation by the semi-quantitative methods of García-Cortés [40] and Brilha [32], including technical details of its particularities. In addition, a qualitative analysis was carried out to determine geoconservation strategies. Section 3 shows the results obtained from the study, while Section 4 discusses the findings and comparisons with other glacial geological sites. Finally, Section 5 summarises the conclusions and future implications of the research.

1.1. Study Area

Cajas National Park (CNP) is in the southern part of mainland Ecuador, within the Province of Azuay (UTM WGS 1984, 17 South Zone: 678000–707000 E/9674000–9696000 N) (Figure 1a), with an area of approximately 149 km2, 140 km from the city of Guayaquil and 21 km from the centre of the city of Cuenca, the second and third largest city in terms of demographic, commercial and cultural importance in the country. The administrative framework is within the Molleturo, Sayausí, Chaucha and San Joaquín parishes (Figure 1b). CNP owes its name to the Quichua word “cassa”, which means “door or entrance to the Sierra Nevada” [55], referring to the intra-mountain valleys that facilitated transport and trade between the Coastal and Andean regions since pre-Columbian times.
Within the immediate area of influence of the park, there are nine localities with a markedly low population density, distributed among the four parishes that share space with CNP. In 2011, the Municipality of Cuenca recorded the following figures for the total population of said communities: Miguir (192), Río Blanco (306), Patul (55), Llano Largo (57), Zhin Alto (87), Angas (12), San Antonio de Chaucha (226), Soldiers (177) and Baute (4). This low population density is the product of the population’s motivation to take advantage of the area for productive activities such as agriculture or livestock [56,57].
The biodiversity of CNP has species of both plant and animal origin. Within the flora species, it is contemplated that there are 500 vascular plants, corresponding to half of the genera of this type located in the paramos worldwide, of which 71 are endemic species of Ecuador and 16 are under extinction threats [57,58]. The most dominant vegetation in the area corresponds to the grassland of the páramo. The fauna of the park has the following numbers of species: birds (152), mammals (43), amphibians (15) and reptiles (4) [58].
The climate is characteristic of the high Andean system, strongly influenced by the Andes Mountain range, generating many microclimates depending on the location. The warm waters of the El Niño current cause a marked presence of clouds, unlike other Andean areas in South America [59]. Between 1965–1990, the average precipitation was 1072 mm, with maximum values close to 1400 mm and minimum values of 829 mm [58]. The humidity of the area and other variables, such as the large presence of organic matter and materials of volcanic origin, allow effective water retention in its soils [60]. The temperature fluctuates between −2 °C and 18 °C, presenting moderate to high rainfall without a spatial variability defined by the influence of direction winds and relief [61]. Solar radiation presents a strong daily cycle due to its location in a tropical zone [60].
The altitude varies between 3152 and 4500 m.a.s.l. The lower altitude is found at the easternmost end of CNP, corresponding to the valleys, inter-Andean trenches and many of the towns in the country’s south, including Cuenca city. Meanwhile, the highest points have an elevation close to 4500 m.a.s.l. in the westernmost part of the park [57], highlighting “Cerro Arquitectos” with 4300 m (Figure 1b).

1.2. Geologic Framework

CNP is to the west of the western fringe of the Ecuadorian Andes, better known as the Western Cordillera (Figure 2a). Lithologically, it has various formations of volcanic origin ranging from the Eocene to the Quaternary. Among the geological units reported by Dunkley and Gaibor (Figure 2b) [62] are: the Saraguro group (E-Ms)–middle Eocene to early Miocene, formed by predominantly andesitic to dacitic calc-alkaline volcanic rocks; rhyolites are also common. The Chulo unit (E Sc), composed of a sequence of rhyolitic to dacitic tuffs, breccias and lacustrine sediments, despite the unknown age, is dated to the late Eocene. The Tomebamba unit (O Stb)–early Oligocene is distributed mainly in the Tomebamba valley (northern part of CNP), composed of massive andesitic and basic dacitic tuffs with lapilli. The Chanlud formation (O Scd)–early Oligocene presents lithologies of andesitic lavas with breccias and intercalations of volcanic sediments; in some sectors, there are dacitic and basaltic compositions. The Rio Blanco formation (O Srb)–early Oligocene (the western portion of CNP) comprises feldspathic andesitic lavas, tuffs and breccias intercalated with sandstones and dacitic tuffs. The La Soldados formation (O Ss) compounds by crystal-rich dacitic tuffs divided into three ash flow tuff units, abundant feldspar and quartz and chloritic lapilli. The Plancharumi formation (O Sp) has very poor lithification, composed of rhyolitic volcanoclastic and fluvial–lacustrine sediments; in some places, they present pumice-rich white ash tuffs.
Regionally, most of the main faults in mainland Ecuador follow a NE–SW direction [63] (Figure 2a). Due to the proximity of the CNP area to the Chaucha batholith and Quimsacocha caldera [64,65,66], both located in the surrounding areas of CNP, some faulting was generated inside its official protected area; most of these faulted sectors facilitated the passage of giant masses of ice (glaciers), represented by landforms of ice transport erosion, such as “U-shaped” valleys.
The historical geology of the sector can be summarised by several important events in the area: (i) the formation of the continental island arc due to the subduction of the Nazca (oceanic) Plate with the South American (Continental) Plate; (ii) a tectonic uplift, which started in the Miocene and currently continues; (iii) deposition of volcanic materials towards both sides of the main axis of elevations of the Western Cordillera in the course of high activity during the Miocene being less frequent in the Pliocene [66]; (iv) filling with fluvial–lacustrine sediments from depressions in the eastern part of the study area (Cuenca Basin) [67,68].

1.3. Geomorphological Framework

CNP topography is marked by volcanic arc activity in the northern part of the park [73] and the events of great glaciations of the late Pleistocene, which covered about 450 km2, distributed in the vicinity of the study area [74]. In addition, these volcaniclastic sediments (mostly tephras) are near the larger CNP lakes [73].
It presents a great abundance of small to medium-size lakes (about 250) [75], which are above 3000 m.a.s.l.; these are permanent, being fed by both the surface and groundwater [76]. Similarly, there are exhibited characteristic landforms, such as a “U” shaped valley and different kinds of moraines, which is a product of said glaciations (see Figure 3) [77]. Navarrete [53] reports several sites that stand out, for example, the glacial evidence of Llaviucu lake (709090 E; 9685471 N), the fluvial–glacial landforms of the Tomebamba River valley in the NE limit of the park, and the lakes that surround the segments of the Inca Trail or Qhapaq Ñan (north-central part of CNP).

2. Materials and Methods

This work was carried out through four phases (Figure 4): (i) base information analysis of the study area, (ii) identification and selection of sites of geological interest, (iii) geosites and geomorphosites assessment, and finally (iv) qualitative assessment, through the development of a strengths, weaknesses, opportunities and threats (SWOT) analysis.

2.1. Stage I: Base Information Analysis of the Study Area

The first stage considers the review of the information available on CNP, which includes: international scientific publications, park tourism websites, government documents, basic cartography of the study area and other technical documents from park management (e.g., plan management of CNP). All these data will serve to make a list of possible sites of interest, notified in said literature and the tourist trails at the park. Similarly, within this same phase, a literature review of methodologies used to assess geosites and geomorphosites was carried out to select the most appropriate methods to assess the sites.

2.2. Stage II: Identification and Selection of Sites of Geological Interest

The second stage consists of the identification of sites of geological interest (SGI) by processing the cartographic information collected in the previous phase, which includes: topographic maps, geological maps, geomorphological maps (e.g., MAG-IEE-SENPLADES Landforms Inventory on 1:25,000 scale [78]), digital terrain models (DTM) and ArcGIS digitisation of current tourist trails within the CNP boundary. Additionally, this stage included the public database’s geolocated points, which refer to the protected area’s natural resources, sites available in the CNP tourist inventory files and other geological elements presented in the scientific literature. Finally, potential sites for fieldwork were chosen using the Delphi method [79,80], which were directed to five experts with knowledge of the study area. The tool used was an interview (questionnaire) based on (i) importance of the geological heritage of CNP, (ii) proposal of potential sites, and (iii) expert’s self-evaluation. For this study, 14 SGIs were selected based on the representativeness of geological elements, accessibility and other additional values (e.g., cultural aspects) of the 51 SGIs registered and documented in the literature.

2.3. Stage III: Geosite and Geomorphosite Assessment

The third stage, the evaluation of the 14 geosites and geomorphosites, was developed using two methods, Brilha [32] and the Spanish Inventory of Places of Geological Interest (IELIG), designed by García-Cortés and Carcavilla [40]. These methods present great versatility in evaluating sites of any type, including those of geomorphological interest dominant in the CNP area. Both methods have been used for national inventories of sites of geological importance in Brazil and Spain [81].

2.3.1. Brilha Method

This method includes the evaluation of sites of geological importance based on four dimensions: scientific (S), use educational potential (UEP), potential tourism use (PTU) and degradation risk (DR) (Table 1). When evaluating the S, UEP and PTU indicators, a high value has a positive connotation for the geosite. However, when assessing the DR indicators, high values represent negative connotations (risk) for the geosite. It has been used since 1997 by the Geological Survey of Brazil (known as CPRM in Portuguese) and currently through the GEOSSIT platform, which includes a geological heritage inventory proposal available in an online database with national coverage in Brazil [82]. Furthermore, it has been used for inventory and the assessment of geosites and geomorphosites (e.g., [83,84,85]).
Based on the total degradation risk (DR) obtained per site, Brilha [32] proposed a classification of DR in three categories (Table 2). These categories allow the configuration of a geosite improvement strategy plan.

2.3.2. IELIG Method

IELIG, also known as the Spanish Inventory of Places of Geological Interest, was developed by the Geological and Mining Institute of Spain (IGME), in response to Law 42/2007 of the Government of Spain, on natural heritage and biodiversity to establish a national database. IELIG collaborated with ministries, local and regional governments and educational and scientific institutions in the country. This method considers three values or interests: scientific, academic and tourist. Each assessment parameter has a scale from 0 to 4; the score assigned is multiplied by its corresponding value weight (Equation (1)) (see the weight in Table 3). The structure of the evaluation criteria is detailed in Table 3, while the criteria for the evaluation of degradation susceptibility (DS) are in Table 4.
Total   ( value ) =   ( Score   ×   Weight ) ,
The degradation susceptibility was evaluated based on 12 parameters, of which 3 have a fragility focus, while the others focus on vulnerability (Table 4). Both values are used to calculate degradation susceptibility (Equation (2)).
DS = Fr . × Vul . 400 ,
Additionally, this method also includes a series of metrics that relate the three values (S, A and T), for the calculation of the protection priority (Pp), according to said values: scientific priority (S-Pp), priority academic (A-Pp) and recreation/tourism priority (T-Pp). Table 5 presents the detail of these equations.

2.4. Stage IV: Qualitative Assessment

Stage IV includes the qualitative analysis through a SWOT matrix [86], which explores four variables: strengths, weaknesses, opportunities and threats on the total list of evaluated geosite–geomorphosites. For the SWOT analysis, an evaluation was carried out under the criteria of five experts in general geology, mines, geomorphology, geological heritage and tourism, who are members of the academy and who know the area. The SWOT matrix allowed for a comprehensive evaluation considering positive and negative aspects of the sites and their environment to generate strategies that combine the development of geotourism and geoconservation with the current conservation plan of CNP.

3. Results

3.1. Geosites and Geomorphosites of CNP

This section presents a list of 14 geosites and geomorphosites and their main features, resulting from the inventory and evaluation process using the qualitative and semi-quantitative methodologies proposed in this study, in addition to the field trips carried out at CNP (Table 6). The geosites are classified into four areas of interest: (i) geomorphological, (ii) glacial, (iii) lithological and (iv) periglacial. In addition, Figure 5 shows some examples of the most representative geosites of CNP and Figure 6 shows the location of the 14 CNP geosites.
Figure 5 shows some of the 14 geosites that highlight the scenic value of glacial environments and volcanic origin elements present within CNP limits.

3.2. Geosite and Geomorphosite Assessment

The assessment of geosites and geomorphosites, through both methods, is detailed below.

3.2.1. Assessment by the Brilha Method

Table 7 shows the assessment through the Brilha method on the 14 geosites selected; the first 3 columns contain the study values of the method (S–scientific, UEP–educational and PTU–touristic). In all three values, the “moderate” category obtained the highest number of geosites (Figure 7a). Within the scientific category (S), the highest value was obtained by volcanic flows (G4), while the lowest value corresponds to the camp of erratic boulders/Llaviucu river (G9). In terms of educational potential, two geosites obtained the highest value (Larga and Negra lakes (G2) and Toreadora lake (G6)); when analysing the PTU, the highest value in the category is found in the Llaviucu end moraine (G13), while the lowest value in this category belongs to volcanic flows (G4) (Table 7).
The results obtained for the degradation risk (DR) of the 14 geosites (Figure 7b) show that 57.14% of the total present values are within the “moderate” category. The remaining 42.86% falls within the “low” degradation risk category. Of the 14 geosites, the highest values of the degradation risk evaluation, with a value of 250/400, correspond to the camp of erratic boulders/Llaviucu river (G10) and Llaviucu end moraine (G12) (Table 3).

3.2.2. Assessment by the IELIG Method

Table 8 presents the values obtained from the application of the IELIG method to the 14 evaluated geosites; the evaluation considered the values of the interests (S, A and T). The Llaviucu valley (310) obtained the highest average score of the indices (S, A and T), while the rocky volcanic outcrop G3 (196.67) had the lowest average score of the above indices. According to the classification suggested by García-Cortés and Carcavilla [40], in the 3 interests, the 14 evaluated geosites present values in the 2 classification categories with the highest score: “very high” and “high” (Figure 8).
On the other hand, the degree of susceptibility of the 14 geosites (Figure 9a) presented results within the categories: “high”, “medium” and “low”. It mostly reflected a classification within the “medium” category, with 71.43% corresponding to 10 geosites; some geosites that fall into this category are Drumlins camp G5 (19.50), Toreadora lake G6 (42.50) and Llaviucu valley viewpoint G9 (23.25). The second category with the highest proportion was “high” with 21.43%, equivalent to 3 out of 14 sites, where the rocky volcanic outcrop G3 geosite obtained the highest score (98.00). The “low” category only presents one geosite: Tres Cruces peak G1 (12.00) (Table 4), equivalent to 7.14% of all geosites.
Additionally, Table 8 shows the global protection priority (Pp) calculated for the 14 geosites. Figure 9b demonstrates that 7/14 geosites were classified within the “high” category, corresponding to 50.00% of the total evaluated. The highest value within this category was the rocky volcanic outcrop G3 geosite (23.69). Meanwhile, the remaining one constitutes 50.00% of the geosites classified within the “medium” category. Tres Cruces peak G1 (5.95) was the lowest value within this category and the total number of geosites.

3.3. Qualitative Assessment

Application of the SWOT analysis allowed us to define the park’s main strengths, weak points and opportunities if geotourism is developed in the area. In addition, this study detected some natural (e.g., the risk of erosion and landslides) and anthropogenic (e.g., water and soil contamination) threats. Table 9 presents a summary of the qualitative evaluation developed.
Based on the SWOT analysis (Table 9), five lines of action are proposed, which focus on the characterisation, protection, promotion and monitoring of the evaluated geosites–geomorphosites.
  • Design an inventory and characterisation protocol for the park’s geosites, articulated with current natural heritage inventory programs (flora and fauna), that integrates all the stakeholders dedicated to the conservation of CNP;
  • Creation and installation of geological and geomorphological interpretive panels. These panels must combine a real photograph of the geosites–geomorphosites with graphic schemes illustrating the geological process of interest. A good practice is the implementation of the ABC (abiotic, biotic and cultural) interpretive concept [88] in the CNP panel system;
  • Development of panoramic view areas, signage, information panels and safety means (e.g., installation of safety railings) at specific points where geomorphosites can be better observed and continue with the construction of trails and other facilities for tourism in some geosites (e.g., Miguir meteorite);
  • Propose pilot projects for the development of geotourism that integrate the values of cultural and natural interest present in the current tourist trails of the park with the geological value of the sites characterised in this study;
  • Include geotourism guides and descriptions of the geological value in the information brochures and other informative material promoting geoheritage and geomorphological heritage education.

4. Discussion

Cajas National Park has elements of geoheritage where the glacial and periglacial landscapes and landforms stand out, evidencing the glaciation and deglaciation processes that occurred during the Pleistocene. These events have left clear signs of their occurrence, represented by the sites of geological importance suggested in this study (Table 6). Depending on the type of main geological interest, the glacial interest constitutes a considerable part of the total. Additionally, the 14 evaluated sites of geological importance (Table 7 and Table 8) present in both methods have predominantly scientific values, with categories mainly varying from “moderate” to “high” (Figure 7 and Figure 8).
The results of the assessment of geological interest sites of CNP indicate that the high scientific/educational value comes from sites where a high aesthetic value is concentrated, resulting from the modelling of more than one geological–geomorphological process (e.g., Llaviucu valley). This opens up the possibility of analysing other types of studies (e.g., hydrological, climatological and biodiversity). One example is the lakes, which constitute an essential cover letter to disseminate the scientific content of the area to people of any educational level.
The tourism component reflected significant qualities for its use in recreational activities due to sports activities (e.g., hiking and recreational fishing) and cultural activities (e.g., birdwatching and religious ceremonies). There is a clear influence of sites with high landscape value over those with high scientific/educational value (e.g., volcanic flows), due to their limited promotional facilities and tourist infrastructure.
The degradation risk parameter is a key factor in evaluating geoheritage due to its influence on the level of use to develop geotourism activities and geoconservation strategies [8,89,90,91]. Both methods (Brilha [32] and IELIG [40]) assess similar parameters (e.g., legal status and proximity to degrading activities). However, IELIG [40] provides more detail on the indicators/sub-indicators, classifying them into fragility (natural conditions) and vulnerability due to anthropic threats.
The study area has many lakes that have historically been influenced by disturbances of the natural state of soil and water [92,93] and by the effects of climate change, which reduce the extension of water bodies within the park [75].
The assessments show that most geosites have medium levels of risk (Figure 7b and Figure 9). This is because many geosites, such as the lakes, have a monitoring system, which is a product of the conservation plan that CNP has, and there are studies of natural threats that have led to the identification and execution of certain prevention actions by natural agents (e.g., erosion and floods).
In general, sites of geological interest, due to their characteristics, provide an opportunity to raise awareness about climate change for tourists and surrounding communities. For example, U-shaped valleys, hanging valleys, eroded relief forms, large lakes and other landforms of glacial origin are witnesses to the fluctuations of the masses of glaciers influenced by Quaternary climate change modelled in this region. Moreover, climate change constitutes one of the key aspects of the role of geology in the sustainable development goals (SDGs) [94].
Within the country, some of the few evaluations of geosites with glacial origin are those carried out in the Chimborazo Province [95] and Ruta Escondida (Quito) [96,97]. These studies presented significant values for promotion and sustainable management. Similarly to CNP, they are located within protected areas with many tourist services, thereby increasing administration opportunities under an emerging geotourism approach. At an international level, the experiences of geosite evaluations of glacial origin show that, as a whole, they present values of geological interest in the highest categories of the classification [98,99]. The geosites with the highest scores are located where aesthetic values are complemented by other topics of importance to the user (e.g., accessibility) and a level of educational and tourist use [98]. Similarly, the sites with the highest scores in both methods in this study (e.g., Llaviucu valley and Toreadora lake) are those where the administration uses greater outreach resources and tourist facilities.
Regarding the methodology applied, the first step was the selection of sites of initial geological interest recommended by several authors [40,100,101,102]. In this case, that meant applying the Delphi method [79,80] with the information collected in the first phase. This step is crucial to understanding the geological and geomorphological context. After this stage, the selected sites are assessed by the semi-quantitative methods of Brilha [32] and IELIG [40]; both methods integrally evaluate the geosites in the different sections that compose them. The IELIG method [40] provides the protection priority value, which is essential for establishing improvement strategies. Additionally, the chosen methods have parameters that include the types of interest of CNP and have the aesthetic value within the evaluation, which is an important factor in geomorphosites [103].
In the Brilha method, the aesthetic value is indirectly in the educational potential and potential touristic use (scenery in Table 1) [32], as the beauty of the site and its frequency of use in national and local campaigns. On the other hand, the IELIG method [40] (beauty in Table 3) assessed beauty in terms of the coincidence of three characteristics: (i) relief amplitude, (ii) mighty river courses/large sheets of water (or ice), and (iii) remarkable chromatic variety. In the first case, the indicators focus on attractiveness to be visited. The second case is on the visual quality of its features. Reynard et al. [100] highlight the subjectivity of the aesthetic value and the difficulty of measurement; it is clear that this is a limitation of the evaluations since it depends on visual perception. However, the use of both methods is adequate for the present study.
Given the difficulty of making direct comparisons between the values of geosites characterised by different methodologies [104], some studies have approached a comparative description of the values obtained with different methods. In general, for an adequate comparison, quantitative assessment values are transformed into percentages (e.g., in the IELIG and Brilha methods, 400 to 0 into 100% to 0%) in the first phase [104]. Furthermore, according to Berrezueta et al. [54], using geosites that have been assessed by more than one method allows us to calculate a factor that relates the IELIG method to the Brilha method (e.g., the Brilha to IELIG factor is 0.82 in geosites from Ecuador). In an attempt to unify the process of characterisation and assessment of sites of geological interest in Ibero-American countries, in 2018 the Association of Geological Services of Ibero-America (ASGMI in Spanish) presented [105] an action protocol based on the IELIG method [40]. As complement to this method, the Brilha method [32] was also used.
This study shows a general trend in the application of the two geosite evaluation methods. Geosites (e.g., Miguir meteorite (G7), Llaviucu valley (G9) and Llaviucu lake (G11)) are among the top 5 best scored in both methods for their scientific value (Table 7 and Table 8). However, the difference lies in the sub-indicators of each method and the ranges established for their classification, as can be seen in the graphs of the global evaluations of the 14 sites in the 3 dimensions for both the Brilha method (Figure 7, whose categories range from “low”, “moderate” to “high”) and the IELIG method (Figure 8, which is dominated by the “high” and “very high” categories).
The SWOT analysis results recommend a protocol for the inventory and characterisation of geological and geomorphological heritage, even though CNP has national [50] and international recognition for its biotic resources [51,52]. The abiotic part of CNP requires greater visibility. This study evidences the lack of a holistic approach that integrates the abiotic, biotic and cultural domains of the natural heritage of CNP. Therefore, it is essential to encourage geoscientific research and promote geoeducation in the interpretive panels and other tourist services offered by the park. Implementing the ABC interpretive concept is an example of one approach used in geoparks (e.g., Colca and Volcanoes Andagua and Muroto geoparks) [88].
Geomorphosites are natural tourist resources with interesting economic benefits, especially in protected areas such as national parks and natural monuments [106]. Natural parks have geological characteristics relevant to the development of geotourism and geoeducation, and present opportunities for the protection of geological elements (e.g., Fforest Fawr Geopark [11] and Cilento Vallo di Diano Geopark [12]). It is important to highlight the identity of natural parks with cultural and natural aspects [11]. Articulating the abiotic axis through geotourism will complement its protection and promote sustainable development in addition to the opening of research projects, marketing, events, workshops and scientific publications.
Finally, the geosites and geomorphosites evaluated in this research make an initial contribution to the knowledge of the geological heritage of natural parks. The geoheritage examples from this investigation illustrate the unique geological features that marked the area’s geological history. However, more geoheritage characterisation studies and geoheritage value information within the CNP conservation plan are needed. It is also important to highlight that the inventory and characterisation of the geomorphosites which have been carried out represent a basis for the proposal of other initiatives. For example, the creation of a UNESCO geopark [107] or the presentation of specific geological itineraries. In this study, since there were official protection figures such as national parks in the study area, the contributions have been framed within this official figure and are intended to be an example extended to other areas with existing protection figures.

5. Conclusions

The present investigation reveals the singularity of glacial and periglacial landscapes and landforms that tell the geological and geomorphological history of the ate Pleistocene of the region. This study evaluated 14 geologically significant sites using 2 international geosite evaluation methods. Four types of main geological interest were identified: (i) geomorphological, (ii) glacial, (iii) lithological and (iv) periglacial. Geosites represented scientific and educational value with exceptional aesthetic value (e.g., Llaviucu valley) and their educational value to illustrate the evolution of geological processes in CNP. In addition, the area has high geodiversity and geoheritage, as well as connections with the biotic and cultural dimensions recognised internationally.
Threats against site degradation reflect the “medium” degree of vulnerability against natural and anthropogenic events (e.g., landslides, agricultural and cattle activities). Additionally, the degree of priority of the sites was evidenced, reaching significant scores concerning the total and illustrating the priority that geosites and the CNP environment should have to protect their geoheritage.
The SWOT analysis results identified the main strengths (e.g., the existence of a system of tourist services that includes a large part of the places of geological interest) and weaknesses (e.g., the absence of studies and scientific–informative publications of geological characterisation). This qualitative analysis allowed us to establish five strategies focused on: (i) the design of an inventory protocol and characterisation of geoheritage; (ii) the establishment of interpretive panels using the ABC interpretive concept; (iii) continuing with the construction of trails and other facilities for tourists; (iv) generating pilot projects for the development of geotourism that integrate abiotic, biotic and cultural dimensions; (v) creating geological guides and including the value of the geoheritage and geomorphological heritage of the park in the informative material.
Finally, this work strengthens geological heritage, biodiversity and cultural benefits that the natural park houses, and shows us the need for more exhaustive investigations on the subject of glaciation for its application in geoheritage.

Author Contributions

Conceptualisation, E.N., F.M.-C., J.D.-T., P.C.-M., M.J.-M. and E.B.; methodology, E.N., F.M.-C., J.D.-T., P.C.-M., M.J.-M. and E.B.; investigation, E.N., F.M.-C., J.D.-T., P.C.-M., M.J.-M. and E.B.; writing—original draft preparation, E.N., F.M.-C., J.D.-T., P.C.-M., M.J.-M. and E.B.; writing—review and editing, E.N., F.M.-C., J.D.-T., P.C.-M., M.J.-M. and E.B.; supervision, P.C.-M. and E.B. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by the ‘Registry of geological and mining heritage and its impact on the defence and preservation of geodiversity in Ecuador’ academic research project by ESPOL University. CIPAT-01-2018.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Not applicable.

Acknowledgments

The authors would like to thank Instituto de Investigación Geológico y Energético (IIGE), Cajas National Park (CNP) and the Red Minería XXI (CYTED: 407310RT0402, IGME) for their information and support in developing the present scientific work. In addition, the authors would like to acknowledge the international project entitled ‘Ibero-American mineral routes and territorial planning: an integral factor for the sustainable development of society’ (the acronym in Spanish is RUMYS).

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Serrano, E.; Ruiz-Flaño, P. Geodiversity: A theoretical and applied concept. Geogr. Helv. 2007, 62, 140–147. [Google Scholar] [CrossRef]
  2. ProGEO. Conserving Our Shared Geoheritage—A Protocol on Geoconservation Principles, Sustainable Site Use, Management, Fieldwork, Fossil and Mineral Collecting. 2011. Available online: https://www.sigeaweb.it/geoheritage/documents/progeo-protocol-definitions-20110915.pdf (accessed on 7 June 2021).
  3. De Wever, P.; Guiraud, M. Geoheritage and Museums. In Geoheritage; Reynard, E., Brilha, J., Eds.; Elsevier: Amsterdam, The Netherlands, 2018; pp. 129–145. [Google Scholar]
  4. Brilha, J. Geoheritage: Inventories and evaluation. In Geoheritage; Reynard, E., Brilha, J., Eds.; Elsevier: Amsterdam, The Netherlands, 2018; pp. 69–85. [Google Scholar]
  5. Gordon, J.E.; Barron, H.F.; Hansom, J.D.; Thomas, M.F. Engaging with geodiversity—Why it matters. Proc. Geol. Assoc. 2012, 123, 1–6. [Google Scholar] [CrossRef] [Green Version]
  6. Wimbledon, W.A.P.; Smith-Meyer, S. Geoheritage in Europe and Its Conservation; ProGEO: Oslo, Norway, 2012. [Google Scholar]
  7. Erikstad, L. Geoheritage and geodiversity management—the questions for tomorrow. Proc. Geol. Assoc. 2013, 124, 713–719. [Google Scholar] [CrossRef]
  8. Garavaglia, V.; Diolaiuti, G.; Smiraglia, C.; Pasquale, V.; Pelfini, M. Evaluating Tourist Perception of Environmental Changes as a Contribution to Managing Natural Resources in Glacierized Areas: A Case Study of the Forni Glacier (Stelvio National Park, Italian Alps). Environ. Manag. 2012, 50, 1125–1138. [Google Scholar] [CrossRef]
  9. Vukoičić, D.; Milosavljević, S.; Valjarević, A.; Nikolić, M.; Srećković-Batoćanin, D. The evaluation of geosites in the territory of National park “Kopaonik” (Serbia). Open Geosci. 2018, 10, 618–633. [Google Scholar] [CrossRef]
  10. Bucci, F.; Tavarnelli, E.; Novellino, R.; Palladino, G.; Guglielmi, P.; Laurita, S.; Prosser, G.; Bentivenga, M. The History of the Southern Apennines of Italy Preserved in the Geosites Along a Geological Itinerary in the High Agri Valley. Geoheritage 2019, 11, 1489–1508. [Google Scholar] [CrossRef]
  11. Ramsay, T. Fforest fawr geopark—A UNESCO global geopark distinguished by its geological, industrial and cultural heritage. Proc. Geol. Assoc. 2017, 128, 500–509. [Google Scholar] [CrossRef]
  12. Santangelo, N.; Romano, P.; Santo, A. Geo-itineraries in the Cilento Vallo di Diano Geopark: A Tool for Tourism Development in Southern Italy. Geoheritage 2015, 7, 319–335. [Google Scholar] [CrossRef]
  13. Keever, P.J.M.; Zouros, N. Geoparks: Celebrating Earth heritage, sustaining local communities. Episodes 2005, 28, 274–278. [Google Scholar] [CrossRef] [Green Version]
  14. ProGEO. Digne-Les-Bains Declaration-Declaration of the Rights of the Memory of the Earth; ProGEO: Digne, France, 1991. [Google Scholar]
  15. Bruno, D.E.; Crowley, B.E.; Gutak, J.M.; Moroni, A.; Nazarenko, O.V.; Oheim, K.B.; Ruban, D.A.; Tiess, G.; Zorina, S.O. Paleogeography as geological heritage: Developing geosite classification. Earth Sci. Rev. 2014, 138, 300–312. [Google Scholar] [CrossRef]
  16. Reynard, E. Geosite. In Encyclopedia of Geomorphology; Goudie, A., Ed.; Routledge: London, UK, 2014; Volume 1, ISBN 9780415863001. [Google Scholar]
  17. Ruban, D.A.; Kuo, I. Essentials of geological heritage site (geosite) management: A conceptual assessment of interests and conflicts. Nat. Nascosta 2010, 41, 16–31. [Google Scholar]
  18. Panizza, M. Geomorphosites: Concepts, methods and examples of geomorphological survey. Chinese Sci. Bull. 2001, 46, 4–5. [Google Scholar] [CrossRef]
  19. Reynard, E. Geomorphosites: Definitions and characteristics. In Geomorphosites; Verlag Dr. Friedrich Pfeil: München, Germany, 2009; pp. 9–20. ISBN 978-3-89937-094-2. [Google Scholar]
  20. Pescatore, E.; Bentivenga, M.; Giano, S.I.; Siervo, V. Geomorphosites: Versatile Tools in Geoheritage Cultural Dissemination. Geoheritage 2019, 11, 1583–1601. [Google Scholar] [CrossRef]
  21. Coratza, P.; De Waele, J. Geomorphosites and Natural Hazards: Teaching the Importance of Geomorphology in Society. Geoheritage 2012, 4, 195–203. [Google Scholar] [CrossRef]
  22. Quesada-Román, A.; Pérez-Umaña, D. Tropical Paleoglacial Geoheritage Inventory for Geotourism Management of Chirripó National Park, Costa Rica. Geoheritage 2020, 12, 58. [Google Scholar] [CrossRef]
  23. Jon, W.-S.; Ryang, D.-Z.; Ri, H.-Y. Natural Heritage Value of Mt. Kumgang and Global Comparative Analysis. Geoheritage 2020, 12, 32. [Google Scholar] [CrossRef]
  24. Tavera Escobar, M.A.; Estrada Sierra, N.; Errázuriz Henao, C.; Hermelin, M. Georutas o itinerarios geológicos: Un modelo de geoturismo en el Complejo Volcánico Glaciar Ruiz-Tolima, Cordillera Central de Colombia. Cuad. Geogr. Rev. Colomb. Geogr. 2017, 26, 219–240. [Google Scholar] [CrossRef]
  25. Perret, A.; Reynard, E. A method for selecting potential geosites. The case of glacial geosites in the Chablais area (French and Swiss Prealps). In Proceedings of the EGU General Assembly Conference Abstracts, Vienna, Austria, 27 April–2 May 2014; p. 12035. [Google Scholar]
  26. Vieira, G.; de Castro, E.; Gomes, H.; Loureiro, F.; Fernandes, M.; Patrocínio, F.; Firmino, G.; Forte, J. The Estrela Geopark—From Planation Surfaces to Glacial Erosion. In Landscapes and Landforms of Portugal; Springer: Cham, Switzerland, 2020; pp. 341–357. [Google Scholar]
  27. Pérez Alberti, A. El patrimonio glaciar y periglaciar del Geoparque Mundial UNESCO Montañas do Courel (Galicia). Cuatern. Y Geomorfol. 2021, 35, 79–98. [Google Scholar] [CrossRef]
  28. Mucivuna, V.C.; Reynard, E.; da Garcia, M.M. Geomorphosites Assessment Methods: Comparative Analysis and Typology. Geoheritage 2019, 11, 1799–1815. [Google Scholar] [CrossRef]
  29. Zafeiropoulos, G.; Drinia, H.; Antonarakou, A.; Zouros, N. From Geoheritage to Geoeducation, Geoethics and Geotourism: A Critical Evaluation of the Greek Region. Geosciences 2021, 11, 381. [Google Scholar] [CrossRef]
  30. Fassoulas, C.; Mouriki, D.; Dimitriou-Nikolakis, P.; Iliopoulos, G. Quantitative Assessment of Geotopes as an Effective Tool for Geoheritage Management. Geoheritage 2012, 4, 177–193. [Google Scholar] [CrossRef]
  31. 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]
  32. Brilha, J. Inventory and Quantitative Assessment of Geosites and Geodiversity Sites: A Review. Geoheritage 2016, 8, 119–134. [Google Scholar] [CrossRef] [Green Version]
  33. 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]
  34. De Lima, F.F.; Brilha, J.B.; Salamuni, E. Inventorying Geological Heritage in Large Territories: A Methodological Proposal Applied to Brazil. Geoheritage 2010, 2, 91–99. [Google Scholar] [CrossRef] [Green Version]
  35. Kubalíková, L. Koncepce Geomorphosites v Kontextu Ochrany Neživé Přírody; Technická Univerzita v Liberci: Liberec, Czech Republic, 2012; ISBN 978-80-7372-932-5. [Google Scholar]
  36. Pralong, J.-P.; Reynard, E. A proposal for a classification of geomorphological sites depending on their tourist value. Ital. J. Quat. Sci. 2005, 18, 315–321. [Google Scholar]
  37. Pralong, J.-P. A method for assessing tourist potential and use of geomorphological sites. Géomorphol. Reli. Process. Environ. 2005, 11, 189–196. [Google Scholar] [CrossRef]
  38. Bruschi, V.M.; Cendrero, A. Geosite evaluation; can we measure intangible values? Quaternario 2005, 18, 293–306. [Google Scholar]
  39. Coratza, P.; Giusti, C. Methodological proposal for the assessment of the scientific quality of geomorphosites. Quaternario 2005, 18, 307–313. [Google Scholar]
  40. García-Cortés, A.; Carcavilla, L. Documento Metodológico Para la Elaboración del Inventario Español de Lugares de Interés Geológico (IELIG); Instituto Geolớgico y Minero de España (IGME): Madrid, Spain, 2013. [Google Scholar]
  41. Vujičić, M.D.; Vasiljević, D.A.; Marković, S.B.; Hose, T.A.; Lukić, T.; Hadžić, O.; Janićević, S. Preliminary geosite assessment model (GAM) and its application on Fruška gora mountain, potential geotourism destination of Serbia. Acta Geogr. Slov. 2011, 51, 361–376. [Google Scholar] [CrossRef]
  42. Comănescu, L.; Nedelea, A. Geomorphosites Assessments of the Glacial and Periglacial Landforms from Southern Carpathians. In Landform Dynamics and Evolution in Romania; Radoane, M., Vespremeanu-Stroe, A., Eds.; Springer: Cham, Switzerland, 2017; pp. 215–246. [Google Scholar]
  43. Ólafsdóttir, R.; Tverijonaite, E. Geotourism: A Systematic Literature Review. Geosciences 2018, 8, 234. [Google Scholar] [CrossRef] [Green Version]
  44. 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]
  45. Hose, T.A. Selling the story of Britain’s stone. Environ. Interpret. 1995, 10, 16–17. [Google Scholar]
  46. Convention on Biological Diversity. Available online: https://www.cbd.int/countries/profile/?country=ec (accessed on 23 August 2021).
  47. 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. Ecodynamics 2020, 15, 39–46. [Google Scholar] [CrossRef]
  48. Rieckmann, M.; Adomßent, M.; Härdtle, W.; Aguirre, P. Sustainable Development and Conservation of Biodiversity Hotspots in Latin America: The Case of Ecuador. In Biodiversity Hotspots; Zachos, F., Habel, J., Eds.; Springer: Berlin/Heidelberg, Germany, 2011; pp. 435–452. [Google Scholar]
  49. Sánchez-Cortez, J.L. Conservation of geoheritage in Ecuador: Situation and perspectives. Int. J. Geoheritage Park. 2019, 7, 91–101. [Google Scholar] [CrossRef]
  50. Ministerio de Ambiente y Agua-República del Ecuador Sistema Nacional de Áreas Protegidas del Ecuador. Available online: http://areasprotegidas.ambiente.gob.ec/es (accessed on 31 October 2021).
  51. United Nations Educational, Scientific and Cultural Organization (UNESCO). Macizo del Cajas Biosphere Reserve, Ecuador. Available online: https://en.unesco.org/biosphere/lac/macizo-del-cajas (accessed on 23 August 2021).
  52. SISR Parque Nacional Cajas | Servicio de Información sobre Sitios Ramsar. Available online: https://rsis.ramsar.org/es/ris/1203?language=es (accessed on 23 August 2021).
  53. Navarrete, E. Huellas de un “frío pasado”. Patrimonio Geológico del Parque Nacional “El Cajas.” In El Patrimonio Geominero en el Contexto de la Ordenación Territorial; Mistral: Guayaquil, Ecuador, 2005; pp. 181–197. [Google Scholar]
  54. Berrezueta, E.; Sánchez-Cortez, J.L.; Aguilar-Aguilar, M. Inventory and Characterization of Geosites in Ecuador: A Review. Geoheritage 2021, 13, 93. [Google Scholar] [CrossRef]
  55. De la Vega, G. Comentarios Reales de los Incas; Biblioteca de Autores Españoles: Madrid, Spain, 1609; ISBN 9681648927. [Google Scholar]
  56. ETAPA E.P.; GAD Municipal Cuenca; MAE. Plan de Manejo del Parque Nacional El Cajas; INSIGMA: Cuenca, Ecuador, 2018. [Google Scholar]
  57. ETAPA E.P.; MAE. Expediente para la Inscripción del Parque Nacional El Cajas y los Tramos Transversales del Qhapaq Ñan en la Lista de Patrimonio Mundial de la UNESCO; UNESCO: Cuenca, Ecuador, 2007. [Google Scholar]
  58. Ministerio de Ambiente Agua y Transición Ecológica-República del Ecuador Parque Nacional El Cajas—Ministerio del Ambiente, Agua y Transición Ecológica. Available online: https://www.ambiente.gob.ec/parque-nacional-cajas/ (accessed on 11 November 2021).
  59. Báez, S.; Ospina, P.; Ramón, G. Una Breve Historia del Espacio Ecuatoriano; CAMAREN: Instituto de Estudios Ecuatorianos (IEE): Quito, Ecuador, 2004. [Google Scholar]
  60. Buytaert, W.; Sevink, J.; De Leeuw, B.; Deckers, J. Clay mineralogy of the soils in the south Ecuadorian páramo region. Geoderma 2005, 127, 114–129. [Google Scholar] [CrossRef]
  61. Buytaert, W.; Poyck, S.; Iñiguez, V.; De Bièvre, B. Hill slope flow in organic soils, a case study from the Ecuadorian páramo ecosystem. In Geophysical Research Abstracts; European Geosciences Union: Munich, Germany, 2006. [Google Scholar]
  62. Dunkley, P.; Gaibor, A. Mapa Geológico de la Cordillera Occidental del Ecuador Entre 2°–3° S, Escala: 1:200,000; Instituto de Investigación Geológico y Energético (IIGE): Quito, Ecuador, 1997. [Google Scholar]
  63. Eguez, A.; Alvarado, A.Y.H.; Machette, M.N.; Costa, C.; Dart, R.L. Map of Quaternary Faults and Folds of Ecuador and Its Offshore Regions. Available online: https://pubs.usgs.gov/of/2003/ofr-03-289/ (accessed on 23 August 2021).
  64. Beate, B.; Monzier, M.; Spikings, R.; Cotten, J.; Silva, J.; Bourdon, E.; Eissen, J.-P. Mio–Pliocene adakite generation related to flat subduction in southern Ecuador: The Quimsacocha volcanic center. Earth Planet. Sci. Lett. 2001, 192, 561–570. [Google Scholar] [CrossRef]
  65. Schütte, P.; Chiaradia, M.; Barra, F.; Villagómez, D.; Beate, B. Metallogenic features of Miocene porphyry Cu and porphyry-related mineral deposits in Ecuador revealed by Re-Os, 40Ar/39Ar, and U-Pb geochronology. Miner. Depos. 2012, 47, 383–410. [Google Scholar] [CrossRef] [Green Version]
  66. Hall, M.L.; Calle, J. Geochronological control for the main tectonic-magmatic events of Ecuador. Earth Sci. Rev. 1982, 18, 215–239. [Google Scholar] [CrossRef]
  67. Coltorti, M.; Ollier, C.D. Geomorphic and tectonic evolution of the Ecuadorian Andes. Geomorphology 2000, 32, 1–19. [Google Scholar] [CrossRef]
  68. Barberi, F.; Coltelli, M.; Ferrara, G.; Innocenti, F.; Navarro, J.M.; Santacroce, R. Plio-Quaternary volcanism in Ecuador. Geol. Mag. 1988, 125, 1–14. [Google Scholar] [CrossRef]
  69. Aspden, J.A.; Litherland, M. The geology and Mesozoic collisional history of the Cordillera Real, Ecuador. Tectonophysics 1992, 205, 187–204. [Google Scholar] [CrossRef]
  70. BGS-CODIGEM. Mapa Geologico de la República Del Ecuador. Escala, 1:1000000; Instituto de Investigación Geológico y Energético (IIGE): Quito, Ecuador, 1983. [Google Scholar]
  71. Cediel, F.; Shaw, R.; Cáceres, C. Tectonic assembly of the northern Andean block. Assoc. Pet. Geol. Mem. 2003, 79, 815–848. [Google Scholar]
  72. Jackson, L.J.; Horton, B.K.; Vallejo, C. Detrital zircon U-Pb geochronology of modern Andean rivers in Ecuador: Fingerprinting tectonic provinces and assessing downstream propagation of provenance signals. Geosphere 2019, 15, 1943–1957. [Google Scholar] [CrossRef] [Green Version]
  73. Rodbell, D.T.; Bagnato, S.; Nebolini, J.C.; Seltzer, G.O.; Abbott, M.B. A Late Glacial–Holocene Tephrochronology for Glacial Lakes in Southern Ecuador. Quat. Res. 2002, 57, 343–354. [Google Scholar] [CrossRef]
  74. Rodbell Donald, T.; Nebolini Jeffrey, C.; Seltzer Geoffrey, O.; Hansen Barbara, C.S.; Abbott Mark, B. Tephrochronology, sedimentology, and palynology of late glacial-Holocene lake sediment cores from southern Ecuador. Abstr. Programs-Geol. Soc. Am. 1996, 28, 471. [Google Scholar]
  75. Hansen, B.C.; Rodbell, D.; Seltzer, G.; León, B.; Young, K.; Abbott, M. Late-glacial and Holocene vegetational history from two sites in the western Cordillera of southwestern Ecuador. Palaeogeogr. Palaeoclimatol. Palaeoecol. 2003, 194, 79–108. [Google Scholar] [CrossRef]
  76. Carrasco, M. Estudio Limnológico y de Macroinvertebrados Bentónico en la Zona de Surocucho; Universidad del Azuay: Cuenca, Ecuador, 1996. [Google Scholar]
  77. Mosquera, P.V.; Hampel, H.; Vázquez, R.F.; Alonso, M.; Catalan, J. Abundance and morphometry changes across the high-mountain lake-size gradient in the tropical Andes of Southern Ecuador. Water Resour. Res. 2017, 53, 7269–7280. [Google Scholar] [CrossRef] [Green Version]
  78. MAG-IEE-SENPLADES. Mosaico homologado a nivel nacional, 1:25,000. 2009-2015 Geomorfología (Formas del Relieve). Available online: http://geoportal.agricultura.gob.ec/ (accessed on 7 June 2021).
  79. McKenna, H.P. The Delphi technique: A worthwhile research approach for nursing? J. Adv. Nurs. 1994, 19, 1221–1225. [Google Scholar] [CrossRef]
  80. Hasson, F.; Keeney, S. Enhancing rigour in the Delphi technique research. Technol. Forecast. Soc. Chang. 2011, 78, 1695–1704. [Google Scholar] [CrossRef]
  81. Mucivuna, V.C.; Motta Garcia, M.d.G.; Reynard, E. Comparing quantitative methods on the evaluation of scientific value in geosites: Analysis from the Itatiaia National Park, Brazil. Geomorphology 2022, 396, 107988. [Google Scholar] [CrossRef]
  82. Garcia, M.d.G.M.; Brilha, J.; de Lima, F.F.; Vargas, J.C.; Pérez-Aguilar, A.; Alves, A.; Campanha, G.A.d.C.; Duleba, W.; Faleiros, F.M.; Fernandes, L.A.; et al. The Inventory of Geological Heritage of the State of São Paulo, Brazil: Methodological Basis, Results and Perspectives. Geoheritage 2018, 10, 239–258. [Google Scholar] [CrossRef] [Green Version]
  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, 35. [Google Scholar] [CrossRef]
  84. Santos, D.S.; Reynard, E.; Mansur, K.L.; Seoane, J.C.S. The Specificities of Geomorphosites and Their Influence on Assessment Procedures: A Methodological Comparison. Geoheritage 2019, 11, 2045–2064. [Google Scholar] [CrossRef]
  85. 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]
  86. Dyson, R.G. Strategic development and SWOT analysis at the University of Warwick. Eur. J. Oper. Res. 2004, 152, 631–640. [Google Scholar] [CrossRef]
  87. Toulkeridis, T.; Echegaray-Aveiga, R.C.; Martinez-Maldonado, K.P. Shock metamorphism in volcanic rock due to the impact of the Miguir-Cajas meteorite in 1995 and its importance for Ecuador. Geoj. Tour. Geosites 2021, 35, 315–321. [Google Scholar] [CrossRef]
  88. Pásková, M.; Zelenka, J.; Ogasawara, T.; Zavala, B.; Astete, I. The ABC Concept—Value Added to the Earth Heritage Interpretation? Geoheritage 2021, 13, 38. [Google Scholar] [CrossRef]
  89. Ilieş, D.C.; Dehoorne, O.; Ilieş, A. Some examples of natural hazards affecting geosites and tourist activities. Geoj. Tour. Geosites 2011, 7, 33–38. [Google Scholar]
  90. Margiotta, S.; Negri, S.; Parise, M.; Quarta, T.A.M. Karst geosites at risk of collapse: The sinkholes at Nociglia (Apulia, SE Italy). Environ. Earth Sci. 2016, 75, 8. [Google Scholar] [CrossRef]
  91. Costa-Casais, M.; Caetano Alves, M.I. Geological Heritage at Risk in NW Spain. Quaternary Deposits and Landforms of “Southern Coast” (Baiona-A Garda). Geoheritage 2013, 5, 227–248. [Google Scholar] [CrossRef]
  92. Fränkl, L.A. A 20th Century Pollution History Reconstruction Using Lake Sediments from Cajas National Park, South Central Ecuador. Master’s Thesis, University of Bern, Bern, Switzerland, 2016. [Google Scholar]
  93. Schneider, T.; Musa Bandowe, B.A.; Bigalke, M.; Mestrot, A.; Hampel, H.; Mosquera, P.V.; Fränkl, L.; Wienhues, G.; Vogel, H.; Tylmann, W.; et al. 250-year records of mercury and trace element deposition in two lakes from Cajas National Park, SW Ecuadorian Andes. Environ. Sci. Pollut. Res. 2021, 28, 16227–16243. [Google Scholar] [CrossRef] [PubMed]
  94. Gill, J.C. Geology and the Sustainable Development Goals. Episodes 2017, 40, 70–76. [Google Scholar] [CrossRef]
  95. Carrión-Mero, P.; Borja-Bernal, C.; Herrera-Franco, G.; Morante-Carballo, F.; Jaya-Montalvo, M.; Maldonado-Zamora, A.; Paz-Salas, N.; Berrezueta, E. Geosites and Geotourism in the Local Development of Communities of the Andes Mountains. A Case Study. Sustainability 2021, 13, 4624. [Google Scholar] [CrossRef]
  96. 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. [Google Scholar] [CrossRef]
  97. Ayala Granda, A.J.; Carrión Mero, P.C.; Gurumendi Noriega, M.; Herrera Franco, G.; Morante Carballo, F.; Paz Salas, N.A. 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 Engineering, Integration, And Alliances for A Sustainable Development” “Hemispheric Cooperation for Competitiveness and Prosperity on A Knowledge-Based Economy, Proceedings of the 18th LACCEI International Multi-Conference for Engineering, Education, and Technology, Virtual, 27–31 July 2020; Latin American and Caribbean Consortium of Engineering Institutions: Buenos Aires, Argentina, 2020. [Google Scholar]
  98. Cruz, R.; Martínez-Graña, A.; Goy, J.L.; Nogueira, N. Analysis of the Geological Heritage and Geodiversity Index of Two Mountainous Areas in Spain: Béjar and El Barco Massifs. Geoheritage 2021, 13, 62. [Google Scholar] [CrossRef]
  99. Fuertes-Gutiérrez, I.; Fernández-Martínez, E. Mapping Geosites for Geoheritage Management: A Methodological Proposal for the Regional Park of Picos de Europa (León, Spain). Environ. Manag. 2012, 50, 789–806. [Google Scholar] [CrossRef]
  100. 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]
  101. Sellier, D. A Deductive Method for the Selection of Geomorphosites: Application to Mont Ventoux (Provence, France). Geoheritage 2016, 8, 15–29. [Google Scholar] [CrossRef]
  102. 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]
  103. Santos, D.S.; Mansur, K.L.; Seoane, J.C.S.; Mucivuna, V.C.; Reynard, E. Methodological Proposal for the Inventory and Assessment of Geomorphosites: An Integrated Approach focused on Territorial Management and Geoconservation. Environ. Manag. 2020, 66, 476–497. [Google Scholar] [CrossRef] [PubMed]
  104. Štrba, L.; Rybár, P.; Baláž, B.; Molokáč, M.; Hvizdák, L.; Kršák, B.; Lukáč, M.; Muchová, L.; Tometzová, D.; Ferenčíková, J. Geosite assessments: Comparison of methods and results. Curr. Issues Tour. 2015, 18, 496–510. [Google Scholar] [CrossRef]
  105. ASGMI Bases para el desarrollo común del Patrimonio Geológico en los Servicios Geológicos de Iberoamérica. Available online: https://asgmi.org/wp-content/uploads/2018/10/Bases-Metodologicas-Patrimonio-Geologico.pdf (accessed on 21 February 2022).
  106. Zouros, N.C. Geomorphosite assessment and management in protected areas of Greece Case study of the Lesvos island—Coastal geomorphosites. Geogr. Helv. 2007, 62, 169–180. [Google Scholar] [CrossRef]
  107. Herrera-Franco, G.; Montalván-Burbano, N.; Carrión-Mero, P.; Jaya-Montalvo, M.; Gurumendi-Noriega, M. Worldwide Research on Geoparks through Bibliometric Analysis. Sustainability 2021, 13, 1175. [Google Scholar] [CrossRef]
Figure 1. Location of Cajas National Park (CNP): (a) mainland Ecuador, with its three natural regions, from left to right: Coastal (blue), Andean (green) and Amazon (red). CNP location is highlighted in the box, inside the Azuay Province (grey); (b) local extension of CNP, within the parishes: Baños, Molleturo, San Joaquín and Sayausí.
Figure 1. Location of Cajas National Park (CNP): (a) mainland Ecuador, with its three natural regions, from left to right: Coastal (blue), Andean (green) and Amazon (red). CNP location is highlighted in the box, inside the Azuay Province (grey); (b) local extension of CNP, within the parishes: Baños, Molleturo, San Joaquín and Sayausí.
Sustainability 14 03120 g001
Figure 2. Geologic map of the study area. (a) Morphotectonic regions and major faults of mainland Ecuador (dotted lines), study area (red square). Modified from Aspden and Litherland [69], BGS-CODIGEM [70], Cediel et al. [71] and Jackson et al. [72]. Abbreviatures stand for: CFB: Coastal Foreland Basin, WC: Western Cordillera, IAD: Interandean Depression, EC: Eastern Cordillera, SAZ: Subandean zone and OB: Oriente Basin. Other features include: 1. Chaucha batholith, 2. Quimsacocha caldera; (b) geological units inside CNP, modified from Dunkley and Gaibor [62].
Figure 2. Geologic map of the study area. (a) Morphotectonic regions and major faults of mainland Ecuador (dotted lines), study area (red square). Modified from Aspden and Litherland [69], BGS-CODIGEM [70], Cediel et al. [71] and Jackson et al. [72]. Abbreviatures stand for: CFB: Coastal Foreland Basin, WC: Western Cordillera, IAD: Interandean Depression, EC: Eastern Cordillera, SAZ: Subandean zone and OB: Oriente Basin. Other features include: 1. Chaucha batholith, 2. Quimsacocha caldera; (b) geological units inside CNP, modified from Dunkley and Gaibor [62].
Sustainability 14 03120 g002
Figure 3. Simplified geomorphological map of Cajas National Park, modified from MAG-SENPLADES-IEE (2007) [78].
Figure 3. Simplified geomorphological map of Cajas National Park, modified from MAG-SENPLADES-IEE (2007) [78].
Sustainability 14 03120 g003
Figure 4. Methodology diagram applied in this study. Abbreviations: SGI, sites of geological interest.
Figure 4. Methodology diagram applied in this study. Abbreviations: SGI, sites of geological interest.
Sustainability 14 03120 g004
Figure 5. Geosite and geomorphosite examples of CNP according to Table 2: (a) Larga and Negra lakes, viewpoint (G2); (b) volcanic flow (G4); (c) Drumlins camp, viewpoint (G5); (d) Toreadora lake (G6); (e) Miguir meteorite (G7); (f) Llaviucu valley, viewpoint (G9); (g) camp of erratic boulders/Llaviucu river (G10).
Figure 5. Geosite and geomorphosite examples of CNP according to Table 2: (a) Larga and Negra lakes, viewpoint (G2); (b) volcanic flow (G4); (c) Drumlins camp, viewpoint (G5); (d) Toreadora lake (G6); (e) Miguir meteorite (G7); (f) Llaviucu valley, viewpoint (G9); (g) camp of erratic boulders/Llaviucu river (G10).
Sustainability 14 03120 g005
Figure 6. Geosite and geomorphosite map.
Figure 6. Geosite and geomorphosite map.
Sustainability 14 03120 g006
Figure 7. Assessment results according to the Brilha method: (a) accumulated evaluation criteria results over 14 geosites in three dimensions (S–scientific, UEP–use educational potential and PTU–potential tourism use); (b) degradation risk results for the 14 geosites.
Figure 7. Assessment results according to the Brilha method: (a) accumulated evaluation criteria results over 14 geosites in three dimensions (S–scientific, UEP–use educational potential and PTU–potential tourism use); (b) degradation risk results for the 14 geosites.
Sustainability 14 03120 g007
Figure 8. Classification of interests of the geosites: (a) scientific, (b) academic and (c) touristic.
Figure 8. Classification of interests of the geosites: (a) scientific, (b) academic and (c) touristic.
Sustainability 14 03120 g008
Figure 9. Assessment of degradation susceptibility (fragility and vulnerability due to anthropogenic threats) and protection priority of the geosites: (a) degree of susceptibility (DS); (b) global protection priority (Pp).
Figure 9. Assessment of degradation susceptibility (fragility and vulnerability due to anthropogenic threats) and protection priority of the geosites: (a) degree of susceptibility (DS); (b) global protection priority (Pp).
Sustainability 14 03120 g009
Table 1. Details of the evaluation criteria for each type of value (S, UEP, PTU and DR) with their respective scores and weight (constant values in %) (Brilha [32]). Interpretation: S, UEP and PTU classification, high (400–301), moderate (300–201), low (200–101) and very low (100).
Table 1. Details of the evaluation criteria for each type of value (S, UEP, PTU and DR) with their respective scores and weight (constant values in %) (Brilha [32]). Interpretation: S, UEP and PTU classification, high (400–301), moderate (300–201), low (200–101) and very low (100).
CriteriaValues
Scientific
(S)
Educational Potential (UEP)Potential
Tourism Use
(PTU)
Degradation
Risk
(DR)
R 1W 2R 1W 2R 1W 2R 1W 2
Representativeness1–4301–401–401–40
Key locality20000
Scientific knowledge5000
Integrity15000
Geological diversity51000
Rarity15000
Use limitations10550
Vulnerability010100
Accessibility0101015
Safety010100
Logistics0550
Density of population05510
Association with other values0550
Scenery05150
Uniqueness05100
Observation conditions01050
Didactic potential02000
Interpretative potential00100
Economic level0050
The proximity of recreational areas0050
Deterioration of geological elements00035
Proximity to areas/activities with potential to cause degradation00020
Legal protection00020
Total-100-100-100-100
1 R: score range; 2 W: weight.
Table 2. Degradation risk (DR) classification.
Table 2. Degradation risk (DR) classification.
Total (DR)Degree of Degradation
<200Low
201–300Moderate
301–400High
Table 3. Details of the evaluation parameters for each value type with their respective scores and weight (constant values in %) (García-Cortés and Carcavilla [40]). Interpretation: maximum (400), very high (267–400), high (134–266), medium (50–134) and low (<50).
Table 3. Details of the evaluation parameters for each value type with their respective scores and weight (constant values in %) (García-Cortés and Carcavilla [40]). Interpretation: maximum (400), very high (267–400), high (134–266), medium (50–134) and low (<50).
ParametersScore RangeValue (Weight)
Scientific (S)Academic
(A)
Touristic
(T)
Representativeness0 to 4305-
Standard of reference site105-
Knowledge of the site15--
State of conservation105-
Conditions of observation1055
Scarcity, rarity155-
Geological diversity1010-
Educational values-20-
Logistics infrastructure-155
Population density-55
Accessibility-1510
Size of site--15
Association with other natural elements-55
Beauty-520
Informative value--15
Possibility of recreational/leisure activities--5
Proximity to other places--5
Socio-economic situation--10
Total (weight) 100100100
Table 4. Parameters for calculating degradation susceptibility (DS). Interpretation of DS: maximum (400), very high (400–200), high (199–68), medium (67–13) and low (<13).
Table 4. Parameters for calculating degradation susceptibility (DS). Interpretation of DS: maximum (400), very high (400–200), high (199–68), medium (67–13) and low (<13).
ParameterFragility (F)Vulnerability (V)
ScoreWeightScoreWeight
Geosite size0 to 4400 to 4-
Vulnerability to looting30-
Natural hazards30-
Proximity of infrastructure-20
Mining exploitation interest-15
Protected area designation-15
Indirect protection-15
Accessibility-15
Ownership status-10
Population density-5
Proximity of recreational areas-5
Total (weight) 100 100
Table 5. Protection priority equations proposed by García Cortés and Carcavilla [40]. Interpretation of Pp: very high (400–113), high (112–17), medium (16–1) and low (<1).
Table 5. Protection priority equations proposed by García Cortés and Carcavilla [40]. Interpretation of Pp: very high (400–113), high (112–17), medium (16–1) and low (<1).
Protection PriorityEquation
Scientific Priority (S-Pp)   ( IS ) 2 ×   DS   × ( 1 / 400 2 )
Academic Priority (A-Pp)   ( IA ) 2 ×   DS   × ( 1 / 400 2 )
Tourist/recreational priority (T-Pp)   ( IT ) 2 ×   DS   × ( 1 / 400 2 ) ,
Protection Priority (Pp) ( IS   +   IA   +   IT 3 ) 2 ×   DS   × ( 1 / 400 2 )
Table 6. List of geosites and geomorphosites in the study area, typological, primary geological interest and main features.
Table 6. List of geosites and geomorphosites in the study area, typological, primary geological interest and main features.
No.NameType
of
Site
Type of Primary
Geological Interest
Type of Secondary Geological InterestMain Features
G1Tres Cruces peak ViewpointGeomorphologicalHydrological,
glacial
(695563 E; 9692860 N) It presents glacial valley views as well as rocky sub-vertical outcrops and highlights the highest elevations in the western part of the geosite, constituting the watershed between the Pacific and Atlantic hydrographic basins. Due to the wind currents crossing from both systems, there are landforms by wind erosion, such as the “Cerro Amarillo”. The erosion phenomena exposed peak of this tuff hill has a yellow colour that gives it its name. The site has an additional cultural/historical value due to the presence of “apachetas” (stone mounds that mark milestones along the route and are related to religious offerings).
G2Larga
and
Negra
lakes
ViewpointGlacialHydrological,
geomorphological
(695528 E; 9692725 N) The geosite shows late Pleistocene glacial modelling, such as glacial cirques, hanging valleys, ridges and several lakes. The staggered geomorphological arrangement of the lakes (paternoster lake) represents the erosive action of glacial advance and retreat that characterised the place. The most representative lakes are Larga lake (24 ha.) and Negra lake (2.5 ha.). In addition, there are several rocky outcrops, the product of natural erosion of the highest parts of the elevation where they are located. Due to the proximity of the lakes, the space constitutes a point of high biological diversity.
G3Rocky outcrop
Larga and
Negra
lakes
PointLithologicalHydrological,
geomorphological,
glacial
(695743 E; 9692562 N) It constitutes a vertical rocky outcrop on an arête that separates two valleys; one houses the Larga and Negra (West) lakes. According to Dunkley and Gaibor [62], the area is within the Chulo unit (Eocene), composed of sequences of lithologies of volcanic origin, such as rhyolitic tuffs and breccias.
G4Volcanic flowsPointLithologicalGlacial,
geomorphological,
hydrological
(695860 E; 9693813 N) It is in the highest areas (4102 m.a.s.l.) of CNP. An outcrop is presented, showing a volcanic flow of andesitic tuffs. In the surroundings, there are rocks with a marked andesitic composition. In addition, there are panoramic views of lakes of glacial origin (e.g., Palcacocha), hanging valleys and roches moutonnées.
G5Drumlins campViewpointGlacialHydrological,
geomorphological,
landscape
(698949 E; 9691621 N) The geosite located at the foot of the main road offers a panoramic view of an extensive field of asymmetric mounds of drumlins-type landforms. In addition, there are paternoster type lakes and fluvial slopes to the west. Finally, at the bottom of the valley there are examples of roches moutonnées.
G6Toreadora
lake
PointGlacialHydrological,
erosional
(697359 E; 9692361 N) It is in the central part of the main tourist attractions of the park, consisting of a 19-hectare lake of glacial origin. You can see the peaks of Cerro San Luis (4295 m.a.s.l.), one of the highest elevations in CNP. The area has a natural interest due to the present examples of the endemic flora and fauna of the Andean Paramo. In addition, it presents a recognised regional historical value due to the García Moreno trail, one of the first modern transport routes between the Coastal and Andean regions.
G7Miguir
meteorite
PointLithologicalHydrological,
geomorphological
(698370 E; 9692187 N) The geosite is an example of impact metamorphism; it constitutes a block of andesite metamorphosed by the impact caused by a meteorite fragment a few centimetres in size, which occurred in 1995 [87]. The block physically presents radial fissures around a 7–10 cm diameter crater.
G8Tomebamba hanged
valley and glacial
groove zone
ViewpointGlacialFluvial,
geomorphological
(703276 E; 9692966 N) It is located in a section of the main road, where there is evidence of the glacier’s passage through grooves generated by the transport of an ice mass on the slopes of a hill located to the east, taking into account the direction of Cuenca–CNP. In addition, there is one of the best examples of hanging valleys in the southern part of the park.
G9Llaviucu
valley
AreaGlacialFluvial,
geomorphological
(708968 E; 9685563 N) The geosite has a characteristic parabolic geometry or “U-shaped” valley. In addition, fluvial modelling can be observed, caused by the rivers that flow from the lakes located at higher altitudes.
G10Camp of
erratic
boulders/
Llaviucu
river
AreaGlacialFluvial,
geomorphological,
lithological
(708089 E; 9685467 N) The site presents a field of erratic blocks arranged on the flood plain formed by the Llaviucu river, which comes from the lake of the same name. In addition, there is part of the meandering system of the Llaviucu river.
G11Llaviucu
lake
AreaGlacialHydrological,
geomorphological
(706065 N; 9685420 N) Lake of glacial origin located at the end of the valley of the same name. The chain of lakes found at higher altitudes (e.g., Taitachugo and Osohuaycu) supply Llaviucu lake. It presents on the sides of the valley, where vertical rocky outcrops are a product of the last stages of glacial erosion in the area.
G12End part
of glacial
valley
Llaviucu
AreaGlacialGeomorphological(705571 E; 9685621 N) The geosite is a representative example of a terminal moraine, which signifies the maximum glacier passage. Additionally, a segment of a meandering stream of high sinuosity is the product of a postglacial stage. In addition to G9 and G11, this site is part of the ancient Inca Trail, which connected several places across South America in pre-Hispanic times.
G13Burines
route
peaks
ViewpointGlacialGeomorphological,
hydrological
(698307 E; 9691400 N) Geosite constitutes eutrophic lakes, which connect by small streams with a smaller lake at a lower elevation than the previous one. There are also views of hanging valleys, peaks and a small valley with a stationary lake.
G14Burines
route
periglacial
AreaPeriglacialGeomorphological,
glacial,
hydrological,
erosional
(698054 E; 9691342 N) Periglacial landforms with swampy plains, which in some sections are confused with the course of the creeks that come from the high-altitude elevations surrounding them. Likewise, it presents a landslide of about 20 to 30 m in height, extending vertically from the small glacial valley base. It exhibits examples of periglacial morphogenesis where there is highlighted hummock.
Table 7. Assessment of geosites is based on four dimensions: scientific (S), use educational potential (UEP), potential tourism use (PTU) and degradation risk (DR).
Table 7. Assessment of geosites is based on four dimensions: scientific (S), use educational potential (UEP), potential tourism use (PTU) and degradation risk (DR).
No.GeositesSUEPPTUDRDegradation Risk Category
G1Tres Cruces peak195290315180Low
G2Larga and Negra lakes300330315215Moderate
G3Rocky volcanic outcrop185250285200Moderate
G4Volcanic flows230210220170Low
G5Drumlins camp255280315130Low
G6Toreadora lake215330315215Moderate
G7Miguir meteorite315215240230Moderate
G8Tomebamba hanged valley and glacial grooves zone260280245150Low
G9Llaviucu valley320320285150Low
G10Camp of erratic boulders/Llaviucu river165225240250Moderate
G11Llaviucu lake320315310210Moderate
G12Llaviucu end moraine320330340250Moderate
G13Burines peaks210210255210Moderate
G14Burines periglacial plains210220255135Low
Table 8. Assessment of geosites in terms of scientific (S), academic (A), touristic (T) and average (Av.) interest, susceptibility to degradation (DS), vulnerability due to anthropic threats (Vul.), and scientific (S-Pp), academic (A-Pp), touristic (T-Pp) and global (Pp) protection priority.
Table 8. Assessment of geosites in terms of scientific (S), academic (A), touristic (T) and average (Av.) interest, susceptibility to degradation (DS), vulnerability due to anthropic threats (Vul.), and scientific (S-Pp), academic (A-Pp), touristic (T-Pp) and global (Pp) protection priority.
No.GeositesSATAv.DSVulS-PpA-PpT-PpPp
G1Tres Cruces peak235280330281.6712.001604.145.888.165.95
G2Larga and Negra lakes255240305266.6727.7518511.281016.1312.33
G3Rocky volcanic outcrop190210190196.6798.0014022.1227.0122.1123.69
G4Volcanic flows295220250255.0024.0016013.057.269.389.75
G5Drumlins camp305255310290.0019.5013011.347.9211.7110.25
G6Toreadora lake260270290273.3342.5017017.9619.3622.3419.85
G7Miguir meteorite370270180273.3384.0012071.8738.2717.0139.22
G8Tomebamba hanged valley and glacial grooves zone295255305285.0042.0010522.8417.0724.4221.32
G9Llaviucu valley335275320310.0023.2515516.3110.9914.8813.96
G10Camp of erratic boulders/
Llaviucu river
235245300260.0032.5013011.2212.1918.2813.73
G11Llaviucu lake335285265295.0033.7513523.6717.1314.8118.36
G12Llaviucu end moraine245225255241.6770.0014026.2622.1528.4525.56
G13Burines peaks180230260223.3339.001307.9012.8916.4812.16
G14Burines Periglacial plains295225245255.0042.508523.1213.4515.9417.27
Table 9. Strengths, weaknesses, opportunities and threats (SWOT) matrix.
Table 9. Strengths, weaknesses, opportunities and threats (SWOT) matrix.
Internal
Environment
Strengths (S)Weaknesses (W)
External
Environment
S1. A great diversity of geological elements of glacial origin.
S2. It has a park conservation plan based on five major programs that focus on administration and planning, control and surveillance, communication, environmental education and participation, public use and tourism and biodiversity management.
S3. It has a system of tourist services, marked trails and organised recreation and leisure activities.
S4. Infrastructure dedicated to the promotion of the natural environment of the park.
S5. The park has a status of protection under the national system of protected areas, and its area has two international denominations: a Ramsar site and a biosphere reserve.
S6. High cultural and ecological value.
W1. Scarce information and promotion about the park’s geoheritage.
W2. The trail system and other tourist facilities do not cover all sites of geological interest.
W3. Absence of plans dedicated to geoconservation.
W4. Sometimes the base information of the park may be outdated.
W5. There is little interaction on social media.
W6. Absence of scientific–informative studies and publications on the geological characterisation of the park.
W7. Extensive pine plantations hinder the scenic value of the sites.
W8. Limited community participation.
W9. Lack of interaction, protocol and communication on the logistics processes for the geoscientific research of the park.
Opportunities (O)Strategies: S + OStrategies: W + O
O1. Creation of geotourism itineraries.
O2. Generation of geoproducts and activities related to the park’s geoheritage for local development.
O3. Development of awareness of the value of geoheritage in conjunction with biodiversity values in local communities and tourists.
O4. Increase in the tourist value of the park.
S1. O1. O2. Develop a geotourism approach related to glacial environments that promotes education and research.
S1. O1. O3. Establish a national and international technical–scientific cooperation network for the sustainable development of the park.
S3. O1. O2. Articulate tourist routes and other services offered by the park with geological characteristics, highlighting the value of the geological heritage.
W4.O3. Update the general information exhibited in the park’s tourist infrastructure and develop and install interpretive panels of the geological–geomorphological processes of the place.
W6.O4. Promote geoscientific research on geoheritage issues to add more sites of geological interest to the list of geosites–geomorphosites proposed in this study.
W6.O1. Design an inventory and characterisation protocol for geosites compatible with the natural heritage inventory process.
W7.O1. O4. Generate a management plan to prohibit the advance of pine plantations that hinder the landscape value of geosites.
Threats (T)Strategies: S + TStrategies: W + T
T1. Greater park promotion will increase the influx of tourists and hurt geosites and the environment.
T2. Alteration in certain park areas due to human activities (e.g., cattle, agriculture, mining and wildfires).
T3. Natural threats due to erosion and landslides.
S2.T2. Establish awareness campaigns on the conservation of geoheritage for the local community and seminars on this subject in the current interpretation centres.W3.T1. T3. Generate and implement a comprehensive geoconservation plan for the natural park.
W5.T1. Design virtual guides on GIS 1-based websites compatible with mobile applications that facilitate geological–geomorphological interpretation and education on geosite conservation measures.
1 GIS: geographical information system.
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Navarrete, E.; Morante-Carballo, F.; Dueñas-Tovar, J.; Carrión-Mero, P.; Jaya-Montalvo, M.; Berrezueta, E. Assessment of Geosites within a Natural Protected Area: A Case Study of Cajas National Park. Sustainability 2022, 14, 3120. https://doi.org/10.3390/su14053120

AMA Style

Navarrete E, Morante-Carballo F, Dueñas-Tovar J, Carrión-Mero P, Jaya-Montalvo M, Berrezueta E. Assessment of Geosites within a Natural Protected Area: A Case Study of Cajas National Park. Sustainability. 2022; 14(5):3120. https://doi.org/10.3390/su14053120

Chicago/Turabian Style

Navarrete, Edison, Fernando Morante-Carballo, Jairo Dueñas-Tovar, Paúl Carrión-Mero, María Jaya-Montalvo, and Edgar Berrezueta. 2022. "Assessment of Geosites within a Natural Protected Area: A Case Study of Cajas National Park" Sustainability 14, no. 5: 3120. https://doi.org/10.3390/su14053120

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