Next Article in Journal
Development and Implementation of a Novel Optimization Algorithm for Reliable and Economic Grid-Independent Hybrid Power System
Next Article in Special Issue
Assessment of Bacterial Contamination of Air at the Museum of King John III’s Palace at Wilanow (Warsaw, Poland): Selection of an Optimal Growth Medium for Analyzing Airborne Bacteria Diversity
Previous Article in Journal
ERS-1/2 and Sentinel-1 SAR Data Mining for Flood Hazard and Risk Assessment in Lima, Peru
Previous Article in Special Issue
Biocidal Activity of Phyto-Derivative Products Used on Phototrophic Biofilms Growing on Stone Surfaces of the Domus Aurea in Rome (Italy)
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Frequent Microalgae in the Fountains of the Alhambra and Generalife: Identification and Creation of a Culture Collection

by
Fernando Bolívar-Galiano
1,*,
Clara Abad-Ruiz
1,
Pedro Sánchez-Castillo
2,
Maurizio Toscano
1,* and
Julio Romero-Noguera
3
1
Department of Painting, Faculty of Fine Arts (Facultad de Bellas Artes), University of Granada, 18014 Granada, Spain
2
Department of Botanic, Faculty of Sciences (Facultad de Ciencias), University of Granada, 18071 Granada, Spain
3
Department of Painting, Faculty of Fine Arts (Facultad de Bellas Artes), University of Seville, 41003 Seville, Spain
*
Authors to whom correspondence should be addressed.
Appl. Sci. 2020, 10(18), 6603; https://doi.org/10.3390/app10186603
Submission received: 4 August 2020 / Revised: 10 September 2020 / Accepted: 17 September 2020 / Published: 21 September 2020
(This article belongs to the Special Issue Microbial Communities in Cultural Heritage and Their Control)

Abstract

:
Cyanobacteria, green algae and diatoms are significant factors in the biodeterioration of stone cultural heritage sites, and specifically fountain monuments, due to the constant presence of water. In this study, samples were taken from different fountains in the Alhambra and Generalife, which are among the Spanish monuments of greatest historical and artistic value and which together were declared a World Heritage Site by UNESCO in 1984. The aim was to identify which species of colonising microalgae are most frequent and to obtain monoalgal cultures from them. From a conservation point of view, it is interesting to identify which algae are growing in these fountains and how they behave in order to develop new methods to control their growth. The most abundant groups of algae in our samples were green algae and cyanobacteria. The most common genera in the former group were Bracteacoccus, Chlorosarcina, Chlorosarcinopsis, Apatococcus and Klebsormidium. As for cyanobacteria, the most abundant genera were Phormidium, Calothrix, Leptolyngbya, Chamaesiphon, Pleurocapsa and Chlorogloea. Using our collected samples, 10 genera of green algae and 13 genera of cyanobacteria were isolated, thereby constituting the base samples for the creation of a reference collection of living algae from the Alhambra and Generalife contexts, which can be used in subsequent studies to develop new types of treatment against biodeterioration.

1. Introduction

Biodeterioration is defined as “any undesirable change in the properties of a material caused by the vital activities of organisms” [1]. Microorganisms such as bacteria, fungi and algae are some of the agents that affect cultural heritage. These include microalgae (microscopic algae), which are photoautotrophic organisms capable of growing in media not abundant in nutrients, such as stone, and which proliferate in places with water or high relative humidity levels. In particular, they play an especially relevant role in fountains and ornamental pools, because of the constant presence of water. These organisms are the main agents involved in the biodeterioration of such cultural assets, causing a range of aesthetic, physical and chemical damage, which has been studied over time around the world [2,3,4,5,6,7,8].

1.1. Biodeterioration Caused by Microalgae

Algae are a polyphyletic group of organisms, i.e., not all of them come from the same evolutionary ancestor. Within this group we can find enormous diversity in all aspects. For decades, cyanobacteria were called blue-green algae and were grouped with other algae, mainly because of their ability to photosynthesise. Today, however, cyanobacteria are no longer considered to be algae, but rather are considered to be prokaryotic organisms belonging to the domain of bacteria, whereas all algae are eukaryotic organisms. Nevertheless, in this work we have included cyanobacteria in the term “microalgae”, since both cyanobacteria and eukaryotic microalgae share important similitudes from an ecological point of view—they both carry out oxygenic photosynthesis, are pioneers in stone colonisation and cause biodeterioration through similar processes.
The microalgae that affect the state of conservation of fountains and other stone monuments can be epilithic (growing on the surface of the stone) or endolithic (growing inside it) [9], with the latter being the most damaging and difficult to remove. In any case, the microalgae grow and produce biofilms, which are composed of a set of organisms and the substances they excrete. The microalgae are often embedded in the matrix formed by extracellular polymeric substances (EPS) secreted by the cells themselves. These biofilms affect the stone in different ways.
Firstly, they produce an aesthetic alteration, since they form a patina with different natural colourings (light or dark green, orange, brown, etc.) on the stone. Furthermore, biofilms can retain atmospheric dust, pollutants from the air and substances dissolved in water that give them a darker colour. This phenomenon is known as “soiling” [10].
These biofilms also cause physical deterioration. Many species of algae and cyanobacteria have sheaths that produce mucilage, which allows them to adhere to the substrate. The mucilage is also used for gliding [11] and to retain water, so that they can survive drought conditions. On absorbing water, the sheaths’ volumes vary considerably and their continual expansion and contraction causes the gradual destruction of stony material, which progressively loses granules that go on to form part of the biofilm [12]. The water retention is a fundamental aspect—it fosters mechanical damage due to frost weathering with sharp temperature changes and also allows for the growth of other algae with greater water requirements, so that the communities become more complex. It may also help to support other, more damaging organisms such as fungi and bryophytes [10]. If there is an excessive proliferation of algae, they can affect how the fountains work, for example by blocking spouts, which can speed up other kinds of deterioration.
Lastly, microalgae also cause chemical damage as a result of their metabolism. They directly damage the stone by producing organic acids (pyruvic, glycolic, lactic, acetic, succinic or alpha-ketoglutaric) as a result of respiration. They can also give rise to biotransference phenomena by producing polysaccharides capable of reacting with metals and leading to salts [10]. However, the main damage caused by microalgae is the formation of carbonate crusts on stone as a result of photosynthetic metabolism. Due to the withdrawal of carbon dioxide from the medium and to other biochemical processes, the pH increases and fosters the precipitation and aggregation of calcium carbonate, forming layers of mineral crusts that can end up covering the original materials. All of the taxonomical groups of microalgae are capable of producing this phenomenon, but it is more apparent in green algae and cyanobacteria [13]. In addition to causing aesthetic damage, these crusts foster physical biodeterioration due to thermal contraction and expansion to a different extent than the substrate, leading to stresses that weaken it, and also due to frost weathering, since they have a porous structure that increases water retention further [14]. They even end up burying the algae that have generated them, which then become endolithic and protected from environmental adversity and treatments, such as cleaning or biocides, thereby making it much more difficult to remove them.

1.2. Microalgae in the Alhambra

The Alhambra and the Generalife, together named a UNESCO World Heritage site in 1984, form one of the monument complexes of greatest historical and artistic value in Spain, meaning their conservation is of great importance. Given the importance of water in Nasrid and Islamic art in general, “alberca” reservoirs abound throughout the site, together with channels and fountains (including iconic ones such the Fountain of the Lions), which are affected by colonisation from microalgae (Figure 1a,b). Although the fountains undergo continual maintenance, these microorganisms are not completely eliminated. Despite the continual investment in chemical and mechanical means of control, the natural process of colonisation starts over.
From a conservation point of view, an important step will be to learn the structure and functioning of the biofilms—which species they are composed of, which species are the most resistant or harmful and which conditions help them develop—in order to know how to create new methods in order to halt the proliferation of the microorganisms. There are a variety of previous studies on the biodeterioration of fountains in the Alhambra, many of which focus on biodeterioration due to microalgae, their classification and possible treatments [13,14,15,16,17,18]. Our work was intended to update the knowledge on the types of microalgae growing in different fountains throughout the Alhambra and Generalife today, establishing which genera are the most common and beginning to create a culture collection of algae with species from these sites.

1.3. Objectives

The main objectives of this study were as follows:
To identify the most common types of microalgae in the fountains of the Alhambra and Generalife today.
To make unialgal cultures of the species found in the Alhambra in order to begin creating a culture collection of living microalgae available for subsequent studies on new kinds of treatment to keep biodeterioration under control.
To design a data model and develop a database for storing and retrieving sample details and analytical values from the collection.

2. Materials and Methods

2.1. Sampling

In total, 21 fountains in the Alhambra and Generalife were studied, with 120 microalgae samples being taken (Table 1, Figure 2 and Figure 3). The greatest number of samples was taken from the monumental fountains of the Nasrid Palaces (60%) due to their enormous cultural interest, although fountains from other areas were also selected in order to make a general assessment of the most common microalgae species colonising the monument complex.
The main materials used in the construction of the fountains studied are white marble from Macael (Almería, Spain) and limestone from Sierra Elvira (Granada, Spain). Samples were also taken from the surface of the metal spouts in several fountains, and occasionally from mortar, brick, plaster and glazed tile.
Seasonal samples were taken in spring, summer, autumn and winter throughout 2017 and 2018. The sampling points chosen were the surfaces of fountains, where generally abundant macroscopic growth of epilithic microalgae had been observed (Figure 4a and Figure 5a). For each of them, the appearance of the microalgal communities was taken into account, as well as the environment they belong to, according to the water available—submerged, amphibious or aerial. The submerged zones are those below the water level, as well as those exposed to continually running water, for example inside the spouts or the outer surfaces of basins where the water overflows. The amphibious zones are those with water available intermittently, such as surfaces that are splashed by the spouts. Lastly, we considered aerial zones to be those that do not receive water directly, such as the inner surfaces of basins above the waterline. As for the appearance of the communities, we can distinguish: patinas (sheens) when they are of scant thickness and less structural complexity; mats of greater thickness and complexity with filamentous species; and mineral crusts, which may arise due to any of the aforementioned formations. On taking the different samples, a variety of fountains and environments were chosen to take into account the greatest possible variation. In some cases, samples were also taken in the same zone at different times of year to check if there were variations in the types of species. All of the sampling processes were photographically documented (i.e., fountain, sampling point, sample, etc.), giving a total of around 450 high quality images.

2.2. Procedures

The samples were taken with the help of tweezers and a scalpel, taking the utmost care not to damage the original materials (Figure 4a and Figure 5b). The fresh samples were stored in a polyethylene sample collection vial with water from the same fountains from which they were taken (Figure 5c). In the laboratory, they were refrigerated until their observation. After observation and creation of the cultures, the samples were fixed in a glutaraldehyde and 25% glycerol solution (Sigma Aldrich, Madrid, Spain) to be permanently conserved for subsequent study.
The microalgae present in each sample were identified in the phycology lab at the University of Granada, using a binocular Olympus SZX10 microscope and an inverted optical Nikon Eclipse TS100 microscope. The following references were used for the taxonomical identification [19,20,21,22,23]. In each sample, the different microalgal species were identified according to their proportion by assigning an index (+–5) representing the following percentages of the total: + = 0–1%; 1 = 1–10%; 2 = 10–25%; 3 = 25–50%; 4 = 50–75%; 5 = 75–100%.
Microalgal cultures were made in a solid medium in agar plates at 2%. A high concentration of agar (Sigma-Aldrich, Madrid, Spain) was used because we intended to cultivate benthonic species that grow on stone. The media for cultivation used were Bold’s Basal Medium (BBM) (Phyto-Tech, Lenexa, KS, USA) for green algae and BG11 (Sigma-Aldrich, Madrid, Spain) for the cyanobacteria. Diatoms were not cultivated since they have stricter nutritional requirements and are the most uncommon group of the three.
The streaking method was used to isolate microalgae on plates using fresh samples, successive isolations and re-streaking until unialgal cultures were obtained. The plates were incubated at 22 °C with a cycle of 12 h of light/darkness using fluorescent lamps that provided a light intensity of 800 lux. On average, the cultures took about two weeks to develop, but the time varied depending on multiple factors, so the re-streaking was carried out depending on the needs in each case.

2.3. The Database

From the planning phase of this study, it was clear that the information collected would have to be processed digitally in order to provide the required features of data consistency, automated backups, multimedia, versioning, automatic quantifications and simultaneous user access. For this reason, two of the fundamental steps towards the creation of the algae culture collection were the design of a specific data model (tables, attributes, constraints and relationships) and creation of a multimedia, geographically enabled database to store and retrieve data collected during the sampling process and in the lab.
The main tables of the database are used to register the collected samples, the monuments (fountains, pools, etc.) and the identified microalgae; the sample represent the central table, which is related to both the monuments (one-to-many cardinality) and microalgae (many-to-many cardinality). Several attributes were added in order to register a wide variety of values, such as the date, location, environment, media and description; classes, genera, species, colour, communities, biotype and biotope; applied treatments, cultivation procedures and microphotographs; material and chronology.
The database is simultaneously accessible by several users with individual accounts and different roles. Stored data can be retrieved using a unified search box or through a multifaceted search engine, supporting filters and Boolean operators. Every change in each record is registered in a different version, so individual states can be restored or compared between each other. The database’s architecture is modular so that additional workflows can be introduced in the future, for example to continue tracing the evolution of biodeterioration through time and space.

3. Results and Discussion

3.1. Microalgae Identification

Firstly, it should be noted that the fountains in the Alhambra and Generalife undergo continual maintenance. In addition to general cleaning, all of the studied fountains are treated with algicide. Chlorine tablets are also added to many of them. In the specific case of the Fountain of the Lions, the water flowing through gets special treatment, since the circulating water is chlorinated and purified independently from the rest of the fountains of the monumental complex. This means that the algae colonising these fountains exist in conditions that are far from natural and that they are generally resistant to such treatment, since it does not manage to eliminate them completely.
Fundamentally, as seen in previous studies [15,16], the microalgae colonising the selected fountains largely belong to the large groups of green algae (Chlorophyta and Charophyta), cyanobacteria (Cyanobacteria) and diatoms (Bacillariophyta). Once in a while, some Euglenophyceae have been found in the “guitar” fountain of the Ladies Tower (Torre de la Damas, Granada, Spain), and a Phaeophyceae (Pleurocladia sp.) sample was found in the southern “guitar” fountain of the Court of the Myrtles (Patio de los Arrayanes), but those species were only seen in spring in those two fountains. Table 2 shows a list of the fountains studied and the most common genera of microalgae in each.
In general, diatoms are the least common type of microalgae of the three groups studied. The most common genera are Navicula, Nitzschia, Cymbella and Achnanthes, all of which are unicellular; in other words, the cells appear individually and do not form aggregates adhering strongly to surfaces. Of these four genera, Navicula, Nitzschia and Cymbella were the ones that generally appeared most often. All of these diatoms appear to a greater or lesser extent as part of more complex communities of cyanobacteria or green algae, except one specific case in which a monospecific community of Achnanthes sp. was found on the surface of the basin at the Fountain of the Lions in a submerged environment.
Among the green algae, the unicellular coccoid algae were very widespread in many fountains. The most common genera were Chlorococcum, Bracteacoccus, Chlorosarcina and Chlorosarcinopsis (Figure 6a). The latter two are able to form generally compact cellular aggregates; less frequent (but still common) were the Chlorella and Choricystis genera.
In addition to the coccoid algae, we also found pseudoparenchymatic green algae—Pleurastrum sp. appears on the “guitar” of the Ladies Tower and Gongrosira sp. appears in the pool of the central fountain of the Court of the Sultana, both in submerged zones; while Leptosira sp. and Pseudopleurococcus sp. grow in the Water Stairway’s handrail in amphibious and submerged zones. As for filamentous algae, the Klebsormidium genus is common, which is also capable of growing in aerial zones. The filamentous alga Cladophora sp. has also been found sporadically (in the Lindaraja fountain and the mirador del Partal basin), as has Stigeoclonium sp. (Patio de los Arrayanes southern “guitar” fountain). Stichococcus sp. appears in amphibious and aerial zones of the Fountain of the Lions and in the handrail of the Escalera del Agua, although not abundantly. This alga is normally unicellular, although it can form lax filaments that disaggregate easily. Other green algae genera that are very widespread, although their proportions in the samples are usually small, were Scenedesmus sp., which is characteristic for the formation of coenocytes; Cosmarium sp. (unicellular; Figure 6b); and Apatococcus sp. (pseudoparenchymatic).
As for the cyanobacteria, we can distinguish between the Chlorococcales order, whose cells group together in colonies (and are occasionally filamentous in appearance), and truly filamentous algae. The most common chlorococcal cyanobacteria among all the fountains studied were Pleurocapsa sp. (Figure 7a), Chamaesiphon sp. (Figure 7b) and Chlorogloea sp., followed by Chroococcidiopsis sp., which is less common or abundant. Cyanosarcina sp. is very common in amphibious zones in the Fountain of Lions and the fountain of the Glorieta del Secano, while Myxosarcina sp. grows in crusts on the southern “guitar” of the Patio de los Arrayanes and the Escalera del Agua. On the Fountain of the Lions and the basin of the Puerta de las Granadas there were species of the genus Chroococcopsis. Among the filamentous cyanobacteria, the most common genera were Phormidium (Figure 8a,b), Calothrix, Leptolyngbya, Lyngbya and Schizothrix. The former three appear in a large number of fountains and are relatively abundant. Symploca sp. appears abundantly in submerged zones of the “guitar” of the Torre de las Damas and in the fountain of the Patio de la Reja. In the spout of the northern fountain in the Patio de los Arrayanes, where samples have been taken in different seasons, a dark mat forms in which there is a great presence of Dichothrix sp.
Normally, in each sample we can find generally complex communities made up of two or more different species of microalgae. However, occasionally the sheens (structures of less complexity) are composed of a single species. In the vast majority of cases, these species belong to the unicellular coccoid green algae group, such as Chlorosarcinopsis sp. (Figure 6a) and Bracteacoccus sp. (Figure 8b and Figure 9a), which form green sheens in submerged environments. On the one hand, this could show that these genera are especially resistant in adverse conditions and capable of surviving in environments with chlorine and biocides; on the other hand, this could indicate that they are the first colonisers of the stone. Mats are usually found in submerged and amphibious environments and are typical of the presence of filamentous algae, especially cyanobacteria. Phormidium, Leptolyngbya (Figure 9c,d) and Calothrix are genera that are often the majority species in mats. On some fountains, these mats also have a high proportion of filamentous green algae, such as Gongrosira sp. and Klebsormidium sp. (Figure 9b). As for the mineral crusts, we found variability; they are usually made up of complex communities of various species of green algae and cyanobacteria (which sometimes form separate strata), but sometimes they are monospecific, made up of Chlorosarcina sp. and other indeterminate coccoid green algae.
Most of the samples belong to amphibious or submerged environments. Only a few were taken in aerial zones far from water, although sometimes with high levels of relative humidity. In such zones, genera such as Leptolyngbya, Klebsormidium, Stichococcus, Chlorococcum, Bracteacoccus, Cosmarium, Phormidium, Navicula and Chroococcopsis were observed, forming coloured sheens and sometimes crusts. These same genera appear in other submerged and amphibious zones, so they are not exclusive to aerial environments. In the sampling points where samples were taken in different seasons, it was observed that there may be small variations between the observed species of algae, however the general composition is largely maintained and the predominant algae are usually the same. The small variations may be due to seasonal environmental factors or cleaning and maintenance work on the fountains.
If we compare the results with previous studies carried out on the fountains of the Alhambra [14,15,17,18], only the genera Bracteacoccus, Klebsormidium, Choricystis, Stichococcus, Chroococcopsis, Cyanosarcina and Dichothrix had never been recorded before. The rest of the observed microalgae had already been recorded in the past. It could be considered that new genera have appeared, but it has to be remembered that the same fountains have not always been sampled. In the thesis by Bolívar-Galiano [15] only seven fountains in the Nasrid Palaces were sampled, while in the study by Cuzman et al. [18] only samples from the Lindaraja fountain and the central fountain of the Patio de la Sultana were studied. In addition to some of the genera currently identified, in the previous works there are also some different ones. Specifically, we can observe a much greater number of species in the thesis by Bolívar-Galiano in 1994 [15], even though only the Nasrid Palaces’ fountains were considered. One of the explanations for this phenomenon is that the presence of microalgae in these fountains was much greater 26 years ago, because today much more attention is paid to these microorganisms and the fountains undergo much stricter maintenance (Figure 10).
The genera we have identified are also common in other monumental fountains in Spain and Italy, where the following genera have been found: Phormidium, Calothrix, Leptolyngbya, Lyngbya, Symploca, Schizothrix, Myxosarcina, Chamaesiphon, Chroococcidiopsis, Pleurocapsa, Chlorogloea, Pseudopleurococcus, Pleurastrum, Apatococcus, Chlorella, Chlorosarcina, Chlorosarcinopsis, Cosmarium, Chlorococcum, Scenedesmus, Stichococcus, Navicula, Nitzschia, Achnanthes and Cymbella [18,24,25]. In addition to fountains, microalgae are also pioneers in colonising other stone monuments and are significant agents in their biodeterioration. In fact, algae and cyanobacteria are the majority components in the biofilms found in stone buildings in Europe, while cyanobacteria are the majority components in those found in Latin America [26]. In a study carried out on the most common green algae and cyanobacteria in monuments of different types of stone in the entire Mediterranean basin [27], all of the genera we found in the Alhambra were present. In fact, Klebsormidium sp., which did not appear in the fountains of the Alhambra in the past, is among the most common species in stone monuments. All of this information makes it clear that the microalgae growing in the Alhambra and Generalife are species that are very widespread and which are commonly found in stone monuments.

3.2. Obtaining Unialgal Cultures

Using some of the fresh samples, and before fixing them permanently, agar plates were inoculated and cultures were made in them to isolate the greatest possible number of microalgae species (Figure 11). The media used to isolate green algae were BBM (Phyto-Tech, Lenexa, KS, USA), while BG11 (Sigma-Aldrich, Madrid, Spain) media were used for cyanobacteria. From the beginning, the cultivation of diatoms was discarded, since they have greater nutritional demands and are a less common group. Finally, it was possible to isolate a total of 23 genera—10 green algae and 13 cyanobacteria (Table 3).
The set of data is not completely representative of the site’s microbiota, since it is known that the different species have different capacities for developing within in vitro cultures. Some species do not prosper in cultivation, even if they are predominant in the sample. Likewise, species that are not at all abundant in the samples may grow when cultivated, including species that are not observed in the sample, as has occurred with Neochloris sp., Pseudanabaena sp., some species of Nostoc sp., Gloeocystis sp. and Leptolyngbya sp. The latter genus grows very well when cultivated (Figure 9c,d), with up to 4 different strains having been isolated that are unidentified at the species level. This shows that there is a great diversity within this genus in the Alhambra and Generalife study areas. Moreover, in communities with great diversity it can be difficult to isolate different algae, especially filamentous cyanobacteria that form mats.
It is useful to have isolated cultures of species, firstly for identification, since the microalgae’s life cycles can be observed, as well as for other different characteristics that are often difficult to appreciate in samples. Furthermore, through this process a culture collection of living algae has been initiated that may be used in future studies. All of the microalgae from the Alhambra and Generalife are to a greater or lesser extent resistant to the treatments currently used to eliminate them (for example, many are resistant to chlorine), which makes them very interesting for use in trials with new treatments aiming to halt the proliferation of these organisms in research projects currently in progress.

4. Conclusions

The fountains of the Alhambra and the Generalife are affected mainly by the growth of green algae, cyanobacteria and diatoms, which grow in spite of the physical and biocide treatments that are periodically carried out. Within these groups, the green algae are the most abundant, as is the case in many other examples of cultural heritage sites involving water and ornamental stone. In most cases different algae grow, forming complex communities and giving rise to sheens, mats or crusts. Coccoid unicellular green algae such as Chlorococcum sp., Bracteacoccus sp., Chlorosarcina sp. and Chlorosarcinopsis sp. are especially common, which sometimes form monospecific sheens. Other common genera of green algae are Klebsormidium, Scenedesmus, Chlorella, Choricystis, Pleurastrum, Gongrosira, Leptosira and Pseudopleurococcus. As for cyanobacteria, the most notable genera are Pleurocapsa, Chamaesiphon, Chlorogloea, Chroococcidiopsis, Cyanosarcina, Myxosarcina and Chroococcopsis (chroococcal); and Phormidium, Calothrix, Leptolyngbya, Lyngbya and Schizothrix (filamentous). The species of microalgae found here hardly vary with seasonal changes, and the most common microalgae are always the same ones. All of the genera found in this study correspond to widespread algae that are common in the colonisation of monumental fountains and stone cultural assets in general.
Ten green algae genera and thirteen cyanobacteria genera have been isolated in cultures (with different species of some of them), thereby initiating the creation of a collection of algal cultures from the Alhambra and Generalife. Although the identification of each species among them is still in progress, this will gradually be carried out as new observations are made in the cultures. This collection may be used in the future to carry out trials of new treatments or to study the synthesis of certain products using microalgae in the study area.

Author Contributions

Conceptualization and work design, F.B.-G. and P.S.-C.; methodology and analyses, F.B.-G, C.A.-R. and P.S.-C.; software, M.T.; writing—original draft preparation, F.B.-G., C.A.-R. and J.R.-N.; writing—review and editing, J.R.-N., M.T. and P.S.-C.; data curation, C.A.-R. and M.T.; project administration, funding acquisition, F.B.-G and M.T. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by the following projects: “Desarrollo de Nuevas Sinergias Arte-Ciencia aplicadas a la Conservación y Restauración de los Palacios y Jardines de la Alhambra y el Generalife” (VIRARTE), MINECO, with reference no. HAR2016-79886-P; “Métodos sinérgicos Arte-Ciencia-Tecnología para la Conservación-Restauración de la Alhambra y otros Bienes Culturales” (VIRARTE II), MICINN, with reference no. PID2019-109713RB-I00; “La aplicación de las algas procedentes de la Alhambra y el Generalife en técnicas artísticas y de conservación-restauración” (FICOARTE), Universidad de Granada, with reference no. A-HUM-279-UGR18. And the contract “Aplicación de nuevas técnicas de diagnosis y tratamiento de las microalgas presentes en la arquitectura del agua de la Alhambra, mediante la combinación de disciplinas”, Patronato de la Alhambra y el Generalife, with reference no. Adenda 1/2017.

Acknowledgments

We would like to thank the Alhambra and Generalife Trust for their logistical support, without which it would not have been possible to do this work. We would also like to thank the Vice-Rector for Research and Knowledge Transfer, Enrique Herrera-Viedma, for his support on this research and for his initiative in creating the Unit of Excellence “Science in the Alhambra”, as well as the anonymous reviewers for their constructive criticism and suggestions.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Hueck, H.J. The biodeterioration of materials as a part of hylobiology. Mater. Org. 1965, 1, 5–34. [Google Scholar]
  2. Anagnostidis, K.; Economou-Amilli, A.; Roussomoustakaki, M. Epilithic and chasmolithic microflora (Cyanophita, Bacillariophyta) from marbles of the Parthenon (Acropolis-Athens, Greece). Nova Hedwigia 1983, 38, 227–287. [Google Scholar]
  3. Ortega-Calvo, J.J.; Sánchez-Castillo, P.M.; Hernández-Marine, M.; Sáiz-Jiménez, C. Isolation and characterization of epilithic chlorophytes and cyanobacteria from two Spanish cathedrals (Salamanca and Toledo). Nova Hedwigia 1993, 57, 239–253. [Google Scholar]
  4. Bolivar-Galiano, F.; Peraza-Zurita, Y.; Romero-Noguera, J.; Sanchez-Castillo, P. L’Alhambra a Granada:15 anni di indagini sulla conservazione delle fontane. In L’Acqua, le Pietre i Bronzi. le Fontane Monumentali; Palombi Editori: Roma, Italy, 2010. [Google Scholar]
  5. Vázquez-Nion, D.; Rodriguez-Castro, J.; López-Rodríguez, M.; Fernández-Silva, I.; Prieto, B. Subaerial biofilms on granitic historic buildings: Microbial diversity and development of phototrophic multi-species cultures. Biofouling 2016, 32, 657–669. [Google Scholar] [CrossRef] [PubMed]
  6. Ibarra-Gallardo, C.E.; Novelo, E. Algas y cianoprocariontes epilíticos de la Zona Arqueológica de Yaxchilán, Chiapas, México. Rev. Mex. Biodivers. 2018, 89, 590–603. [Google Scholar] [CrossRef]
  7. Mohammadi, P.; Gholami-Nejad, P.; Asghari-Daryasari, R.; Asgarani, E. The study of microbial communities of rudkhan castle. Geomicrobiol. J. 2019, 37, 119–129. [Google Scholar] [CrossRef]
  8. Favero-Longo, S.E.; Viles, H. A review of the nature, role and control of lithobionts on stone cultural heritage: Weighing-up and managing biodeterioration and bioprotection. World J. Microbiol. Biotechnol. 2020, 36, 1–18. [Google Scholar] [CrossRef] [PubMed]
  9. Golubic, I.F.S. The lithobiontic ecological niche, with special reference to microorganisms. J. Sediment. Res. 1981, 51, 475–478. [Google Scholar] [CrossRef]
  10. Galiano, F.C.B.; Castillo, P.M.S. Biodeterioro del patrimonio artístico por cianobacterias, algas verdes y diatomeas. Rev. PH 1998, 24, 52. [Google Scholar] [CrossRef]
  11. Hoiczyk, E. Gliding motility in cyanobacteria: Observations and possible explanations. Arch. Microbiol. 2000, 174, 11–17. [Google Scholar] [CrossRef] [PubMed]
  12. Saiz-Jimenez, C. Biogeochemistry of weathering processes in monuments. Geomicrobiol. J. 1999, 16, 27–37. [Google Scholar] [CrossRef]
  13. Bolívar-Galiano, F.; Sánchez-Castillo, P.M. Biomineralization processes in the fountains of the Alhambra, Granada, Spain. Int. Biodeterior. Biodegrad. 1997, 40, 205–215. [Google Scholar] [CrossRef]
  14. Peraza Zurita, Y. Biodeterioro por Microalgas en Fuentes de Mármol. Ph.D. Thesis, University of Granada, Granada, Spain, 2004. [Google Scholar]
  15. Bolívar-Galiano, F.C. Diagnosis y Tratamiento del Deterioro por Microalgas en los Palacios Nazaríes de la Alhambra. Ph.D. Thesis, University of Granada, Granada, Spain, 1994. [Google Scholar]
  16. Sánchez Castillo, P.M.; Bolívar Galiano, F.C. Caracterización de comunidades algales epilíticas en fuentes monumentales y su aplicación a la diagnosis del biodeterioro. Limnetica 1997, 13, 31–46. [Google Scholar]
  17. Sarró, M.I.; García, A.M.; Rivalta, V.M.; Moreno, D.A.; Arroyo, I. Biodeterioration of the lions fountain at the Alhambra Palace, Granada (Spain). Build. Environ. 2006, 41, 1811–1820. [Google Scholar] [CrossRef]
  18. Cuzman, O.A.; Ventura, S.; Sili, C.; Mascalchi, C.; Turchetti, T.; D’Acqui, L.P.; Tiano, P. Biodiversity of phototrophic biofilms dwelling on monumental fountains. Microb. Ecol. 2010, 60, 81–95. [Google Scholar] [CrossRef] [PubMed]
  19. Bourrelly, P. Les Algues d’eau Douce. Algues Bleues et Rouges; Éditions N. Boubée and Cie: Paris, France, 1970. [Google Scholar]
  20. Komárek, J.; Anagnostidis, K. Süßwasserflora von Mitteleuropa, Bd. 19/1: Cyanoprokaryota. 1st Part: Chroococcales; Springer Spektrum: Berlin/Heidelberg, Germany, 1998. [Google Scholar]
  21. Galiano, F.C.B.; Castillo, P.M.S. Claves de identificación de microalgas frecuentes en monumentos. Rev. PH 1999, 26, 93. [Google Scholar] [CrossRef]
  22. Komárek, J.; Anagnostidis, K. Süßwasserflora von Mitteleuropa, Bd. 19/2: Cyanoprokaryota. 2nd Part: Oscillatoriales; Elsevier GmbH: Munich, Germany, 2005. [Google Scholar]
  23. Komárek, J. Süßwasserflora von Mitteleuropa, Bd. 19/3: Cyanoprokaryota. 3nd Part: Heterocytous Genera; Springer Spektrum: Berlin/Heidelberg, Germany, 2013. [Google Scholar]
  24. Peraza, Y.; Cultrone, G.; Sánchez-Castillo, P.; Bolívar-Galiano, F. Il biodeterioramento delle fontane dei Reales Alcazares di Siviglia e dell’Alhambra di Granada (Spagna). Sci. Tech. Cult. Heritage 2002, 11, 111–118. [Google Scholar]
  25. Peraza Zurita, Y.; Cultrone, G.; Sánchez Castillo, P.; Sebastián, E.; Bolívar, F.C. Microalgae associated with deteriorated stonework of the fountain of Bibataubín in Granada, Spain. Int. Biodeterior. Biodegrad. 2005, 55, 55–61. [Google Scholar] [CrossRef]
  26. Gaylarde, C.C.; Gaylarde, P.M. A comparative study of the major microbial biomass of biofilms on exteriors of buildings in Europe and Latin America. Int. Biodeterior. Biodegrad. 2005, 55, 131–139. [Google Scholar] [CrossRef]
  27. Macedo, M.F.; Miller, A.Z.; Dionisio, A.; Saiz-Jimenez, C. Biodiversity of cyanobacteria and green algae on monuments in the Mediterranean Basin: An overview. Microbiology 2009, 155, 3476–3490. [Google Scholar] [CrossRef] [PubMed] [Green Version]
Figure 1. (a) Two different phycofilms on the Fuente del Patio de la Reja at the Nasrid Palaces. (b) Algal mat growing on the Fuente del Patio de la Sultana at Generalife. Photos by G. Alfano.
Figure 1. (a) Two different phycofilms on the Fuente del Patio de la Reja at the Nasrid Palaces. (b) Algal mat growing on the Fuente del Patio de la Sultana at Generalife. Photos by G. Alfano.
Applsci 10 06603 g001
Figure 2. Sampling from the Fountain of the Lions. Photo by G. Alfano.
Figure 2. Sampling from the Fountain of the Lions. Photo by G. Alfano.
Applsci 10 06603 g002
Figure 3. Sampling points in the monumental complex of the Alhambra and the Generalife. The labels correspond to Table 1.
Figure 3. Sampling points in the monumental complex of the Alhambra and the Generalife. The labels correspond to Table 1.
Applsci 10 06603 g003
Figure 4. (a) Sampling a mineralising crust with a scalpel at Fuente N del Patio de los Arrayanes. (b) Two sheeted crusts from perpendicular angles with buried communities. Photos by G. Alfano.
Figure 4. (a) Sampling a mineralising crust with a scalpel at Fuente N del Patio de los Arrayanes. (b) Two sheeted crusts from perpendicular angles with buried communities. Photos by G. Alfano.
Applsci 10 06603 g004
Figure 5. (a) South fountain of the Lower Gardens. The base of the basin, where a dark green mat can be seen, was one of the sampling points; (b) Taking a sample of mineral crust in the Escalera del Agua (Water Stairway); (c) Sample of mineral crust taken in the Washington Irving basin. Photos by G. Alfano.
Figure 5. (a) South fountain of the Lower Gardens. The base of the basin, where a dark green mat can be seen, was one of the sampling points; (b) Taking a sample of mineral crust in the Escalera del Agua (Water Stairway); (c) Sample of mineral crust taken in the Washington Irving basin. Photos by G. Alfano.
Applsci 10 06603 g005
Figure 6. Photographs of samples taken using optical microscopy: (a) Chlorosarcinopsis sp. cells can be seen on the left; (b) Cosmarium sp. cells can be seen on the right, along with a filament of cyanobacteria.
Figure 6. Photographs of samples taken using optical microscopy: (a) Chlorosarcinopsis sp. cells can be seen on the left; (b) Cosmarium sp. cells can be seen on the right, along with a filament of cyanobacteria.
Applsci 10 06603 g006
Figure 7. Photograph of samples taken using optical microscopy: (a) Pleurocapsa sp. cells are shown on the left; (b) Chamaesiphon sp. cells are shown on the right.
Figure 7. Photograph of samples taken using optical microscopy: (a) Pleurocapsa sp. cells are shown on the left; (b) Chamaesiphon sp. cells are shown on the right.
Applsci 10 06603 g007
Figure 8. Photograph of samples taken using optical microscopy: (a) diatoms of the Cymbella, Navicula and Achnanthes genera and filaments of Phormidium sp. can be seen on the left; (b) on the right, there are cells of Bracteacoccus sp. and filaments of Phormidium sp.
Figure 8. Photograph of samples taken using optical microscopy: (a) diatoms of the Cymbella, Navicula and Achnanthes genera and filaments of Phormidium sp. can be seen on the left; (b) on the right, there are cells of Bracteacoccus sp. and filaments of Phormidium sp.
Applsci 10 06603 g008
Figure 9. Photographs of the algal cultures Bracteacoccus sp. (a) and Klebsormidium sp. (b) and two different species of Leptolyngbya (c,d).
Figure 9. Photographs of the algal cultures Bracteacoccus sp. (a) and Klebsormidium sp. (b) and two different species of Leptolyngbya (c,d).
Applsci 10 06603 g009
Figure 10. Presence of microalgae on one of the lions in the Fountain of the Lions: (a) an image from 1994 by F. Bolivar is shown on the left; (b) a modern image by G. Alfano in which one can see the effect of the cleaning treatment and lack of large sheens on the surface is shown on the right.
Figure 10. Presence of microalgae on one of the lions in the Fountain of the Lions: (a) an image from 1994 by F. Bolivar is shown on the left; (b) a modern image by G. Alfano in which one can see the effect of the cleaning treatment and lack of large sheens on the surface is shown on the right.
Applsci 10 06603 g010
Figure 11. Microalgal cultures: (a) an initial culture using a sample is shown on the left; (b) a unialgal culture of Neochloris sp is shown on the right.
Figure 11. Microalgal cultures: (a) an initial culture using a sample is shown on the left; (b) a unialgal culture of Neochloris sp is shown on the right.
Applsci 10 06603 g011
Table 1. Fountains studied in different areas of the Alhambra and numbers of samples collected. In the last two columns, in brackets, the numbers of samples taken corresponding to different materials and environments are shown.
Table 1. Fountains studied in different areas of the Alhambra and numbers of samples collected. In the last two columns, in brackets, the numbers of samples taken corresponding to different materials and environments are shown.
ZoneFountainNo. of SamplesMaterialEnvironment
Nasrid Palaces
  • Fuente del Patio del Cuarto Dorado (Mexuar fountain)
7marble (5), marble and metal (1), mortar (1)submerged (6), amphibious (1)
2.
Fuente-guitarra N del Patio de los Arrayanes (North “guitar” fountain of the Court of the Myrtles)
15marble (10), metal (5)submerged (9), amphibious (5), aerial (1)
3.
Fuente-guitarra S del Patio de los Arrayanes (South “guitar” fountain of the Court of the Myrtles)
6marble (5), metal (1)submerged (4), amphibious (2)
4.
East channel of the Court of the Myrtles
2marble (2)submerged (1), amphibious (1)
5.
Fuente de los Leones (Fountain of the Lions)
21marble (15), mortar (3), metal (3)submerged (4), amphibious (13), aerial (4)
6.
Fuente del Patio de la Reja (Fountain of the Court of the Grille)
11marble (11)submerged (8), amphibious (3)
7.
Fuente del Patio de Lindaraja (Fountain of the Court of the Lindaraja)
10marble (10)submerged (1), amphibious (5), aerial (4)
Partal
8.
Fuente-guitarra junto a la Torre de las Damas
6marble (5), metal (1)submerged (4), amphibious (2)
9.
Pilar (Basin) debajo del Mirador del Partal
5brick (5)submerged (1), amphibious (2), aerial (2)
Casa del Arquitecto
10.
Inner fountain of the court with oranges
1painted plaster (1)aerial (1)
Alcazaba
11.
Pilar E. del Jardín de los Adarves (East basin of the Garden of the Ramparts)
3marble (3)amphibious (2), aerial (1)
Secano
12.
Fuente de la Glorieta del Secano
5marble (5)submerged (3), amphibious (2)
Generalife
13.
Fuente S de los Jardines Bajos (South fountain of the Lower Gardens)
2marble (2)submerged (1), amphibious (1)
14.
Fuente central del Patio de la Sultana (Central fountain of the Sultana’s Court)
7marble (6), brick (1)submerged (2), amphibious (2), aerial (3)
15.
Pretil de la Escalera del Agua (The handrail of the Water Stairway)
4glazed tile (2), plaster (2)submerged (2), amphibious (2)
Walkways
16.
Fuente del Tomate (Tomato fountain)
3marble (3)submerged (3)
17.
Fuente de Ángel Ganivet
3bronze (1), mortar (2)submerged (2), amphibious (1)
18.
Pilar de Carlos V
3marble (3)submerged (1), amphibious (2)
19.
Pilar de Washington Irving
3marble (3)amphibious (2), aerial (1)
20.
Pilar de la Puerta de las Granadas (Basin of the Gate of the Pomegranates)
2marble (2)aerial (2)
Carmen de Bellavista
21.
Fuente derecha de la 2ª terraza (Right fountain of the 2nd terrace)
1marble (1)amphibious (1)
Total samples120
Table 2. The most common genera of microalgae in each of the fountains studied. For each fountain, the set of samples taken was considered, along with the different sampling points, and where applicable the different times of year. The genera that appear most often and in greater abundance were chosen. For each genus, the total number of samples in which it is present (first number) and the number of samples in which its index is 3 or higher (second number; see 2.2. Procedures) are shown in brackets.
Table 2. The most common genera of microalgae in each of the fountains studied. For each fountain, the set of samples taken was considered, along with the different sampling points, and where applicable the different times of year. The genera that appear most often and in greater abundance were chosen. For each genus, the total number of samples in which it is present (first number) and the number of samples in which its index is 3 or higher (second number; see 2.2. Procedures) are shown in brackets.
FountainMost Common Genera
Green AlgaeCyanobacteriaDiatoms
  • Mexuar fountain
Bracteacoccus (3-2), Chlorosarcinopsis (2-1) Navicula (3-1), Nitzschia (2-2)
2.
North “guitar” fountain of the Court of the Myrtles (Patio de los Arrayanes)
Chlorosarcinopsis (5-2)Dichothrix (5-3), Phormidium (6-5), Leptolyngbya (9-6),
3.
South “guitar” fountain of the Court of the Myrtles
Phormidium (4-2), Calothrix (4-1), Chlorogloea (1-1), Pleurocapsa (3-1)
4.
East channel of the Court of the Myrtles
Leptolyngbya (2-2), Pleurocapsa (2-1)
5.
Fountain of the Lions
Klebsormidium (3-2)Phormidium (7-6), Leptolyngbya (9-2), Cyanosarcina (5-4)
6.
Patio de la Reja fountain
Chlorosarcina (7-5)Symploca (1-1), Leptolyngbya (5-4), Chamaesiphon (3-2), Calothrix (2-1)
7.
Patio de Lindaraja fountain
Chlorococcum (4-2), Chlorosarcinopsis (4-1), Klebsormidium (5-2)Chamaesiphon (6-3), Phormidium (2-2)
8.
Ladies Tower (Torre de las Damas)
Pleurastrum (2-2)Symploca (3-2), Phormidium (3-2), Leptolyngbya (2-1)
9.
Mirador del Partal basin
Bracteacoccus (4-4) Navicula (2-1)
10.
Inner fountain of the court with oranges
Klebsormidium (1-1)
11.
Jardín de los Adarves east basin
Bracteacoccus (3-2), Choricystis (1-1)
12.
Glorieta del Secano fountain
Bracteacoccus (4-3)Chamaesiphon (2-1), Cyanosarcina (1-1)
13.
Jardines Bajos South fountain
Phormidium (2-2)
14.
Patio de la sultana central fountain
Gongrosira (3-1), Bracteacoccus (3-3)Phormidium (3-1), Lyngbya (1-1)Navicula (4-1), Cymbella (3-1)
15.
Escalera del Agua handrail
Pseudopleurococcus (1-1), Leptosira (2-1), Chlorosarcina (2-2)Myxosarcina (1-1), Chlorogloea (1-1)
16.
Fuente del tomate
Bracteacoccus (2-1)Chlorogloea (1-1), Chamaesiphon (1-1), Phormidium (2-1)
17.
Ángel Ganivet pool fountain
Bracteacoccus (2-2)Leptolyngbya (1-1)
18.
Carlos V basin
Bracteacoccus (3-1)Chamaesiphon (2-1)
19.
Washington Irving basin
Bracteacoccus (2-1)
20.
Puerta de las Granadas basin
Bracteacoccus (1-1)Chroococcopsis (1-1)
21.
Right fountain of the 2nd terrace of the Carmen de Bellavista
Calothrix (1-1), Phormidium (1-1)
Table 3. Species obtained from unialgal cultures using samples collected in the Alhambra and Generalife. The fountain numbers (Table 1 and Table 2) from which the cultures came are indicated in the left column.
Table 3. Species obtained from unialgal cultures using samples collected in the Alhambra and Generalife. The fountain numbers (Table 1 and Table 2) from which the cultures came are indicated in the left column.
FountainsGreen AlgaeFountainsCyanobacteria
4Neochloris sp.2Dichothrix sp.
14Stigeoclonium sp.4, 8, 15, 17Leptolyngbya spp.
14, 8Chlorococcum sp.20, 7Chroococcopsis sp.
9, 12, 16, 19Bracteacoccus sp.2, 8Pseudophormidium sp.
4, 8, 14Scenedesmus spp.13Pseudanabaena sp.
15Pseudopleurococcus sp.6, 21Calothrix sp.
5, 10Klebsormidium sp.3Pleurocapsa sp.
11, 13Choricystis sp.3Phormidium sp.
3Apatococcus sp.5Cyanosarcina sp.
4Gloeocystis sp.5Schizothrix sp.
5Nostoc sp.
5Chroococcidiopsis sp.
4Ammatoidea sp.

Share and Cite

MDPI and ACS Style

Bolívar-Galiano, F.; Abad-Ruiz, C.; Sánchez-Castillo, P.; Toscano, M.; Romero-Noguera, J. Frequent Microalgae in the Fountains of the Alhambra and Generalife: Identification and Creation of a Culture Collection. Appl. Sci. 2020, 10, 6603. https://doi.org/10.3390/app10186603

AMA Style

Bolívar-Galiano F, Abad-Ruiz C, Sánchez-Castillo P, Toscano M, Romero-Noguera J. Frequent Microalgae in the Fountains of the Alhambra and Generalife: Identification and Creation of a Culture Collection. Applied Sciences. 2020; 10(18):6603. https://doi.org/10.3390/app10186603

Chicago/Turabian Style

Bolívar-Galiano, Fernando, Clara Abad-Ruiz, Pedro Sánchez-Castillo, Maurizio Toscano, and Julio Romero-Noguera. 2020. "Frequent Microalgae in the Fountains of the Alhambra and Generalife: Identification and Creation of a Culture Collection" Applied Sciences 10, no. 18: 6603. https://doi.org/10.3390/app10186603

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