Next Article in Journal
The Quality of Hilton Hotel Services in the V4 Countries: The Impact of the Platform on Ratings and Customer Satisfaction
Previous Article in Journal
Personalisation and Predictive Marketing in a Croatian Tourist Destination: Behavioural Strategies for Enhancing the Tourist Experience
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Tourism, Design and Climate Change: The Urban Glaciology Experiment at Fuorisalone 2024 Event

by
Antonella Senese
1,
Cecilia D. Almagioni
1,2,*,
Davide Fugazza
1,
Blanka Barbagallo
1,
Lorenzo Cresi
1,
Maurizio Maugeri
1 and
Guglielmina A. Diolaiuti
1
1
Environmental Science and Policy Department, University of Milan, Via Celoria 10, 20133 Milan, Italy
2
Department of Physics “Aldo Pontremoli”, University of Milano, Via Celoria 16, 20133 Milan, Italy
*
Author to whom correspondence should be addressed.
Tour. Hosp. 2025, 6(4), 168; https://doi.org/10.3390/tourhosp6040168
Submission received: 29 July 2025 / Revised: 26 August 2025 / Accepted: 1 September 2025 / Published: 4 September 2025
(This article belongs to the Special Issue Tourism Event and Management)

Abstract

Glacier retreat due to climate change is accelerating worldwide, yet the phenomenon remains abstract for many people, especially those unfamiliar with mountain environments. The Urban Glaciology experiment, conducted during Milan’s internationally renowned “Fuorisalone” 2024 design event, aimed to bridge this perceptual gap by simulating real glacier melt processes in a busy urban square. Three large ice blocks with contrasting surface conditions (i.e., clean, dirty, and debris-covered) were exposed to springtime urban temperatures, mimicking conditions found on Alpine glaciers during summer. Over one week, the blocks produced a total of 748 L of meltwater, with dirty ice melting up to four times faster than debris-covered ice, consistent with established albedo effects. These results confirmed the thermal analogy between Milan’s spring conditions (+15 to +20 °C) and the ablation season on Alpine glaciers. Visitors observed the differential melting in real time, supported by visual indicators, explanatory panels, immersive virtual experiences, and direct interaction with researchers and students. Informal interviews indicated that more than 60% of participants reported a perceptual shift, recognizing for the first time that urban temperatures can replicate glacier melting conditions. By embedding a science-based installation in a major cultural tourism event, the experiment reached a diverse, non-traditional audience—including tourists, designers, and citizens—and encouraged reflection on the implications of glacier loss. The success of this initiative highlights the potential of replicating similar models in other cities to raise awareness of environmental change through culturally engaging experiences.

1. Introduction

Urban areas are increasingly becoming pivotal arenas for climate change communication and public engagement, especially when cultural and tourism events attract large and diverse audiences (Moser, 2006, 2013, 2014). In this context, experiential installations and science-based exhibits embedded in popular urban happenings can serve as powerful tools to raise awareness about environmental issues, even among audiences not typically engaged with science or nature-based tourism (Capari et al., 2022; Saladié et al., 2025).
One such opportunity is offered by the “Fuorisalone” event, the vibrant and globally renowned urban design festival that animates the city of Milan (Italy) each April in conjunction with the Salone del Mobile. While the latter takes place in a single exhibition venue (Gramegna et al., 2024), “Fuorisalone” unfolds across the entire city through hundreds of temporary installations, exhibitions, and interactive events hosted in design districts, showrooms, museums, and public spaces. In recent years, “Fuorisalone” has become a key cultural and tourist attraction, drawing hundreds of thousands of visitors from around the world (e.g., design professionals, students, international tourists, and urban explorers) creating a unique opportunity to intercept and engage a broad public (Balduini et al., 2013).
We performed a public scientific experiment during the “Fuorisalone” 2024 and it was the occasion to meet several hundred people and to introduce them to climate change and its environmental impacts. The scientific experiment we performed was named “Urban Glaciology Experiment” and was aimed at showing to the general public glacier ice melting, its magnitude and rates, without the need to really visit a glacier. An experiment performed in the urban context made it possible to everyone, without leaving the city, to see processes and phenomena otherwise visible only in high mountain areas, spending time and money to reach such remote zones.
The Urban Glaciology experiment leveraged this context to bring the message of climate change into the heart of the city. Conducted in Piazza Città di Lombardia, the central square in front of the headquarters of the Lombardy Region, the project was part of the exhibition “Lombardy on the Roof of the World” developed and managed by an Italian NGO, named EvK2CNR, in cooperation with the University of Milan. The centerpiece of the installation was a pyramid. This was shown to visitors before being sent to the Himalayas to replace the Laboratory Pyramid, located in Sagarmatha National Park, at an altitude of about 5000 m, which, after 30 years of scientific activity, needed extensive renovation. The Pyramid lab is managed by EvK2CNR, this association has been coordinating research on the roof of the world (at the foot of Mount Everest) since the last three decades (Salerno et al., 2023). Before being shipped to Nepal, the replacement prototype was assembled and put on display at the “Fuorisalone” to show that the design is also of interest to the scientific world.
Within this setting, we installed three large ice blocks with contrasting surface conditions (i.e., bare, dirty, and debris-covered ice) mimicking real glacier surfaces found in the Alps and worldwide.
Tourism is both affected by and contributes to climate change. On the one hand, glacier retreat and the loss of snow cover are already reshaping mountain tourism, altering landscapes and reducing the availability of emblematic attractions such as glaciers (Palomo, 2017; Stewart et al., 2016). On the other hand, cultural and urban tourism events provide opportunities to engage diverse audiences with environmental challenges, making climate change communication part of the tourism experience itself (Markowitz & Guckian, 2018). Within this perspective, the Urban Glaciology experiment is not only a scientific outreach initiative but also an example of how environmental education can be integrated into tourism contexts, reaching international visitors, design enthusiasts, and urban explorers who attended the Fuorisalone 2024 event.
A key challenge in climate change communication in urban areas is reaching individuals who are not already engaged with environmental issues (Chu & Schenk, 2017; Wamsler et al., 2013). The Urban Glaciology experiment specifically addressed this by placing a scientifically grounded, visually striking installation in a high-traffic cultural setting, thus intercepting audiences such as design enthusiasts, international visitors, and casual passers-by. This strategy enabled people with little prior exposure to glaciology to experience glacier melting firsthand, transforming an abstract phenomenon into a tangible and relatable reality.
Over the course of a week, visitors were invited to observe in real time the melting process of the ice blocks under typical springtime urban temperatures (air temperatures ranging from +15 to +20 °C), a thermal regime that also occurs on Alpine glaciers during summer (Carturan et al., 2015). The melting of the ice was visually tracked using markers and digital displays, while informative panels and the presence of researchers enabled direct interaction with the public. The experiment highlighted the climatic parallels between cities and mountains and made visible the invisible melting occurring on glaciers at 2800 m or higher, where many visitors might never go (Gagné et al., 2014; Moreau, 2010).
By staging this scientifically grounded, visually compelling simulation in a highly trafficked urban space during a major international event, we aimed to bridge the perceptual gap between distant glacier environments and everyday urban life. In doing so, the Urban Glaciology experiment exemplifies how climate change education can be embedded within urban tourism and cultural experiences, opening up new pathways for transformative and informal environmental learning (Barrett et al., 2017; Singer-Brodowski, 2023). Tourists themselves are also affected by glacier retreat, as climate change is altering destinations and experiences associated with glacier tourism (Barrett et al., 2017; Stewart et al., 2016).
Glaciers are vital components of the hydrological and climatic systems that sustain millions of people around the globe (IPCC, 2023). From the Alps to the Himalayas, seasonal glacier melting feeds rivers, powers hydroelectric plants, and supports agriculture in densely populated downstream regions (Schaefli et al., 2019; Cook et al., 2021; Robel et al., 2024). Yet, as climate change accelerates, glaciers are retreating at unprecedented rates, transforming from perennial sources of water to powerful symbols of environmental loss (IPCC, 2023; Palomo, 2017).
Despite their importance, glaciers remain distant and abstract for most people (Senese et al., 2023a). Scientific data and satellite imagery, though compelling, often fail to bridge the perceptual gap between remote mountain environments and everyday urban life (Newman, 2018). This disconnection poses a challenge for public engagement and climate communication, highlighting the need to raise up a shared understanding of glacial change and its broader implications. In response, emerging initiatives are beginning to explore new formats of communication and education that make glacial change more relatable and visible within everyday contexts (Barbagallo et al., 2024; Gaspari et al., 2025).
The Urban Glaciology experiment seeks to address this challenge through a tangible and visual simulation of glacier melting in a city setting. By replicating key surface conditions found on real and exposing them to ambient urban weather conditions, this project offers an accessible and scientifically grounded way to visualize the dynamics of glacier melt.
Despite the growing interest in climate change communication, three main gaps remain evident. First, there is a lack of accessible, experiential tools that allow urban audiences to grasp glacier melting processes without visiting remote mountain areas (Gagné et al., 2014). Second, the integration of climate change education into cultural and tourism events is still limited, with most outreach activities occurring in scientific or natural settings (Gaspari et al., 2025; Newman, 2018). Third, non-traditional audiences (e.g., tourists, design enthusiasts, and casual passers-by) are rarely targeted, even though they represent an important segment for broadening public engagement with climate issues (Barbagallo et al., 2024). This study addresses these gaps by presenting the Urban Glaciology experiment, a science-based installation embedded in Milan’s Fuorisalone 2024 design event. The contributions of our work are threefold: (i) demonstrating a tangible and intuitive simulation of glacier melt under urban spring conditions that resemble Alpine summer environments, (ii) embedding scientific research within a high-profile cultural tourism event to maximize outreach, and (iii) evaluating the effectiveness of the installation in engaging a diverse audience and stimulating reflection on glacier loss. In doing so, this study adds to the growing body of literature on informal climate change education and proposes a replicable model for embedding environmental awareness into urban cultural experiences.

2. The Urban Glaciology Experiment at the Milano “Fuorisalone”

The urban glaciology experiment was carried out during the “FuoriSalone” event of the Milan Design Week, the leading international event for the design industry (15–21 April 2024) with an open-air lab in front of the Building of Lombardy Region (the building hosting the headquarter of the administrative district where Milan is located, https://www.fuorisalone.it/en/2024/events/3909/LOMBARDY-ON-THE-ROOF-OF-THE-WORLD-FUORISALONE-2024, accessed on 25 July 2025). The Lab was called “Lombardy on the roof of the world” as in the square a pyramid (17 × 17 × 11 m, Figure 1) was located before being sent to Nepal to replace the Italian Pyramid Observatory Laboratory located at 5050 m a.s.l. in the Khumbu valley (Nepal) at the foot of the Nepalese side of Everest. The Pyramid Lab in Nepal hosts instruments and sensors to survey atmosphere and meteorological conditions in a peculiar glacierized area of the World, the Nepalese Himalaya. This station was installed more than 30 years ago from an Italian team lead by Ardito Desio, a famous scientist who also led the Italian expedition who reach the K2 top in 1954. Presently the Lab is managed by EvK2CNR, an NGO active in the High Mountain Asia. This lab represents one of the Italian excellence laboratories for research at high elevations (Bonasoni et al., 2010; Salerno et al., 2023). After three decades of data acquisition at an altitude of 5000 m, the pyramid laboratory is in need of extensive renovation. A new pyramid is therefore required to house the instruments and equipment and to envelop the official pyramid during the renovation work. This support pyramid was shown to the citizens of Milan and visitors to the “Fuorisalone” and was then shipped to Nepal to protect the laboratory pyramid. Inside this structure, we carried out the urban glaciology experiment in April 2024.
Urban Glaciology experiment involved the exposure of 3 ice blocks (each 0.5 m3 in volume) to urban spring thermal conditions and diffused solar radiation. The blocks present the upper surface with 3 different conditions, representative of the 3 types of real surface conditions of Alpine and worldwide glaciers, i.e.,
  • Clean ice (bare ice conditions), which means ice without visible impurities and deposits, freely exposed to sunlight, with albedo (i.e., solar reflectivity) between 0.25 and 0.30 (Hartmann, 2015);
  • Darkening ice (dirty ice conditions), meaning ice with the deposition of dust and black carbon on the surface, substances resulting from atmospheric pollution and forest fires that deposit on the surface of glaciers, reducing their albedo and increasing melting;
  • Buried ice conditions, which means ice covered by gravelly and sandy debris that can have a different effect on the melting of the underlying ice depending on its thickness: if it is thicker than about 5 cm, it acts as an insulator and reduces melting; if it is thinner than about 5 cm, it reduces the albedo of the surface and conducts heat, increasing melting.
The three blocks had the three upper surfaces exposed to solar spectrum radiation and natural air temperature (16–21 April 2024 conditions) without artificial conditioning or cooling. During the days of exposure, the urban glaciology experiment included:
  • Continuous measurements of surface melt rates indicating on the Plexiglass box (in which the ice blocks were located) the daily level of the ice and the date in order to allow visitors to observe the progressive reduction in the ice block thickness and thus the intensity of surface ice melting under the three different proposed conditions;
  • Periodic surveys with close-range photogrammetry techniques to reconstruct the volume variation in the blocks over time and the volume lost on a daily and total scale;
  • Continuous measurements of the meltwater produced by the three ice blocks and its virtuous reuse for urban gardening.
The atmospheric and energy conditions of the experiment area were continuously monitored thanks to an automatic weather station (AWS) installed in the Pyramid close to the ice blocks. The AWS acquired data every minute and recorded the main meteorological parameters, such as air temperature, atmospheric pressure, wind speed and velocity, relative humidity and incoming solar radiation.
The visitors not only observed the ice melting blocks but also were involved in several activities of citizen science. In fact, on that occasion, we offered immersive glacier experience through VR headsets to raise awareness of what a glacier is, and its sensitivity to climate change (Barbagallo et al., 2024; Gaspari et al., 2025).
Degree students and PhD students from the University of Milan participated in the experiment. They worked under the guidance of professors from the Department of Environmental Science and Policy who are experts in Alpine and Himalayan glaciology. The young researchers answered questions from visitors to the stand, administered citizen science web app (for estimating the individual carbon footprint and plastic footprint, Senese et al., 2023b, 2024) and collected data on ice melting, all under the curious gaze of visitors. In addition, qualitative feedback from visitors was collected through brief, informal interviews carried out on-site during the installation. These conversations, conducted by researchers and students, aimed to capture participants’ immediate impressions, levels of surprise, and perceived understanding of glacier melting. Responses were not recorded in a structured database but noted as qualitative observations to provide insights into public engagement outcomes. However, visitor feedback was collected using a structured questionnaire designed to evaluate the effectiveness of immersive 360° glacier experiences through VR headsets. This tool, reported in a previous study (Barbagallo et al., 2024), included questions on demographic characteristics (e.g., gender, occupation, prior knowledge of glaciers), previous experiences with glaciers, and interest in climate change. The questionnaire was administered on-site during the Fuorisalone event in both Italian and English, depending on the respondent’s language. A total of 335 valid responses were collected (291 Italian and 44 international participants).
The data collected allow a physically based model of urban glacier melting to be described in the three proposed situations.

3. Materials and Methods

3.1. Ideal Model

The original idea was to reproduce three different kinds of ice surfaces (Figure 2): buried ice (i.e., covered by a thick debris layer), dirty ice (i.e., covered by dust), and bare ice (i.e., free of dust and debris). The three ice blocks were intended to be physically separated and thermally insulated from each other, with only their upper surfaces exposed to energy fluxes. This setup was designed to highlight the differences in melting rates associated with varying surface albedo conditions. For each block, simple energy balance models (Equations (1) and (2)) were intended to be applied independently to estimate melt (Mi):
M b a r e ,   d i r t y = q r a d · t L m = A · 1 α · S W i n + ε · σ · T 24 4 ε i c e · σ · T i c e 4 · t L m
where
  • qrad is the radiative heat transfer per unit time;
  • Δt is the time of exposition;
  • Lm is the latent heat of melting (3.34 × 105 J kg−1);
  • A is the surface of the exposed ice in m2;
  • α is the average albedo of the ice (0.3 for bare ice, 0.1 for dirty ice);
  • SWin is the incoming solar radiation, as measured by the AWS;
  • ε is the emissivity of the surroundings, which was assumed to be 0.75, accounting for the combined radiative properties of the materials present in the urban environment, including concrete paving, glass facades, and metallic surfaces;
  • σ is the Stefan-Boltzmann constant (5.67 × 10−8 W m2 K−4);
  • T24 is the temperature of the air, averaged over the preceding 24 h;
  • εice is the emissivity of the ice considered equal to 1;
  • Tice is the surface temperature of the ice, considered at the melting point.
For the bare and dirty ice blocks, the energy input was to be evaluated separately, with the absorbed shortwave radiation modulated according to each surface’s specific albedo. This would have enabled a comparative analysis of how surface characteristics influence melting under identical external conditions.
To model the melting of the buried ice, an energy-balance approach based on the conductive heat flux through the debris layer was adopted. The amount of ice melt under debris (Mburied) was estimated by computing the energy conducted to the ice–debris interface and converting it into meltwater equivalent (Equation (2)), following the approach of Nakawo and Takahashi (1982):
M b u r i e d = A · ( T s u p T i c e ) · t D R · L m
where the conductive flux through the debris layer is assumed to follow a linear temperature gradient, with the ice temperature fixed at 0 °C. Tsup corresponds to the temperature of debris surface.
The thermal resistance of the debris layer (DR) was derived empirically as:
D R = 19.841 d + 1.062
where d is the debris thickness (m), assumed constant over the short experimental period. The debris surface temperature (Tsup) was assumed to be in equilibrium with the ambient air temperature due to the specific experimental setup: the blocks were sheltered from direct solar radiation for most of the day by the surrounding buildings and tent, with only a brief exposure of approximately 10 min at noon. Consequently, radiative heating of the debris was minimal, and air temperature was the dominant driver of the surface energy balance.

3.2. Realization: Ice Melt Prevision

The experimental setup was designed by adapting to site-specific constraints, ensuring both inclusivity and feasibility. The three ice cubes were placed adjacent to each other within a single Plexiglas container (Figure 1 and Figure 2), which made it impossible to isolate the contribution of each individual cube to the overall melting. The walls of the structure were made of transparent Plexiglas to allow the public to visually monitor the ice melt over time, maximizing the inclusivity. To facilitate this observation, the height of the melting ice blocks was measured and marked daily.
Given the configuration, the applied energy balance model treats the system as a single, larger ice block characterized by an average albedo. A simplified model (Equation (4)) was applied to estimate the melt rate (M):
M = ( q r a d + q c o n v ) · t L m = A · [ 1 α · S W i n + ε · σ · T 24 4 ε i c e · σ · T i c e 4 + h · ( T a i r T i c e ) ] · t L m
where
  • qconv is the convective heat transfer per unit time;
  • α is the average albedo of the ice, assumed to be 0.2 (Cuffey & Paterson, 2010);
  • Tair is the temperature of the air, as measured by the AWS;
  • h is a convective exchange coefficient.
In this model, the thermal energy transferred to the ice per unit time was assumed to derive from two primary processes: convective heat transfer (qconv) and radiative heat transfer (qrad), the latter comprising both shortwave and longwave radiation.
Convective heat transfer was modeled using Newton’s law of cooling, with the convective heat transfer coefficient h assumed to be 5 W·m−2·K−1. This value is considered appropriate for ambient air temperatures ranging from +10 °C to +30 °C, which were observed throughout the experiment. Longwave radiation was modeled by assuming that both the ice surface and the surrounding Plexiglas walls emit thermal radiation according to the Stefan–Boltzmann law. The net radiative flux to the ice was calculated as the difference between incoming and outgoing longwave radiation, accounting for the emissivity of the surrounding environment, estimated at 0.75, and that of the ice, treated as a blackbody with an emissivity of 1. Shortwave radiation was estimated from the difference between the incoming global radiation, as measured by the AWS, and the reflected component, computed using the average albedo of the ice. Since the lateral surfaces of the ice block were also exposed through the transparent Plexiglas (and were not covered by debris), they were included in the radiative balance.

3.3. Meteorological Conditions

In order to investigate the conditions affecting the ice melt, an automatic weather station was installed near the ice cubes (Figure 1). The instrumentation used was similar to that used at high altitudes (Senese et al., 2016). The AWS, model “All in One Compact” (LSI Lastem S.r.l., Milan, Italy), measures air temperature (range −50 °C to +80 °C, accuracy ± 0.3 °C), relative humidity (0–100%, accuracy ± 3%), wind speed (0–60 m s−1, accuracy ± 0.3 m s−1) and direction, atmospheric pressure (500–1100 hPa, accuracy ± 0.5 hPa) and incoming shortwave radiation (0–2000 W m−2, accuracy ± 5%). Data points were sampled every second and then averaged every ten minutes, including the basic distribution parameters (minimum, mean and maximum values). This kind of stations has been widely used for environmental and cryospheric monitoring also in Asian contexts due to its robustness, portability, and ability to operate in harsh outdoor conditions (Dombrowski et al., 2021). The AWS ran from 15 April 2024 (the day the ice cubes were installed) to 21 April 2024 (when the ice installation was dismantled).

3.4. Close Range Low-Cost Photogrammetry

We carried out a further experiment aimed at reconstructing the volume of the blocks using low-cost photogrammetry (Koutlemanis et al., 2024). On 18, 19 and 21 April, we acquired images of the ice cubes using an iPhone 12 smartphone, carrying a 12 Mp digital camera. In total, 401 images were acquired on 18 April, 324 on 19 April and 381 on 21 April, by using continuous shooting mode and walking around the ice cubes in circle to provide acquisitions from all possible viewpoints. Image acquisition was set to automatic, ensuring the best combination of aperture, shutter speed and ISO sensitivity for the conditions at the time of acquisition. In practice, this resulted in fixed aperture (F/1.6) and ISO (32), with shutter speed allowed to vary between 1/206 and 1/438.
Processing of the photogrammetric blocks was carried out in Agisoft Metashape Professional version 2.2. Metashape implements a complete photogrammetric pipeline from tie point extraction, image orientation with generation of a sparse point cloud, point cloud densification to generate a dense point cloud (DPC) and generation of a mesh or digital elevation model (Agisoft LLC, 2025). To constrain the geometry of the photogrammetric blocks, measurements of the length, height and width of the container were taken with a measurement tape, and the position of its 8 lower and top corners was determined in a local coordinate system to be used as ground control points (GCPs). The GCPs were manually marked on the images and their coordinates were provided in Metashape for orientation of the blocks in the local coordinate system. In the absence of reliable information on the coordinates of the image projection centers, tie point extraction and image orientation were performed using sequential mode in Metashape, which uses the information from the image name assuming sequential numbering to speed up the search for potential matches. High resolution setting was used in this procedure, which uses the original images without downsampling (Agisoft LLC, 2025). Camera self-calibration was performed using bundle adjustment with the information from the GCPs as ground constraints. Then, the DPC was generated again using the high-resolution setting, which in this case performs a 2 × downsampling (Agisoft LLC, 2025). To finally calculate the volume of the ice blocks, the DPC was rasterized to a digital elevation model in the local coordinate system. A polygon was then created corresponding to the shape of the ice block container, and the volume of the material above the shape was calculated. Ice melt was then computed from the difference in volume in comparison with the measurements from the ice melt model.

4. Results

4.1. Ice Melt Prediction

The ice melt rates for the idealized configurations of bare, dirty, and buried ice (i.e., physically separated and thermally insulated from each other, with only their upper surfaces exposed to energy fluxes) were computed a posteriori by applying the respective energy balance models under the meteorological conditions recorded during the experimental period. The resulting melt rates are illustrated in Figure 3 and highlight the distinct behaviors associated with each surface type. Buried ice, being insulated from radiative fluxes, exhibited the lowest melt rate, corresponding to a total mass loss of 5.6 L of meltwater. Bare ice experienced an intermediate melt rate, leading to a total meltwater of 16 L (equal to about three times the value of the buried ice). In contrast, dirty ice, characterized by the lowest albedo and thus the highest absorption of incoming solar radiation, showed the greatest ice melt rate, with a total of 22 L of meltwater (equal to about four times the value of the buried ice).

4.2. Meteorological Conditions Affecting Ice Melt

During the observation period, daily air temperatures in Piazza Città di Lombardia ranged from a minimum of +9.3 °C (on 19 April) to a maximum of +24.0 °C (on 16 April), with a mean daily temperature of +15.8 °C (Figure 4). Relative humidity exhibited strong variability, with values ranging from a minimum of 11.3% (16 April) to a maximum of 75.3% (on 16 April) and an average of 40.9% over the entire period (Figure 5). These fluctuations reflect the transitional spring weather conditions typical of the region.
The experimental setup was positioned beneath the large canopy of the Pyramid structure (Figure 1), which prevented direct solar radiation from reaching the ice cubes for most of the day. As a result, only the diffuse component of incoming shortwave radiation was effectively measured. Consistently low SWin values were recorded, ranging from 0 W·m−2, which is the values recorded at night, to daytime values generally below 200 W·m−2. However, around local solar noon (approximately 12:30), the sun’s position occasionally allowed a few direct rays to penetrate the canopy and reach the radiometric sensor. On such occasions, peak shortwave radiation reached up to 772 W·m−2 (on 20 April), suggesting transient exposure to direct radiation (Figure 6).
The surrounding urban architecture also played a role in modulating the wind. The square is enclosed by tall buildings, which significantly reduced air circulation near the surface. Wind speed measurements confirm this, with a maximum of only 1.1 m·s−1 (on 16 April) and several instances of complete wind calm (0 m·s−1), especially during the early morning and evening hours. These sheltered conditions likely influenced both convective heat transfer and the overall energy balance at the ice surface.

4.3. Actual Ice Melt Quantification

We compared the modeled cumulative meltwater curve with the measured data during the week (Figure 7 and Table 1). The model predicted a total meltwater equivalent of 745 L, slightly lower than the total measured value (748 L). Analyzing the cumulative curve in Figure 7, the model tends to underestimate the loss between 18 April and 20 April, whereas it overestimates it during the final two days. Additionally, volume loss derived from photogrammetric measurements on three separate days was compared to the modeled estimates (Table 2). Between 18 April (09:00) and 19 April (16:50), the model shows a higher melt, while between 19 April (16:50) and 21 April (11:10) it shows lower losses compared to the photogrammetric data.

5. Discussion

The aim of this Urban Gaciology experiment is to raise awareness among citizens and visitors about the intensity of glacier melting, which is increasing in the Alps and worldwide as a result of the ongoing climate change (Leclercq et al., 2014; Zemp et al., 2025). Lombardy’s Alpine glaciers are featuring accelerate shrinkage rates (Hugonnet et al., 2021), that coincided with a clear local warming and a small local decrease in snow cover depth and duration which resulted strongly related to North Atlantic Oscillation (NAO) winter variability. Moreover, these glaciers currently experience consistently positive temperatures day and night for at least 60 days a year over most of their surface area, with even peaks of +15 °C during the day in July and August (data from AWS installed at the Forni Glacier, e.g., (Senese et al., 2021). Indeed, the weather conditions affecting one week in April in Milan closely resembled those typically observed on Alpine glaciers during the summer, with meteorological parameters comparable to those recorded at high elevations between June and August (Williamson et al., 2025). Thus, being able to observe it represented a unique opportunity for school pupils, teachers and citizens (Scanlon, 2014). This is particularly true for the people who may not have regular access to high-altitude environments. For those who do not frequent the mountains, witnessing the visible effects of ice melt firsthand helped to bridge the gap between distant climate processes and everyday urban life, bringing greater awareness of the ongoing changes affecting the cryosphere and their broader environmental implications. Furthermore, the phenomenon of ‘darkening’ is intensifying, which sees Alpine and worldwide glaciers becoming darker and darker due to the deposition of black carbon and detritus on their surface (Ming et al., 2012; Oerlemans et al., 2009). Thus, being able to observe the different evolution of the 3 ice blocks would have provided a better understanding of the fate of our glaciers. Although practical constraints required adapting the initial design, resulting in a single ice block with three distinct surface conditions rather than three thermally isolated cubes, the experiment successfully raised public awareness about the fragility of glaciers and the climatic processes that threaten their survival. Importantly, this urban setting allowed the experiment to reach audiences who are not usually exposed to glacial environments or scientific fieldwork. By enabling participants to witness glacier melting processes without traveling to remote mountain areas, the installation broadened accessibility and made the phenomenon tangible for a heterogeneous public. This approach not only facilitated genuine understanding of glacier loss but also encouraged participants to reflect on its broader societal implications, demonstrating the potential of urban cultural events as platforms for transformative environmental learning. The recording of the experiment using close range photogrammetry techniques made it possible, at the end of the experiment, to better appreciate the effect of the different surface conditions (clean ice, covered by black carbon or gravelly–sandy debris) and of the general weather conditions, on the melt rates, and to share all the phases and results via web-social media, and to reach other people who were not physically present in the Urban Glaciology Lab.
The Urban Glaciology experiment is grounded in quantifiable observations: the ice blocks showed significantly different melting rates (up to 20% higher for dirty ice compared to clean ice), consistent with known albedo and solar radiation models observed on Alpine glaciers at 2800 m. This evidence validates the thermal analogy between urban spring temperatures (15–18 °C) and summer ablation season conditions in mountain environments, making the experience both intuitive and scientifically sound.
The presence of multiple surface variants (i.e., clean, dirty, and debris-covered) made differences in energy balance visibly clear, enriching the public’s understanding of how surface color and material composition influence glacier melt.

5.1. Comparison with Similar Global Initiatives

This experiment aligns with international environmental outreach projects that use tangible, site-specific elements (Table 3):
Ice Watch by Olafur Eliasson: twelve Arctic ice blocks displayed in a clock-like configuration in Copenhagen (2014), then Paris and London, left to melt in public urban spaces to evoke the passage of time and the accelerated loss of ice (Cucuzzella, 2021).
The Tempestry Project, a participatory fiber art initiative visualizing annual temperature trends through color-coded knitted panels, shared in museums and public installations worldwide (Zambello, 2019).
Moreover, installations like Ghost Forest by Maya Lin in New York (Marley, 2021), or Holoscenes by Lars Jan (Jacolbe, 2025), use dead trees or simulated floods to symbolize climate impacts in city settings.
These initiatives share with Urban Glaciology the goal of creating visceral, visual experiences in public spaces to move beyond abstraction and provoke climate awareness. What distinguishes the Urban Glaciology experiment from other initiatives (e.g., Olafur Eliasson’s Ice Watch) is the integration of a structured educational path grounded in three key thematic dimensions: perception, knowledge, and awareness. Rather than offering a static visual provocation, our installation guided visitors through an experiential narrative designed to activate different levels of engagement. The first phase (i.e., perception) aimed to awaken visitors’ emotional connection to glacier loss through the direct encounter with melting ice in an unexpected urban setting. This was followed by the phase of knowledge, in which immersive reality experiences enabled participants to understand the scientific mechanisms behind glacier retreat, simulating the fieldwork and research conditions of high-altitude environments using real data and glaciological tools (Barbagallo et al., 2024). Finally, the third phase focused on awareness, as visitors were invited to measure their own contribution to climate change through an interactive carbon footprint web application, encouraging personal reflection on everyday behavior (Senese et al., 2024). Each step was supported by the active presence of researchers and university students, who provided scientific explanations, answered questions, and facilitated the entire journey. This multi-step framework transformed a fleeting visual encounter into a meaningful, personalized learning process, an approach that goes beyond symbolic gestures and positions Urban Glaciology as an example of climate communication through transformative tourism.

5.2. Interpretation and Future Potential

The approach succeeded in engaging a non-traditional audience by integrating into a tourism and cultural flow. Testimonials collected through on-site brief interviews indicated a marked increase in climate awareness: over 60% of respondents, previously unaware of the thermal analogy with Alpine glaciers, reported that the experience shifted their perception of the issue. As part of the evaluation of the immersive VR experience offered during the Urban Glaciology installation, visitors were invited to fill in a structured questionnaire previously developed to assess the effectiveness of 360° immersive tools for environmental education (Barbagallo et al., 2024). A total of 335 responses were collected, including 291 Italians and 44 internationals. Italians were mainly employees (51%) and students (26%), while the international group showed a similar profile. Gender distribution was balanced in both groups. Prior exposure to Earth science courses was reported by 36% of Italians and 64% of internationals. Experience with glaciers varied: among Italians, 24% had never seen a glacier, while 16% had walked on one; among internationals, 34% had never seen one, and 14% had walked on one. Prior knowledge was mostly acquired through media and online resources (61% of Italians; 52% of internationals), while a smaller fraction cited formal studies. Interest in climate change was very high: only 3% of Italians and none of the internationals declared no interest. These findings highlight both the diversity of the audience and the strong baseline curiosity about environmental issues, which the Urban Glaciology experiment was able to address effectively.
Compared to larger-scale examples like Ice Watch, this project added the value of a direct climate comparison between urban and mountain contexts, supported by scientific measurements and real-time interaction.
These examples demonstrate how environmental installations, when integrated into high-traffic urban events, can reach diverse audiences, make climate change experiential, and stimulate critical thinking without formal instruction. In this sense, the Urban Glaciology experiment was deliberately designed to intercept audiences beyond those already sensitive to environmental issues. By situating the installation in the heart of a major cultural and design event, we were able to engage individuals who might not actively seek out environmental content, such as international tourists, design professionals, and urban passers-by. The melting ice blocks served as an immediate visual trigger, while the presence of researchers and students provided a bridge from initial curiosity to informed reflection. This dual approach (i.e., emotional impact followed by scientific explanation) allowed visitors from diverse backgrounds to gain a genuine perception of glacier melting, encouraging them to relate the phenomenon to their own lives and to consider its broader implications. Such engagement with non-traditional audiences is essential for widening the social reach of climate change communication and for fostering collective reflection on the shared responsibility to address its impacts.
The Urban Glaciology experiment also illustrates how climate change communication can be embedded in tourism flows. By situating the installation within a major cultural and design event, the initiative directly engaged tourists and international visitors who were exploring Milan for leisure rather than for scientific or environmental purposes. In this way, the experiment broadened the scope of climate change education beyond traditional audiences and transformed a cultural tourism experience into an occasion for environmental learning. This demonstrates that tourism can act not only as a sector highly exposed to climate risks, such as glacier retreat in mountain destinations, but also as a vehicle for raising awareness and fostering dialogue on global environmental change in urban contexts.
Beyond the urban public space, schools could represent a valuable arena for similar outreach initiatives. Bringing a scaled or adapted version of the Urban Glaciology experiment into educational settings would foster early awareness among students and teachers, providing opportunities to integrate experiential climate change education within formal learning environments.

6. Conclusions

The Urban Glaciology experiment demonstrated that climate change education can be effectively integrated into large-scale urban cultural events, offering a novel strategy to engage a wide and heterogeneous audience. The Fuorisalone, traditionally associated with design, creativity, and innovation, provided an unexpected yet ideal platform for environmental awareness. By simulating glacier melting through real ice blocks in an urban spring environment, the installation offered a tangible and intuitive representation of climate change impacts, particularly for individuals who may never visit high-altitude glacial areas.
This public engagement initiative succeeded in capturing the attention of a tourist-driven and culturally curious audience, composed largely of design enthusiasts, international visitors, and urban explorers. Many visitors reported surprise when learning that the spring temperatures in Milan closely resemble summer conditions on Alpine glaciers, and that glacier melt occurs just as rapidly (albeit invisibly) in remote high-altitude settings. This perceptual shift is a key outcome of the installation: by making glacier melting visible and local, the project helped bridge the psychological and geographic distance between urban dwellers and vulnerable mountain environments.
The Urban Glaciology experiment demonstrated that climate change education can be successfully integrated into large-scale urban cultural events, offering a new paradigm for reaching broad and diverse audiences. Hosted during the Fuorisalone 2024 in Milan (i.e., an international design and tourism event), the installation transformed a busy urban square into an interactive learning environment where visitors could observe glacier melting in real time, without needing to access remote mountain environments.
The installation also acted as a form of experiential learning and informal science education, where knowledge was not transmitted through lectures or digital media, but rather emerged from observation, curiosity, and interaction with researchers on-site. This approach aligns with concepts of transformative tourism and environmentally responsible travel, where experiences are designed to provoke critical reflection and potentially inspire behavioral change.
Crucially, the experiment succeeded in intercepting a “non-traditional” audience (e.g., design enthusiasts, urban tourists, and international visitors), many of whom would not normally engage with scientific content or visit glacial areas. This highlights the untapped potential of cultural and tourism events as strategic platforms for informal science education and environmental awareness.
The initiative also provided educational value for students and young researchers and generated useful data through simple yet robust scientific methods, including photogrammetry and energy balance modeling. These tools ensured scientific rigor while maintaining accessibility and inclusiveness.
As cities increasingly seek to promote sustainability and resilience, initiatives like Urban Glaciology offer a replicable model for embedding climate engagement within urban tourism flows. Whether through melting ice blocks, immersive media, or experiential installations, bringing science into public spaces can turn temporary events into lasting environmental encounters.
In conclusion, by placing a melting cube of ice in the heart of Milan’s design week, this initiative transformed a cultural event into a climate classroom, demonstrating that even fleeting encounters in urban public spaces can foster lasting environmental awareness.

6.1. Theoretical Implications

The findings of this study contribute to the growing body of literature on informal and transformative environmental education. By demonstrating that glacier melting can be effectively communicated through an urban installation, the Urban Glaciology experiment confirms the potential of cultural and tourism events as non-traditional arenas for climate change communication (Gaspari et al., 2025; Newman, 2018). It advances theoretical discussions on experiential learning by showing that emotional engagement (through direct observation of melting ice) and cognitive understanding (through interaction with scientists and VR technologies) can be combined to foster reflective learning. Moreover, the study reinforces the theoretical link between tourism studies and climate communication, positioning cultural events not only as spaces of leisure and creativity but also as platforms for environmental awareness and collective reflection.

6.2. Practical Implications

From a practical perspective, the experiment demonstrates a replicable model for embedding science-based installations in urban cultural events. Practitioners and event organizers can adapt similar approaches to reach audiences beyond those traditionally engaged in environmental issues, such as tourists, designers, and casual passers-by. For educators, the integration of immersive media and citizen science tools offers an effective means to contextualize climate change and glacier retreat within familiar urban settings. For policymakers, the initiative illustrates how temporary cultural events can be leveraged as part of broader sustainability and climate communication strategies in cities. By fostering accessibility, inclusiveness, and experiential learning, such initiatives can complement formal education and contribute to building public awareness and resilience in the face of climate change.

Author Contributions

Conceptualization, A.S., M.M. and G.A.D.; methodology, A.S., C.D.A., D.F., M.M. and G.A.D.; software, C.D.A. and D.F.; validation, A.S., C.D.A., D.F., M.M. and G.A.D.; formal analysis, A.S., C.D.A., D.F., M.M. and G.A.D.; investigation, A.S., D.F., B.B., L.C. and G.A.D.; resources, A.S., D.F., B.B., L.C. and G.A.D.; data curation, A.S., D.F., B.B., L.C. and G.A.D.; writing—original draft preparation, A.S., C.D.A., D.F., M.M. and G.A.D.; writing—review and editing, A.S. and G.A.D.; visualization, A.S., C.D.A., D.F., M.M. and G.A.D.; supervision, A.S., M.M. and G.A.D.; project administration, A.S., M.M. and G.A.D.; funding acquisition, A.S., M.M. and G.A.D. All authors have read and agreed to the published version of the manuscript.

Funding

This research received no external funding.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

The original contributions presented in this study are included in the article. Further inquiries can be directed to the corresponding author.

Acknowledgments

This study was carried out in the framework of the “Lombardy on the roof of the world” lab managed by EvK2CNR during the “FuoriSalone” event of the Milan Design Week 2024. These kinds of activities are under the frame of public engagement activities of the Environmental Science and Policy Department of the University of Milan. Researchers involved in the study were also supported by Sanpellegrino Levissima S.p.A. and Stelvio National Park (ERSAF).

Conflicts of Interest

The authors declare no conflicts of interest.

References

  1. Agisoft LLC. (2025). Agisoft metashape user manual professional edition (Version 2.2) [Computer software]. Agisoft LLC.
  2. Balduini, M., Della Valle, E., Dell’Aglio, D., Tsytsarau, M., Palpanas, T., & Confalonieri, C. (2013). Twindex fuorisalone: Social listening of milano during fuorisalone 2013. In The semantic web: ESWC 2013 satellite events (Vol. 7955). Lecture Notes in Computer Science. Springer. [Google Scholar] [CrossRef]
  3. Barbagallo, B., Senese, A., Ambrosini, R., Cresi, L., Franceschini, M., Almagioni, C. D., Arcuri, B., Crosta, A., De Rosa, S., Litholdo, T., Lopopolo, S., Manara, V., Melada, J., Montemurro, A. M., Spezza, A., Maugeri, M., & Diolaiuti, G. A. (2024). Immersive 360° experiences for geoscience learning. In L. T. De Paolis, P. Arpaia, & M. Sacco (Eds.), Extended reality (Vol. 15030, pp. 161–178). Springer Nature. [Google Scholar] [CrossRef]
  4. Barrett, M. J., Harmin, M., Maracle, B., Patterson, M., Thomson, C., Flowers, M., & Bors, K. (2017). Shifting relations with the more-than-human: Six threshold concepts for transformative sustainability learning. Environmental Education Research, 23(1), 131–143. [Google Scholar] [CrossRef]
  5. Bonasoni, P., Laj, P., Marinoni, A., Sprenger, M., Angelini, F., Arduini, J., Bonafè, U., Calzolari, F., Colombo, T., Decesari, S., Di Biagio, C., Di Sarra, A. G., Evangelisti, F., Duchi, R., Facchini, M. C., Fuzzi, S., Gobbi, G. P., Maione, M., Panday, A., … Cristofanelli, P. (2010). Atmospheric brown clouds in the Himalayas: First two years of continuous observations at the Nepal Climate Observatory-Pyramid (5079 m). Atmospheric Chemistry and Physics, 10(15), 7515–7531. [Google Scholar] [CrossRef]
  6. Capari, L., Wilfing, H., Exner, A., Höflehner, T., & Haluza, D. (2022). Cooling the city? A scientometric study on urban green and blue infrastructure and climate change-induced public health effects. Sustainability, 14(9), 4929. [Google Scholar] [CrossRef]
  7. Carturan, L., Cazorzi, F., De Blasi, F., & Dalla Fontana, G. (2015). Air temperature variability over three glaciers in the Ortles-Cevedale (Italian Alps): Effects of glacier fragmentation, comparison of calculation methods, and impacts on mass balance modeling. Cryosphere, 9(3), 1129–1146. [Google Scholar] [CrossRef]
  8. Chu, E., & Schenk, T. (2017). Communicating about climate change with urban populations and decision-makers. In Oxford research encyclopedia of climate science. Oxford University Press. [Google Scholar] [CrossRef]
  9. Cook, D., Malinauskaite, L., Davíðsdóttir, B., & Ögmundardóttir, H. (2021). Co-production processes underpinning the ecosystem services of glaciers and adaptive management in the era of climate change. Ecosystem Services, 50, 101342. [Google Scholar] [CrossRef]
  10. Cucuzzella, C. (2021, October 26). Eco-art, design and architecture can be agents of environmental change in the public realm. SocialCommon.Ca. Available online: https://socialcommons.ca/2021/10/26/eco-art-design-and-architecture-can-be-agents-of-environmental-change-in-the-public-realm (accessed on 25 July 2025).
  11. Cuffey, K. M., & Paterson, W. S. B. (2010). The physics of glaciers (4th ed.). Academic Press. [Google Scholar]
  12. Dombrowski, O., Franssen, H. J. H., Brogi, C., & Bogena, H. R. (2021). Performance of the ATMOS41 all-in-one weather station for weather monitoring. Sensor, 21(3), 741. [Google Scholar] [CrossRef]
  13. Gagné, K., Rasmussen, M. B., & Orlove, B. (2014). Glaciers and society: Attributions, perceptions, and valuations. Wiley Interdisciplinary Reviews: Climate Change, 5(6), 793–808. [Google Scholar] [CrossRef]
  14. Gaspari, F., Barbieri, F., Fascia, R., Ioli, F., Pinto, L., & Migliaccio, F. (2025). Strategies for glacier retreat communication with 3D geovisualization and open data sharing. ISPRS International Journal of Geo-Information, 14(2), 75. [Google Scholar] [CrossRef]
  15. Gramegna, S. M., Mattioli, F., & Pei, X. (2024). What Italian furniture companies do towards sustainable transition? Design actions and strategies showcased during Milan Design Week 2023. Proceedings of the Design Society, 4, 1269–1278. [Google Scholar] [CrossRef]
  16. Hartmann, D. L. (2015). Global physical climatology (2nd ed.). Elsevier Inc. [Google Scholar] [CrossRef]
  17. Hugonnet, R., McNabb, R., Berthier, E., Menounos, B., Nuth, C., Girod, L., Farinotti, D., Huss, M., Dussaillant, I., Brun, F., & Kääb, A. (2021). Accelerated global glacier mass loss in the early twenty-first century. Nature, 592, 726–731. [Google Scholar] [CrossRef]
  18. IPCC. (2023). Climate change 2023: Synthesis report. IPCC. [Google Scholar]
  19. Jacolbe, J. (2025, April 22). Lars Jan’s art goes underwater as warning of climate crisis rises. ALL ARTS. Available online: https://www.allarts.org/2025/04/lars-jan-holoscenes/ (accessed on 25 July 2025).
  20. Koutlemanis, P., Zabulis, X., Stivaktakis, N., Partarakis, N., Zidianakis, E., & Demeridou, I. (2024). A low-cost close-range photogrammetric surface scanner. Frontiers in Imaging, 3, 1341343. [Google Scholar] [CrossRef]
  21. Leclercq, P. W., Oerlemans, J., Basagic, H. J., Bushueva, I., Cook, A. J., & Le Bris, R. (2014). A data set of worldwide glacier length fluctuations. Cryosphere, 8(2), 659–672. [Google Scholar] [CrossRef]
  22. Markowitz, E. M., & Guckian, M. L. (2018). Climate change communication: Challenges, insights, and opportunities. In Psychology and climate change: Human perceptions, impacts, and responses (pp. 35–63). Elsevier Academic Press. [Google Scholar]
  23. Marley, M. (2021, May 10). In maya lin’s new installation, a forest grows in Manhattan. Vogue. Available online: https://www.vogue.com/article/maya-lin-ghost-forrest (accessed on 25 July 2025).
  24. Ming, J., Du, Z., Xiao, C., Xu, X., & Zhang, D. (2012). Darkening of the mid-Himalaya glaciers since 2000 and the potential causes. Environmental Research Letters, 7(1), 014021. [Google Scholar] [CrossRef]
  25. Moreau, M. (2010). Visual perception of changes in a high mountain landscape: The case of the retreat of the Évettes Glacier (Haute-Maurienne, northern French Alps). Geomorphologie: Relief, Processus, Environnement, 16(2), 165–174. [Google Scholar] [CrossRef]
  26. Moser, S. C. (2006). Talk of the city: Engaging urbanites on climate change. Environmental Research Letters, 1(1), 14006–14010. [Google Scholar] [CrossRef]
  27. Moser, S. C. (2013). Navigating the political and emotional terrain of adaptation: Community engagement when climate change comes home. In Successful adaptation to climate change: Linking science and policy in a rapidly changing world. Routledge. [Google Scholar]
  28. Moser, S. C. (2014). Communicating adaptation to climate change: The art and science of public engagement when climate change comes home. Wiley Interdisciplinary Reviews: Climate Change, 5(3), 337–358. [Google Scholar] [CrossRef]
  29. Nakawo, M., & Takahashi, S. (1982). A simplified model for estimating glacier ablation under a debris layer. In Hydrological aspects of alpine and high mountain areas. Proceedings of the exeter symposium, July 1982. Penn State University Press. [Google Scholar]
  30. Newman, C. P. (2018). Geoscience education and public outreach in Rocky Mountain National Park, Colorado, U.S.A. Rocky Mountain Geology, 53(1), 45–55. [Google Scholar] [CrossRef]
  31. Oerlemans, J., Giesen, R. H., & Van Den Broeke, M. R. (2009). Retreating alpine glaciers: Increased melt rates due to accumulation of dust (Vadret da Morteratsch, Switzerland). Journal of Glaciology, 55(192), 729–736. [Google Scholar] [CrossRef]
  32. Palomo, I. (2017). Climate change impacts on ecosystem services in high mountain areas: A literature review. Mountain Research and Development, 37(2), 179–187. [Google Scholar] [CrossRef]
  33. Robel, A. A., Ultee, L., Ranganathan, M., & Nash, M. (2024). For whom and by whom is glaciology? Journal of Glaciology, 70, e68. [Google Scholar] [CrossRef]
  34. Saladié, Ò., Boqué-Ciurana, A., Sevil, J., & Olano Pozo, J. X. (2025). Summer thermal comfort in urban squares: The case of human tower exhibitions in Catalonia. Atmosphere, 16(6), 666. [Google Scholar] [CrossRef]
  35. Salerno, F., Guyennon, N., Yang, K., Shaw, T. E., Lin, C., Colombo, N., Romano, E., Gruber, S., Bolch, T., Alessandri, A., Cristofanelli, P., Putero, D., Diolaiuti, G., Tartari, G., Verza, G., Thakuri, S., Balsamo, G., Miles, E. S., & Pellicciotti, F. (2023). Local cooling and drying induced by Himalayan glaciers under global warming. Nature Geoscience, 16(12), 1120–1127. [Google Scholar] [CrossRef]
  36. Scanlon, E. (2014). Scholarship in the digital age: Open educational resources, publication and public engagement. British Journal of Educational Technology, 45(1), 12–23. [Google Scholar] [CrossRef]
  37. Schaefli, B., Manso, P., Fischer, M., Huss, M., & Farinotti, D. (2019). The role of glacier retreat for Swiss hydropower production. Renewable Energy, 132, 615–627. [Google Scholar] [CrossRef]
  38. Senese, A., Caspani, A. C., Lombardo, L., Manara, V., Diolaiuti, G. A., & Maugeri, M. (2024). A user-friendly tool to increase awareness about impacts of human daily life activities on carbon footprint. Sustainability, 16(5), 1976. [Google Scholar] [CrossRef]
  39. Senese, A., Fugazza, D., Chiatante, G., Kaihoh, K., Trasatti, S., Maugeri, M., & Diolaiuti, G. (2023a). Indoor glaciology: “Melting Landscape” to increase people awareness on climate change impacts. Geografia Fisica e Dinamica Quaternaria, 46, 153–162. [Google Scholar]
  40. Senese, A., Leidi, M., & Diolaiuti, G. (2021). A new enhanced temperature-index melt model including net solar and infrared radiation. Geografia Fisica e Dinamica Quaternaria, 44(1), 5–14. [Google Scholar] [CrossRef]
  41. Senese, A., Maugeri, M., Ferrari, S., Confortola, G., Soncini, A., Bocchiola, D., & Diolaiuti, G. (2016). Modelling shortwave and longwave downward radiation and air temperature driving ablation at the Forni Glacier (Stelvio National Park, Italy). Geografia Fisica e Dinamica Quaternaria, 39(1), 89–100. [Google Scholar] [CrossRef]
  42. Senese, A., Pecci, M., Ambrosini, R., & Diolaiuti, G. A. (2023b). MOUNTAINPLAST: A new Italian plastic footprint with a focus on mountain activities. Sustainability, 15(8), 7017. [Google Scholar] [CrossRef]
  43. Singer-Brodowski, M. (2023). The potential of transformative learning for sustainability transitions: Moving beyond formal learning environments. Environment, Development and Sustainability. [Google Scholar] [CrossRef]
  44. Stewart, E. J., Wilson, J., Espiner, S., Purdie, H., Lemieux, C., & Dawson, J. (2016). Implications of climate change for glacier tourism. Tourism Geographies, 18(4), 377–398. [Google Scholar] [CrossRef]
  45. Wamsler, C., Brink, E., & Rivera, C. (2013). Planning for climate change in urban areas: From theory to practice. Journal of Cleaner Production, 50, 68–81. [Google Scholar] [CrossRef]
  46. Williamson, S. N., Marshall, S. J., & Menounos, B. (2025). Temperature mediated albedo decline portends acceleration of North American glacier mass loss. Communications Earth & Environment, 6(1), 555. [Google Scholar] [CrossRef]
  47. Zambello, E. (2019, April 3). Crafters work to create the national park tempestry project. National Parks Traveler. Available online: https://www.nationalparkstraveler.org/2019/04/crafters-work-create-national-park-tempestry-project (accessed on 25 July 2025).
  48. Zemp, M., Jakob, L., Dussaillant, I., Nussbaumer, S. U., Gourmelen, N., Dubber, S., A, G., Abdullahi, S., Andreassen, L. M., Berthier, E., Bhattacharya, A., Blazquez, A., Boehm Vock, L. F., Bolch, T., Box, J., Braun, M. H., Brun, F., Cicero, E., Colgan, W., … Zheng, W. (2025). Community estimate of global glacier mass changes from 2000 to 2023. Nature, 639, 382–388. [Google Scholar] [CrossRef] [PubMed]
Figure 1. The Urban Glaciology experiment. (a) The pyramid (17 × 17 × 11 m) of the “Lombardy on the roof of the world” Lab during the “FuoriSalone” event of the Milan Design Week; (b) the ice blocks (each 0.5 m3 in volume) with the AWS and the water container used to collect meltwater for measurements inside the pyramid; (c) The different kinds of ice surfaces.
Figure 1. The Urban Glaciology experiment. (a) The pyramid (17 × 17 × 11 m) of the “Lombardy on the roof of the world” Lab during the “FuoriSalone” event of the Milan Design Week; (b) the ice blocks (each 0.5 m3 in volume) with the AWS and the water container used to collect meltwater for measurements inside the pyramid; (c) The different kinds of ice surfaces.
Tourismhosp 06 00168 g001
Figure 2. Conceptual scheme of the Urban Glaciology experiment. (Left): Ideal setup, with three thermally isolated ice blocks (bare, dirty, and debris-covered) analyzed through energy balance models (Equations (1)–(3)). (Right): Experimental setup adopted during the Fuorisalone 2024 event, where the three blocks were placed in a single transparent container. Meteorological data were continuously recorded by the AWS (air temperature, humidity, wind, pressure, solar radiation), while melt was quantified through daily height measurements, photogrammetry, and direct collection of meltwater.
Figure 2. Conceptual scheme of the Urban Glaciology experiment. (Left): Ideal setup, with three thermally isolated ice blocks (bare, dirty, and debris-covered) analyzed through energy balance models (Equations (1)–(3)). (Right): Experimental setup adopted during the Fuorisalone 2024 event, where the three blocks were placed in a single transparent container. Meteorological data were continuously recorded by the AWS (air temperature, humidity, wind, pressure, solar radiation), while melt was quantified through daily height measurements, photogrammetry, and direct collection of meltwater.
Tourismhosp 06 00168 g002
Figure 3. Melt rate of different ice surfaces: bare ice (blue), buried ice (red) and dirty ice (green) in the idealized configurations (i.e., physically separated and thermally insulated from each other, with only their upper surfaces exposed to energy fluxes).
Figure 3. Melt rate of different ice surfaces: bare ice (blue), buried ice (red) and dirty ice (green) in the idealized configurations (i.e., physically separated and thermally insulated from each other, with only their upper surfaces exposed to energy fluxes).
Tourismhosp 06 00168 g003
Figure 4. Air temperature measured by AWS during the “FuoriSalone” event. Data represent hourly averages.
Figure 4. Air temperature measured by AWS during the “FuoriSalone” event. Data represent hourly averages.
Tourismhosp 06 00168 g004
Figure 5. Relative humidity recorded by AWS during the “FuoriSalone” event. Data represent hourly averages.
Figure 5. Relative humidity recorded by AWS during the “FuoriSalone” event. Data represent hourly averages.
Tourismhosp 06 00168 g005
Figure 6. Incoming shortwave radiation (SWin) measured by AWS during the “FuoriSalone” event. Data represent hourly averages.
Figure 6. Incoming shortwave radiation (SWin) measured by AWS during the “FuoriSalone” event. Data represent hourly averages.
Tourismhosp 06 00168 g006
Figure 7. Cumulative ice melting modeled for the actual conditions (blue line) and measured (red dots).
Figure 7. Cumulative ice melting modeled for the actual conditions (blue line) and measured (red dots).
Tourismhosp 06 00168 g007
Table 1. Total daily values of ice melt modeled for the actual conditions and measured during the experiment and their difference.
Table 1. Total daily values of ice melt modeled for the actual conditions and measured during the experiment and their difference.
DaysModeled (L)Measured (L)Difference (L)
15 April (18.30)4350−7
16 April (22.30)2342304
17 April (20.00)103128−25
18 April (19.00)8698−12
19 April (20.00)100102−2
20 April (20.00)11310013
21 April (17.00)674027
Table 2. Ice melt derived from photogrammetric reconstruction compared to modeled values in two different periods.
Table 2. Ice melt derived from photogrammetric reconstruction compared to modeled values in two different periods.
Photogrammetry (L)Modeled (L)
18 April 9.00–19 April 16.50−124−132
19 April 16.50–21 April 11.10−213−171
Table 3. Comparison between events, educational initiatives and visitors reached.
Table 3. Comparison between events, educational initiatives and visitors reached.
Event/InstallationLocationEducational Activity HostedEstimated Visitors
Urban Glaciology (this work)Milan (Fuorisalone 2024)Ice melt simulation in public square~2000 (entire square)
Ice Watch (first edition) https://socialcommons.ca/2021/10/26/eco-art-design-and-architecture-can-be-agents-of-environmental-change-in-the-public-realm/ (accessed on 25 July 2025)Copenhagen 2014Arctic ice blocks arranged as a clockover 100,000
Ice Watch (Paris edition) https://socialcommons.ca/2021/10/26/eco-art-design-and-architecture-can-be-agents-of-environmental-change-in-the-public-realm/ (accessed on 25 July 2025)Paris 2015Same, during COP21tens of thousands
Tempestry Project https://www.nationalparkstraveler.org/2019/04/crafters-work-create-national-park-tempestry-project (accessed on 25 July 2025)global sitesKnitted temperature records displayedhundreds per venue
Ghost Forest (https://www.vogue.com/article/maya-lin-ghost-forrest) (accessed on 25 July 2025)New York (Madison Sq. Park)Dead trees as symbolic foresttens of thousands
Holoscenes (https://www.allarts.org/2025/04/lars-jan-holoscenes/) (accessed on 25 July 2025)NYC, Brisbane, LondonLive performance inside a flooding glass boxthousands (public shows)
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Senese, A.; Almagioni, C.D.; Fugazza, D.; Barbagallo, B.; Cresi, L.; Maugeri, M.; Diolaiuti, G.A. Tourism, Design and Climate Change: The Urban Glaciology Experiment at Fuorisalone 2024 Event. Tour. Hosp. 2025, 6, 168. https://doi.org/10.3390/tourhosp6040168

AMA Style

Senese A, Almagioni CD, Fugazza D, Barbagallo B, Cresi L, Maugeri M, Diolaiuti GA. Tourism, Design and Climate Change: The Urban Glaciology Experiment at Fuorisalone 2024 Event. Tourism and Hospitality. 2025; 6(4):168. https://doi.org/10.3390/tourhosp6040168

Chicago/Turabian Style

Senese, Antonella, Cecilia D. Almagioni, Davide Fugazza, Blanka Barbagallo, Lorenzo Cresi, Maurizio Maugeri, and Guglielmina A. Diolaiuti. 2025. "Tourism, Design and Climate Change: The Urban Glaciology Experiment at Fuorisalone 2024 Event" Tourism and Hospitality 6, no. 4: 168. https://doi.org/10.3390/tourhosp6040168

APA Style

Senese, A., Almagioni, C. D., Fugazza, D., Barbagallo, B., Cresi, L., Maugeri, M., & Diolaiuti, G. A. (2025). Tourism, Design and Climate Change: The Urban Glaciology Experiment at Fuorisalone 2024 Event. Tourism and Hospitality, 6(4), 168. https://doi.org/10.3390/tourhosp6040168

Article Metrics

Back to TopTop