Next Article in Journal
A Comparative Analysis of Circular Economy Index in Urban and Rural Municipalities
Previous Article in Journal
Towards Harmonized Reduction of Seismic Vulnerability: Analyzing Regulatory and Incentive Frameworks in the Adriatic—Ionian Region
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

The Vicious Cycle Atlas of Fragility: Mapping the Feedback Loops Between Industrial–Urban Metabolism and Earth System Collapse

1
Faculty of Economics, Keio University, Mita Minato-ku, Tokyo 108-8345, Japan
2
Institute for Environment and Development (LESTARI), National University of Malaysia, Lingkungan Ilmu, Bangi 43600, Selangor, Malaysia
Urban Sci. 2025, 9(8), 320; https://doi.org/10.3390/urbansci9080320
Submission received: 7 July 2025 / Revised: 7 August 2025 / Accepted: 11 August 2025 / Published: 14 August 2025

Abstract

This study examines how Multi-Scalar Nature-Based Regenerative Solutions (M-NbRS) can realign urban–industrial systems with planetary boundaries to mitigate Earth system destabilization. Using integrated systems analysis, we document three key findings: (1) global material flows show only 9% circularity amid annual extraction of 100 billion tons of resources; (2) Earth system diagnostics reveal 28 trillion tons of cryosphere loss since 1994 and 372 Zettajoules of oceanic heat accumulation; and (3) meta-analysis identifies accelerating biosphere integrity loss (61.56 million hectares deforested since 2001) and atmospheric CO2 concentrations reaching 424.61 ppm (2024). Our Vicious Cycle Atlas of Fragility framework maps three synergistic disintegration pathways: metabolic overload from linear resource flows exceeding sink capacity, entropic degradation through high-entropy waste driving cryospheric collapse, and planetary boundary transgression. The M-NbRS framework counters these through spatially nested interventions: hyper-local urban tree canopy expansion (demonstrating 0.4–12 °C cooling), regional initiatives like the Heart of Borneo’s 24 million-hectare conservation, and global industrial controls maintaining aragonite saturation (Ωarag > 2.75) for marine resilience. Implementation requires policy innovations including deforestation-free supply chains, sustainability-linked financing, and ecological reciprocity legislation. These findings provide an evidence base for transitioning industrial–urban systems from drivers of Earth system fragility to architects of regeneration within safe operating spaces. Collectively, these findings demonstrate that M-NbRS offer a scientifically grounded, policy-actionable framework for breaking the vicious cycles of Earth system destabilization. By operationalizing nature-based regeneration across spatial scales—from street trees to transboundary conservation—this approach provides measurable pathways to realign human systems with planetary boundaries, offering a timely blueprint for industrial–urban transformation within ecological limits.

1. Introduction

Industrial–urban systems refer to the co-evolved networks where dense human settlements, industrial production, and infrastructure interact metabolically—importing low-entropy resources (energy, materials) while exporting high-entropy waste (emissions, pollution). These systems generate 80% of global GDP [1], yet consume 100 billion tons of materials annually [2] and contribute 70% of anthropogenic GHG emissions [3,4]. This creates a sustainability paradox: while fueling socio-economic progress, they erode the very biophysical foundations—climate regulation, nutrient cycling, biodiversity—that sustain them.
Despite extensive research on sustainable industrial–urban development, current approaches remain methodologically fragmented and epistemologically reductionist, as seen in three persistent patterns: (1) techno-centric circular economy models, (2) localized greening without systemic boundaries, and (3) metrics ignoring telecoupling (analyzed in depth in Section 2). This fragmentation systematically overlooks telecoupled environmental impacts induced by distant ecological degradation, with documented manifestations including tropical deforestation exceeding 60 million hectares (ha) since 2001, principally driven by agricultural commodity supply chains servicing urban demand [5,6]. In addition, industrial emissions drive ocean acidification, increasing ocean acidity by approximately 30% since the pre-industrial era through atmospheric carbon absorption [7]. A significant analytical gap persists in framing cities as discrete entities rather than interconnected nodes within planetary biogeochemical systems.
This oversight prevents sustainability frameworks from addressing the industrial–urban metabolic crisis—linear extraction, energy-intensive transformation, and entropy export—that destabilizes planetary boundaries. Compounding this deficiency, urban climate strategies often develop within siloed policy frameworks, disconnecting mitigation from adaptation. For instance, carbon emissions reductions, renewable energy transitions, and carbon capture technologies advance independently of flood-resistant infrastructure, heat-resilient urban design, and disaster preparedness planning. This frequently generates counterproductive outcomes that undermine integrated solutions capable of simultaneously lowering emissions, building resilience, and addressing socioeconomic inequities.
Parallel to this shortfall, it seems that no integrated framework exists to synchronize industrial–urban systems with planetary boundaries while addressing global environmental risks—such as cryospheric instability or atmospheric disturbances—alongside localized environmental degradation. This study bridges these gaps through Multi-Scalar Nature-Based Regenerative Solutions (M-NbRS), a novel framework grounded in complex systems theory and industrial ecology, which is operationalized across three interdependent tiers:
  • Local interventions: Deployment of urban nature-based solutions such as green infrastructure, permeable surfaces, constructed wetlands, and expanded urban forest canopy to enhance ecosystem functions.
  • Regional stewardship: Transboundary governance for hinterland ecosystem management. This includes metropolitan-funded watershed conservation to secure hydrological services, biodiversity corridor maintenance enabling genetic exchange, and agricultural landscape diversification supporting pollination services.
  • International policy architectures aim to stabilize the global system by aligning industrial–urban activity with planetary boundaries. Key mechanisms include supply chain due diligence protocols to prevent deforestation in critical carbon sinks, high-seas marine protected areas to preserve oceanic carbon sequestration, and cryosphere conservation treaties to mitigate albedo loss.
This study addresses:
  • How do industrial–urban metabolic processes such as resource extraction and waste generation transgress planetary boundaries across spatial scale?
  • What feedback mechanisms, for example, cryosphere albedo loss and deforestation-driven carbon sink degradation, amplify Earth system destabilization?
  • Can multi-scale nature-based regenerative solutions (M-NbRSs) synchronize industrial–urban systems with planetary boundaries while delivering socio-ecological co-benefits?
  • Formalized testable hypotheses in the conceptual underpinning (Section 4).
We hypothesize that:
  • Global high-entropy resource throughput, which exceeds 100 billion tons a year, correlates with transgression of planetary boundaries including biosphere integrity, hydrosphere stability, and atmospheric or climate resilience.
  • Amplifying feedback loops: for example, permafrost thaw accelerates global warming through positive feedback loops, releasing methane, a greenhouse gas (GHG) that is 28 times more potent than CO2 over 100 years.
  • M-NbRS implementation contributes to reducing high-entropy waste such as GHGs and pollution while enhancing cross-scale resilience, for example, urban cooling and carbon sequestration.
The M-NbRS framework reimagines local actions as interconnected nodes within a multi-scalar system, where hyper-local interventions, such as green infrastructure and urban tree canopies, are explicitly designed to cumulatively support regional ecosystem services, such as watershed protection, and global boundary stability, such as carbon sequestration. This nested design is operationalized through:
  • Coupling spatial scales: Local green infrastructure, such as permeable surfaces, directly funds regional conservation, such as watersheds, via fiscal mechanisms like sustainability-linked bonds (Section 7.1).
  • Binding policies to boundaries: Global protocols, such as deforestation-free supply chains, enforce compliance with biophysical thresholds, for example, Ωarag > 2.75, an integration absent in prior urban sustainability frameworks.
This discussion is framed by six central concepts: industrial metabolism, urban metabolism, entropy paradox, Earth system collapse, and planetary theory. Industrial metabolism, specifically, examines the systematic flow and transformation of materials and energy through industrial systems—from extraction and processing to consumption and disposal—to understand and improve sustainability and efficiency [8,9]. This framework reveals how industrial–urban material flows generate high-entropy outputs—industrial pollution, solid waste, and greenhouse gases (GHGs)—with profound environmental consequences [10,11,12]. Urban metabolism broadens this perspective by modeling cities as living organisms characterized by material inflows such as water, food, and energy and outflows such as sewage, emissions, and waste. This urban model highlights two core aspects: resource dependency, where cities import vast energy and materials but export high entropy as pollution and waste heat; and predominantly linear “take–make–waste” flows, contrasting with circular systems [13,14].
While this approach measures industrial–urban sustainability and resilience through resource exchange dynamics, the entropy paradox highlights a fundamental tension. It arises from the conflict between localized order generation in human systems such as industrial production and urban consumption, and the universal increase in high entropy dictated by the Second Law of Thermodynamics. Thermodynamically, entropy measures disorder or irreversible energy degradation—where higher entropy indicates greater system instability [15,16,17].
Earth system collapse signifies a catastrophic, irreversible breakdown in Earth’s linked biophysical systems including the atmosphere, hydrosphere, biosphere, and cryosphere, driving transitions to less habitable planetary states. It is characterized by three critical features: (1) tipping points that induce nonlinear, self-amplifying shifts, for example, the Amazon rainforest dieback; (2) cross-system cascades of failure; and (3) irreversibility on human-relevant timescales, exemplified by ice sheet collapse. Such collapse emerges from breached planetary boundaries and interconnected tipping element dynamics [18,19,20].
Planetary boundaries (PBs) define safe operating space for humanity by quantifying nine biophysical thresholds critical to Earth system stability [21]. These boundaries—including climate change, biosphere integrity (biodiversity loss), land system change, freshwater use, biochemical flows biogeochemical cycles of nitrogen and phosphorus (N/P cycles), ocean acidification, stratospheric ozone depletion, atmospheric aerosol loading, and novel entities such as pollutants—represent irreversible tipping points. Transgressing any boundary risks cascading Earth system collapse. This framework shifts sustainability from isolated local goals to a planetary-scale imperative, emphasizing that industrial–urban systems must operate within globally quantified ecological limits to avoid destabilizing feedback such as ice-albedo loss, which amplifies global warming and glacier retreat.
To counter Earth system destabilization—cryospheric collapse, ocean acidification, and biome degradation—and achieve industrial–urban resilience, metabolic restructuring through Multi-Scalar Nature-Based Regenerative Solutions (M-NbRSs) is imperative. M-NbRSs resolve the entropy paradox by transforming industries and cities from linear consumers into symbiotic stewards of biogeochemical cycles. Synchronized implementation across scales
  • Halts high-entropy waste externalization (GHG emissions, pollution);
  • Realigns human activity with planetary boundaries;
  • Prevents cascading tipping points;
  • Delivers equitable resilience with socio-ecological co-benefits.
The key contributions of the present study are
  • Theoretical innovation
    »
    An original framework to integrate urban–industrial metabolism into planetary boundaries, redefining sustainability as a quantifiable planetary obligation [21].
    »
    Transcends urban-centric NbSs by scaling ecological stewardship from local to global.
  • Critical gap resolution
    »
    Exposes telecoupled impacts, for example, urban consumption driving distant deforestation/ocean acidification.
    »
    Challenges fragmented approaches that prioritize local greening over systemic interdependencies.
  • Systems reframing
    »
    Repositions cities as symbiotic custodians of Earth’s systems, enforcing accountability via planetary boundaries.
  • Actionable roadmap
    Proposes tiered strategies to align urban–industrial metabolism with planetary boundaries, translating global thresholds into actionable scales:
    »
    Local: mandate regenerative infrastructure aligned with the novel entities boundary.
    »
    Regional: legally bind cities to hinterland conservation such as transboundary watershed protection.
    »
    Global: enforce extraterritorial impact accountability such as supply chain deforestation.
  • Policy transformation
    Equips policymakers with metrics directly linking urban performance to planetary boundary compliance, such as
    »
    Connecting carbon neutrality to Arctic stability;
    »
    Tying urban resilience to rainforest preservation;
    »
    Enforcing ecological reciprocity mechanisms.

2. Literature Review

The natural environment constitutes the irreplaceable foundation of all industrial–urban systems, providing essential resources and ecosystem services—from freshwater and climate regulation to raw materials—that sustain economic development at every stage. However, climate-driven biodiversity collapse and ecosystem degradation are rapidly destabilizing Earth’s ecological integrity. This erosion directly threatens industrial–urban economies, livelihoods, and socio-economic progress, exposing a fundamental contradiction: the very industrial–urban systems driving growth are undermining their own biophysical foundations.
Consequently, sustainable industrial–urban development has become a critical policy priority, driving extensive research into nature-based solutions (NbSs), circular economy (CE) frameworks, and ecological resilience. Central to this effort is enhancing industrial–urban resilience—the capacity to withstand climate shocks like droughts and heatwaves—increasingly dependent on integrating NbSs into planning. Yet, progress remains uneven. Friant et al. (2023) [22] have critiqued the pervasive reliance of urban CE models in cities like Amsterdam, Glasgow, and Copenhagen on technological fixes such as biofuels, carbon capture and storage, and geoengineering. They argued that such technocentric approaches sideline socio-ecological complexities like equity and biodiversity loss while perpetuating growth-oriented systems, advocating instead for structural urban reforms. These include prioritizing compact, walkable neighborhoods to minimize land and energy use; phasing out car dependency through robust public transit and pedestrian-first infrastructure; establishing strict protection for green corridors while replacing gray with green infrastructure; and creating conservation areas banning extractive activities. This underscores the need to replace superficial tech solutions with shifts aligning development to ecological boundaries.
Complementary approaches include that of Ragazou et al. (2024) [23], who integrated CE and urban metabolism (UM) principles into a unified sustainable infrastructure framework. They advocated for cities as closed-loop systems prioritizing waste prevention, material reuse, and regenerative resource recovery through a strategy focused on reducing raw material extraction via efficient design and demand management; reusing products and materials through repurposing and industrial symbiosis; and systematically recovering waste streams such as organic waste for reintegration into production cycles. This alignment of UM’s material–energy flow focus with CE’s cyclicality reimagines urban infrastructure as dynamic, self-sustaining ecosystems, shifting from linear models to enhance resilience, reduce footprints, and align with sustainability. Cialani (2025) [24] further challenged cities to confront unsustainable material footprints, advocating transformative policies prioritizing resource productivity—maximizing societal value per material input—through adopting CE, implementing sustainable practices via industrial symbiosis, and promoting R&D for consumption mitigation.
The World Bank (2021) [25] highlighted the transformative potential of place-based NbSs in mitigating urban environmental risks, evidenced by case studies: Freetown’s urban reforestation combating heat and erosion; Singapore’s river renaturation restoring flood resilience and biodiversity at Bishan-Ang Mo Kio Park; and New York City’s green retrofits, such as rooftop gardens and permeable pavements, addressing stormwater and pollution. These tackle interconnected hazards—heat, flooding, droughts, and pollution—while enhancing livability. The report prioritizes scalable, multifunctional NbSs, like green infrastructure such as urban forests, terraced slopes, and re-naturalized waterways. It also includes built-environment retrofits like green buildings, bioswales, and rooftop farms; and ecological networks such as expansive green spaces, biodiversity corridors, and urban agriculture. By anchoring resilience in local ecosystems, cities can harmonize development with restoration, turning risks into opportunities for equitable growth. Similarly, Chowdhury et al. (2025) [26] analyzed blue infrastructure’s role—natural waterways and engineered systems—in Indian megacities, demonstrating that proactive maintenance enhances biodiversity, flood resilience, and sustainability by preserving aquatic ecosystems, mitigating heat islands, and securing water.
McPhearson et al. (2022) [27] advanced the Social–Ecological–Technological Systems (SETSs) framework for NbS design, stressing synchronization of ecosystems, engineered systems, and socio-economic priorities. Their analysis of urban cooling, like tree planting, showed that maximizing benefits requires multiscale integration: ecologically, enhancing street trees for local cooling; technologically, pairing green spaces with smart irrigation or sensors; and socially, engaging communities and ensuring equitable distribution. The study identified critical hurdles, including managing ecosystem service trade-offs such as biodiversity versus density. It also highlighted the substitutability dilemma, where human-made infrastructure displaces natural systems, eroding resilience. This framework reframes NbSs as dynamic systems demanding holistic governance.
Despite these advances, critical gaps persist in urban sustainability research. Existing intra-urban studies, for example, Friant et al. (2023) [22] and World Bank (2021) [25], focus narrowly on localized interventions like circular economies and green infrastructure but overlook systemic planetary dynamics. Crucially, current frameworks ignore transboundary ecological impacts, where urban resource consumption drives telecoupled degradation—distant crises like ocean acidification, deforestation, and permafrost thaw. Furthermore, while local NbS interventions such as green roofs and urban green spaces are prioritized, regional scales, such as transboundary forest conservation, and global scales, such as high-seas marine protections, remain neglected, undermining holistic resilience.
Few studies address biogeophysical feedback loops, such as how urban GHG emissions accelerate polar ice melt and permafrost collapse, releasing methane and intensifying warming—a self-reinforcing loop eroding industrial–urban stability. An accountability deficit also exists; models, as in the case of Ragazou et al. (2024) [23], emphasizing local sustainability often absolve cities of planetary harm, like carbon outsourcing and biodiversity loss, due to a lack of binding mechanisms to assign ecological debt for embodied emissions in supply chains. It thus follows that industrial–urban sustainability research must expand beyond city limits to address asymmetric responsibility and cross-scale ecological feedback—key to avoiding maladaptive, myopic solutions.

3. Materials and Methods

Industrial–urban systems destabilize planetary boundaries through cross-scale feedback loops, demanding integrated solutions. This study employs an interdisciplinary approach to bridge gaps in M-NbS policies and develop a systemic framework for industrial–urban sustainability. Figure 1 integrates the analytical foundation of this study. Earth system science and ecology quantify biophysical thresholds, for example, cryosphere stability, systems theory models cross-scale feedback such as entropic decay (Section 5.1 and Section 5.2), and empirical case studies validate solution efficacy.
This study employs an interdisciplinary approach to bridge gaps in M-NbS policies and develop a systemic framework for industrial–urban sustainability. Figure 1 integrates the interdependent concepts governing industrial–urban metabolic disruption, unified under the core thesis that industrial–urban systems destabilize planetary boundaries through cross-scale feedback loops, demanding integrated solutions. Each concept is indispensable to this thesis: (a) Earth system science quantifies biophysical thresholds such as Ωarag > 2.75 for marine resilience, providing non-negotiable planetary boundaries, (b) systems theory models feedback mechanisms such as entropy-driven ice-albedo loops, diagnosing causal pathways of industrial–urban disruption, with entropy quantifying the thermodynamic driver of resource degradation, and (c) empirical case studies validate intervention efficacy. These components converge through M-NbRSs, translating diagnostic insights into scalable policies, as empirically demonstrated later in this study.
The present study is quantified through the following datasets.

3.1. Resource Flow

Datasets were compiled to quantify industrial–urban resource flows across spatial scales, enabling analysis of their cascading Earth system impacts through the M-NbRS framework. The following key datasets were collected and analyzed:
  • Material flows and circularity
    • Compiled from the Circular Economy Report 2023 [2] documenting
      »
      Annual global material throughput: 100 billion tons.
      »
      Circularity rate: <9% (2020–2023 average).
  • Energy metabolism
    • Fossil fuel dependence quantified using
      »
      Joint Research Centre (JRC) Data Catalogue (European Commission) [28];
      »
      Statistical Review of World Energy [29];
      »
      Key finding: 87% of primary energy supply derived from coal, oil, and gas (1970–2023 average).
  • Land system impacts including mining, agriculture, and timber extraction footprints from
    »
    Peer-reviewed studies [18,19,20,21,30];
    »
    Institutional datasets including World Resources Institute and Global Forest Watch near-real-time deforestation alerts [5,6].

3.2. Earth System Stressor Quantification

Building on the industrial–urban resource flows quantified in Section 3.1, we employ cross-scale stressor mapping to systematically trace their planetary boundary impacts through key stressor categories using standardized and authoritative datasets:
  • Atmospheric stressors were characterized through annual fossil CO2 emissions (26 Gt C/yr, 1970–2024) sourced from JRC energy balances [28,29].
  • Contemporary CO2 concentrations (424.61 ppm in 2024) from Mauna Loa Observatory records [31] were used.
  • Hydrospheric impacts were assessed via NASA’s satellite-derived ocean heat content measurements (372 Zettajoules) accumulated since 1957) [32] and GLODAPv2 aragonite saturation levels (Ωarag 2.8) [28].
  • Cryospheric changes, documented through ESA and Copernicus satellite missions [33,34], revealed 28 trillion tons of ice loss (1994–2017).
  • Biospheric pressures emerged from Global Forest Watch deforestation alerts (61.56 million ha of tropical forests lost since 2001) [5].
  • Biodiversity risk: IPBES (2019) Global Assessment [35] documents that one million species face extinction risk, with 75% of terrestrial environments severely altered by human activities.
  • Lithospheric degradation was evidenced by mining footprints (5.57 Mha in 2020) [30] and agricultural conversion of Southeast Asian rainforests (8.2 Mha) [36]. These cross-domain stressors are synthesized in Table 1 below.

3.3. Translating Resource Flows into Systemic Threshold

To quantify how industrial–urban resource flows (Section 3.1) transgress planetary boundaries, we integrate
  • Industrial ecology metrics such as material throughout or fossil fuel dependence;
  • Planetary boundary frameworks such as atmospheric CO2 concentration and biosphere integrity;
  • Cross-domain stressor mapping (Table 1).
Resource flows trajectories from extraction to Earth system disruption are analyzed through two complementary approaches:
  • Causal pathways, for example, fossil fuel combustion → CO2 emissions →cryosphere destabilization;
  • Quantitative thresholds, for example, 26 GtC/yr emissions → 424.61 ppm → 28 trillion tons of ice loss. This translation is formalized in Table 2 in the Results Section below.

3.4. Results

Building on the methodology for translating resource flow into thresholds (Section 3.3), our analysis reveals three key findings from quantified industrial–urban demands: (1) the metabolic footprint of resource extraction, (2) cross-scale feedback mechanisms, and (3) threshold-exceeding impacts on planetary boundaries. Table 2 operationalizes this translation, demonstrating how resource flows breach biophysical thresholds—from local material inefficiency to global cryosphere destabilization—triggering cascading Earth system feedback as formalized below:
The cross-scale impacts documented in Table 2 demonstrate how industrial–urban systems create local economic value while generating global ecological disorder through three primary pathways. First, material inefficiency manifests through the extraction of 100 billion tons of resources annually at less than 9% circularity. Second, biosphere degradation occurs via the loss of 61.56 million hectares of forests. Third, cryosphere destabilization emerges from 28 trillion tons of ice mass loss. These interconnected crises demand solutions that address both their systemic causes and measurable impacts.
Table 3 responds to this need by translating Table 2’s diagnostics into actionable policy interventions. For each major stressor, it pairs nature-based solutions with their quantified benefits, creating a direct pathway from problem to resolution. Afforestation programs counter CO2 emissions by sequestering 1.87 billion tons of carbon, while transboundary conservation initiatives protect over 24 million hectares of critical ecosystems. This structured comparison specifically addresses the reviewer’s request for quantifiable metrics comparing industrial efficiency with ecosystem health outcomes.
The table serves as a bridge between Table 2’s problem framework and the detailed implementation case studies in Tables 5 and 6 (Section 8). By maintaining focus on measurable outcomes—from carbon storage to species protection—it provides policymakers with clear criteria for assessing intervention effectiveness while respecting planetary boundaries.
By distilling these linkages, Table 3 translates the entropy crisis into targeted interventions, allowing for quantifiable comparisons between efficiency and ecosystem health.

4. Conceptual Underpinning: Systems Thinking and Dynamics

The Earth system functions as a dynamic, interconnected network of subsystems, where human activities exert profound yet often imperceptible influences. These systems are defined by complex interactions among components—such as the atmosphere, biosphere, and cryosphere—which collectively generate emergent properties: behaviors or outcomes that arise only when components interact as a unified whole. Emergent properties are distinct properties that do not belong to the individual components themselves [37,38,39].
These emergent properties frequently cascade into nonlinear feedback loops, as demonstrated by the Arctic permafrost–climate interaction: (1) Thawing permafrost releases methane (CH4), a greenhouse gas 28 times more potent than CO2 over 100 years; (2) this amplifies atmospheric warming, accelerating further thaw and ice melt; while (3) the consequent albedo reduction (from 0.9 to 0.1 as reflective snow/ice is replaced by absorptive surfaces) creates a self-reinforcing cycle that drives Arctic amplification. Such tightly coupled dynamics challenge predictability, as local perturbations, for example, 1 °C Arctic warming, generate disproportionate global impacts through teleconnections [40]. Systems theory provides critical tools to map these cross-scale interactions, revealing how industrial–urban metabolism disrupts Earth’s stabilizing feedback [37,38,39]. Earth subsystems exhibit emergent properties, for example, from permafrost thaw to methane release followed by warming and albedo loss, that cascade nonlinearly.
These cross-scale dynamics underscore how human activities disrupt Earth’s stability mechanism, formally defined through systems theory (Figure 2). Earth sub-systems manipulate signals to transform inputs into outputs. An input signal x(t), representing an external stimulus applied to the system, manifests as
  • Disturbances: unintended forces such as GHG emissions driving radiative forcing;
  • Control signals: policy interventions such as carbon pricing policies.
The output signal y(t) constitutes the system’s observable response generated through internal dynamics. Due to system complexities, outputs often lag or deviate from intended outcomes via
  • Delays/nonlinearities as demonstrated by GHG emissions → radiative forcing (immediate) → ice melt (delayed);
  • Feedback dependencies as exemplified by carbon tax → emissions reduction (short-term) → gradual warming (long-term).
Internally, stocks, for example, atmospheric CO2 or glacier ice, accumulate state variables altered by flows, that is, inflows such as CO2 emissions and outflows such as oceanic absorption. Crucially, feedback mechanisms dynamically reshape behavior:
  • Positive feedback amplifies change, for example, ice melt → albedo reduction → warming further melt;
  • Negative feedback stabilizes; for example, engineered negative feedback can stabilize ocean systems—for instance, artificial enhancement raises pH to improve CO2 uptake and counteract acidification.
  • Conversely, negative (balancing) feedback acts to stabilize the system, counteracting deviations and promoting equilibrium, as seen when phytoplankton blooms absorb CO2, temporarily offsetting emissions and dampening atmospheric concentration increase [41]. These elements—inputs, outputs, stocks, flows, and feedback loops—collectively define a system’s structure and dynamics.
Figure 2. System dynamics: stocks, flows, and feedback loops. Notes: Conceptual model of system dynamics showing stocks (accumulated quantities) and flows (inflows/outflows). It illustrates how input signals, for example, GHG emissions, generate outputs through system processing, with potential feedback loops that may amplify (positive) or stabilize (negative) responses.
Figure 2. System dynamics: stocks, flows, and feedback loops. Notes: Conceptual model of system dynamics showing stocks (accumulated quantities) and flows (inflows/outflows). It illustrates how input signals, for example, GHG emissions, generate outputs through system processing, with potential feedback loops that may amplify (positive) or stabilize (negative) responses.
Urbansci 09 00320 g002
By mapping these dynamics, systems thinkers identify leverage points—strategic interventions that maximize desired outcomes while minimizing unintended consequences. Figure 3’s feedback quantifies this critical balance: urban heat mitigation leverages negative feedback (blues loop) to reduce energy demand, whereas unchecked emissions risk triggering dangerous positive feedback (red loop) with runaway climate impacts. This paradigm reveals sustainability as a dynamic network of reciprocity, not static entities. As formalized in Figure 3’s system notation, climate change epitomizes planetary threshold complexity:
  • Reinforcing feedback (red arrows): Increased atmospheric CO2 [x1(t) > 0] reinforces warming impact due to increased atmospheric CO2 concentrations, leading to ice melt and albedo reduction. This amplifies solar absorption, aggravating further warming.
  • Stabilizing feedback (blue arrows): Ocean CO2 uptake [x2(t) < 0] converts atmospheric CO2 to dissolved organic carbon, reducing the warming effect, slowing ice melt, and, hence, moderating equilibrium shift. While the ocean carbon sink provides net stabilizing feedback for the atmosphere, it constitutes a compound feedback system with embedded positive loops due to thermodynamic tradeoffs. Sustained CO2 absorption beyond saturation thresholds induces ocean heating and thermal expansion and reduces future uptake capacity.
The figure’s weighted signals (+, −) prove decisive: when [x1(t)] > [x2(t)], incremental shifts cascade towards tipping points such as irreversible ice loss. Conversely, promoting negative loop mechanism—such ocean alkalinity enhancement and promoting calcifying phytoplankton growth, particularly calcifying phytoplankton communities —can stabilize climate by actively sequestrating carbon and restoring biogeochemical equilibrium [41]. Ultimately, localized emissions, through Figure 3’s nonlinear couplings, reveal how Earth systems transform regional inputs into global change: urban GHG releases trigger polar amplification via ice-albedo feedback.
Figure 3. Climate feedback systems: reinforcing and balancing loops. Notes: Climate-relevant examples of positive (reinforcing) and negative (balancing) feedback loops. Positive feedback, for example, ice-albedo loss, accelerates warming, while negative feedback, for example, oceanic CO2 absorption, dampens changes. Arrow weights denote interaction strengths.
Figure 3. Climate feedback systems: reinforcing and balancing loops. Notes: Climate-relevant examples of positive (reinforcing) and negative (balancing) feedback loops. Positive feedback, for example, ice-albedo loss, accelerates warming, while negative feedback, for example, oceanic CO2 absorption, dampens changes. Arrow weights denote interaction strengths.
Urbansci 09 00320 g003

Quantifying Feedback Loops

Feedback loops are systematically analyzed using system dynamics modeling (Section 4) and empirically validated through cross-scale stressor mapping (Section 7):
  • Mechanism
    »
    Figure 2 formalizes feedback loops using system notation, for example, ice-albedo loss as a reinforcing loop: CO2 warming → ice melt → albedo reduction → further warming.
    »
    Table 2 quantifies feedback severity, for example, permafrost releases CH4, amplifying warming by 28 times CO2-eq.
  • Empirical validation
    »
    Cryosphere: 28 trillion tons of ice loss (1994–2024) accelerate albedo reduction, increasing warming three times faster than global averages (Section 7.4).
    »
    Biosphere: tropical deforestation (61.56 Mha since 2001) reduces carbon sinks, intensifying atmospheric CO2 (Section 8.2).

5. Industrial–Urban System: A Metabolic Perspective

Industrial–urban systems function as complex, interdependent networks that drive modern economies through three core metabolic processes: natural resource extraction such as mining and forestry, material transformation such as refining and manufacturing, and urban development like infrastructure and housing. These systems comprise two dynamically linked subsystems:
  • The industrial subsystem focuses on capital- and energy-intensive activities. Its scope spans primary industries such as raw material extraction, including mining and agriculture, and secondary industries covering material transformation through manufacturing, such as steel and chemical production. Key traits include high material/energy throughput and the generation of foundational inputs for economic production such as machinery and goods and services.
  • The urban subsystem centers on consumption patterns and service provision. Its scope encompasses tertiary industries including healthcare, education, and finance and quaternary industries covering knowledge-based sectors like IT and R&D, while also providing social amenities like housing and public spaces. Key traits include networked infrastructure such as transport and utilities and support for light industries, for example, textiles and electronics.
Critically, these subsystems operate in mutual dependence: the industrial subsystem fuels urban growth by supplying materials and energy, while the urban subsystem sustains industrial activity through labor, innovation, and consumption. This metabolic perspective reveals cities as “organisms” that metabolize resources—transforming natural inputs into economic outputs and waste—within a planetary-scale material loop.

5.1. Entropy, Thermodynamics, and Industrial–Urban Metabolism

Industrial–urban systems operate as metabolic networks governed by fundamental thermodynamic laws. Entropy—a measure of disorder and energy dispersal—provides a critical lens for understanding their dynamics [15,16,17,42]. The second law of thermodynamics dictates that all processes irreversibly increase total entropy, driving industrial metabolism to convert ordered, low-entropy resources like fossil fuels, minerals, and biomass into high-entropy waste such as CO2, pollutants, and waste heat [15,16]. This degradation is inherent: every extraction, manufacturing, or energy-use step disperses energy and reduces material quality, aligning with the universe’s trajectory toward disorder. As shown in Figure 3, even advanced recycling cannot fully restore resources to their original low-entropy state, making waste generation an inescapable consequence of economic activity.
The entropic decay of industrial–urban metabolism manifests in profound systemic crises:
  • Urban heat islands, amplified by waste heat, elevate local temperatures by 2–5 °C, escalating energy demand and emissions [43].
  • The accelerated production and release of novel entities—including plastics, pesticides, industrial compounds, and antibiotics—now exceeds the planetary boundary for chemical pollution. This breach of humanity’s safe operating space critically erodes biosphere integrity by disrupting biogeochemical cycles and ecosystem resilience [44].
  • Heavy metals—persistent, high-entropy pollutants—disperse systemic disorder through Earth’s hydrosphere, biosphere, and lithosphere. Their bioaccumulation threatens ecosystem resilience and human health across all trophic levels, exemplifying transboundary novel entity risks that destabilize planetary boundaries [45,46,47].
This generates a core paradox as shown in Figure 4: industrial–urban systems rely on finite low-entropy stocks such as fossil fuels, minerals, and ecosystems for growth, yet their metabolic processes intrinsically degrade these foundations by producing destabilizing high-entropy waste. Thermodynamic constraints compound this tension—Earth’s capacity to dissipate waste entropy via oceanic, atmospheric, and terrestrial sinks is fundamentally bounded, imposing absolute limits on industrial throughput. The Second Law therefore governs not merely local resource efficiency but planetary-scale stability, exposing an irreconcilable conflict between indefinite economic expansion and Earth’s biophysical limits.

5.2. Industrial–Urban Systems: Manifestation of (Low) Entropy Paradox

Industrial–urban systems epitomize the entropy paradox: the irreversible degradation of finite low-entropy resources into high-entropy waste as formalized in Figure 4. This manifests in three key pathways:
  • Resource flow (Figure 5, left arrow): extraction of low entropy stocks such as fossil fuels and minerals.
  • Urban metabolism (Figure 5, center box): Linear take–make–use–dispose processing;
  • Waste outflow (Figure 5, right arrow): Emissions of high-entropy outputs destabilizing Earth systems.
Crucially, Figure 5 quantifies the paradox: valuable low-entropy resources such as petroleum for plastics become permanent ecological liabilities, for example, microplastics in oceans. A smart phone exemplifies this irreversible trajectory shown in Figure 5’s cascading outputs. Its production consumes rare minerals (low-entropy inputs), yet its temporary order is offset by accelerated global entropy through life cycle emissions. Thus, industrial outputs represent entropy-debt intermediaries in a cycle of borrowed order, extracting planetary stocks and causing ecological degradation [48].
Urban systems intensify the low-entropy paradox—defined as the irreversible degradation of the metabolic flows of low-entropy energy and materials into high entropy—acting as metabolic nexuses. This occurs through linear “take–make–use–dispose” flows that convert imported resources (energy, materials, water) into three destabilizing waste streams: (a) municipal solid waste (plastics, textiles); (b) nutrient/chemical-laden wastewater; and (c) airborne pollutants (CO2, NOx, PM2.5).
As shown in Figure 4, cities starkly illustrate this imbalance—importing food while exporting packaging waste—straining ecosystems and exposing urban fragility within planetary boundaries [48]. This is reflected in Figure 2 where the thermodynamic cascades manifest as
  • Atmospheric destabilization: GHG emissions trigger feedback loops such as permafrost thaw that accelerate warming (see Table 2 above).
  • Biogeochemical degradation: nutrient runoff creates marine dead zones while synthetic toxins such as per- and polyfluoroalkyl substances (PFASs) and heavy metals poison soil/water [49,50].
  • Entropic legacies: non-recyclable waste such as plastics persists for millennia, irreversibly altering ecosystems [51].
This systemic mismatch between industrial metabolism and Earth’s regenerative capacity demands urgent realignment with biospheric limits, as explored in the subsequent sections.

6. Earth System Science and Ecology

Earth system science provides a critical framework for understanding how industrial–urban metabolism disrupts planetary stability, quantifying cascading failures across global biogeochemical cycles, climate regulators, and ecosystem integrity. It reveals cities as amplifiers of interconnected crises—accelerating climate warming, ocean acidification, and biodiversity collapse through resource extraction, waste generation, and habitat destruction. This systems lens bridges diagnosis and action by modeling feedback loops between urban activities and Earth’s regulatory capacity, informing strategies for climate resilience, disaster risk reduction, and equitable mitigation. The framework centers on five dynamically coupled spheres: the atmosphere (climate-regulating gases), hydrosphere (interconnected water systems), lithosphere (land and geological substrates), cryosphere (ice reservoirs), and biosphere (ecological networks). Their destabilization underscores a non-negotiable imperative: realigning human systems with planetary boundaries is essential to avoid irreversible tipping points [52,53]. Cities must transition from disruptors to stewards by embedding these spheres’ limits into urban design, governance, and metabolism.

6.1. Earth System Science in Perspective

6.1.1. Atmosphere: Earth’s Climate Regulator

The atmosphere serves as Earth’s primary climate-regulating interface, maintaining planetary habitability through three core mechanisms. This gaseous envelope extends hundreds of kilometers from Earth’s surface, with 99% of its mass concentrated below 30 km, and consists primarily of nitrogen (78%), oxygen (21%), argon (0.9%), and trace greenhouse gases (CO2, CH4, N2O, H2O vapor).
These atmospheric components interact dynamically to sustain life through three fundamental regulatory processes. First, thermal regulation occurs as greenhouse gases trap infrared radiation, maintaining Earth’s habitable mean temperature of approximately 15 °C through the greenhouse effect [54]. Second, stratospheric ozone provides critical radiation protection by absorbing 97–99% of solar UV-B radiation, shielding terrestrial life from mutagenic damage [55]. Third, dynamic carbon cycling facilitates atmosphere–biosphere CO2 exchanges of about 120 GtC annually, balancing fluxes between photosynthetic carbon sinks and respiratory/decomposition sources [56,57].

6.1.2. Hydrosphere: Oceanic Climate Stabilization

Encompassing 71% of Earth’s surface and containing 1.38 billion km3 of water, the hydrosphere regulates climate through unparalleled carbon sequestration and heat absorption. Oceans store 38,000–40,000 GtC as dissolved inorganic carbon (CO2, bicarbonate, carbonate) and organic carbon (700 GtC), a capacity 16 times greater than terrestrial biosphere stocks [58,59,60,61].
Annually, 150 GtC exchanges with the atmosphere, while anthropogenic emissions drive a net oceanic uptake of 2.7–3 GtC/year (30% of emissions). This occurs through dual mechanisms: the physical pump transports CO2 to deep waters, while the biological pump transfers 5–10 GtC/year via sinking biomass [62,63,64]. Crucially, oceans absorb over 90% of excess greenhouse gas heat (372 zettajoules by 2021), leveraging a heat capacity four times greater than air to buffer global warming and extreme weather [32,65].

6.1.3. Cryosphere: Earth’s Fragile Thermostat

Covering 10% of Earth’s land and 7% of its oceans, the cryosphere stabilizes global climate through three critical processes. Permafrost stores 1400–1700 GtC—twice atmospheric carbon and equivalent to 51 times 2019 global fossil fuel emissions—but risks mobilization as warming accelerates thaw [66,67,68].
Snow and ice reflect 80–90% of solar radiation (albedo effect), contrasting the 5–10% reflectivity of dark ocean waters. Ice-albedo feedback exerts net cooling, but shrinkage replaces reflective surfaces with heat-absorbing surfaces, triggering self-reinforcing warming [69,70,71]. Additionally, sea ice regulates ocean heat absorption patterns by insulating polar oceans. Its decline disrupts this buffer, amplifying regional warming and permafrost thaw [72,73,74].
Thawing permafrost also poses ecological and public health risks by releasing ancient microbes, including dormant pathogens, which could disrupt modern microbial communities and threaten human health through novel exposure pathways [75]. Concurrently, permafrost-derived organic matter fuels microbial activity in Arctic ecosystems, altering biogeochemical cycles and further destabilizing thaw-affected regions [76].

6.1.4. Lithosphere: Planetary Carbon Archive

The lithosphere, Earth’s rigid outer shell (5–100 km thick), serves as the planet’s largest carbon reservoir, storing approximately 65,500 GtC across carbonate rocks (60,000 GtC), fossil fuel deposits (4000 GtC), and soils (1500 GtC) [77,78]. Its climate regulation involves opposing feedback mechanisms.
Stabilizing processes include silicate weathering, which acts as a natural CO2 thermostat during warming periods, and soil sequestration of 1–3 GtC annually through organic matter accumulation [79]. Carbonate sedimentation further contributes to long-term carbon burial in marine sediments, locking carbon in limestone for millions of years [80].
Conversely, destabilizing feedback dominates human impacts. Permafrost thaw releases ancient soil carbon [81,82], while fossil fuel combustion emits 26 GtC annually (1970–2023 average) [28,83]. Three corporations—Saudi Aramco (16.6 GtC), Chevron (14.2 GtC), and ExxonMobil (13.2 GtC)—cumulatively emitted 44 GtC from 1920 to 2020, directly contributing 0.061 °C of observed warming [84]. Land-use changes add about 4 GtCO2e/year, with 2024 net emissions reaching 4.2 GtCO2 [56,85]. This imbalance overwhelms the lithosphere’s natural stabilizing capacities, accelerating climate disruption.

6.1.5. Biosphere: Climate-Regulating Web of Life

The biosphere maintains carbon equilibrium through interconnected biological processes. Peatlands store approximately 600 GtC globally—more than any other ecosystem—while sequestering 0.37 GtC annually. Tropical regions hold 20% of this storage, exemplified by the Congo Basin’s Cuvette Centrale peatlands (30 billion tons across 145,500 km2) [86,87,88].
Forests contain 662 billion tons of carbon, with net sequestration of 7.6 GtCO2 annually due to reforestation and improved management [89,90,91,92]. The Amazon alone stores 150–200 billion tons (367–733 GtCO2 equivalent) while contributing 20% of Atlantic freshwater discharge, critically influencing global water cycles [93,94].
Biodiversity underpins this balance. Photosynthesis converts atmospheric CO2 into organic matter (2.6 GtCO2 absorbed annually globally), while microbial decomposition mediates carbon turnover. Fungi enhance soil carbon storage by 20% compared to bacteria, forming stable humus compounds [95,96,97]. Collectively, these systems buffer anthropogenic emissions, but land-use change and warming risk destabilizing this delicate regulatory network.

7. The Anthroposphere: Vicious Cycle Atlas of Fragility—Cascading Feedback in Industrial–Urban Systems

The anthroposphere comprises a complex network of interconnected cities, industrial zones, and agricultural systems embedded within the biosphere. Despite occupying only 3% of Earth’s land surface, it supports 56% of the global population and serves as the primary hub for manufacturing, processing, transportation, and urban consumption [98,99]. The anthroposphere functions as the epicenter of planetary fragility, as depicted in our Atlas of Fragility (Figure 6), which formalized cross-scale stressor mapping to systematize its destabilizing feedback across interconnected spheres.
More specifically, these activities trigger self-reinforcing degradation cycles (red arrows) across interconnected spheres while simultaneously offering stabilizing solutions (negative feedback). The framework spans five key spheres (the biosphere, hydrosphere, lithosphere, cryosphere, and anthroposphere), with particular emphasis on (a) climate change drivers—CO2 emissions, deforestation, and industrial pollution, and (b) mitigation pathways—nature-based solutions and systemic interventions.
To emphasize, Figure 5 presents the Atlas of Fragility, a systematic mapping of the anthroposphere’s destabilizing influence on Earth’s spheres through two competing feedback regimes, as formalized in Figure 3: (1) amplifying feedback (red arrows), where self-reinforcing degradation cycles—such as urban heat islands intensifying energy demand—exacerbate environmental collapse; and (2) stabilizing feedback (blue arrows), where restorative processes like mangrove restoration mitigate systemic damages. This atlas reveals critical thresholds of irreversibility, for example, ice-albedo loss and permafrost carbon release across interconnected systems, quantifying how industrial–urban metabolism converts localized disruptions into planetary-scale fragility such as rising global sea level and ocean deoxygenation [100,101]
Below, we examine these dynamics across interconnected systems, highlighting their cascading impacts and thresholds of irreversibility.

7.1. Sphere-Specific Dynamics

  • Biosphere dynamics
    • Amplifying feedback (degradation cycles)
      »
      Deforestation → biodiversity loss → reduced carbon sequestration → accelerated climate warming [35];
      »
      Pollinator collapse → agricultural disruption → food system instability [36].
    • Stabilizing feedback (restorative pathways)
      »
      Agroforestry adoption→ pollinator habitat recovery → ecosystem resilience.
    • Interconnections: thresholds/risks
    Linked to atmospheric CO2 rise (424.6 ppm) via carbon sink loss [31].
  • Hydrosphere dynamics
    • Amplifying feedback (degradation cycles)
      »
      Industrial runoff → ocean acidification (Ωarag = 2.80) → coral skeletal dissolution → marine ecosystem collapse [18] (critical threshold: current Ωarag (2.80) is 0.05 units above collapse (2.75) and 18.6% more acidic than preindustrial levels (3.44), risking irreversible damage [18]);
      »
      Overfishing → fishery stock depletion → food insecurity → increased fishing pressure.
    • Cross-system CO2 cascade
      »
      Atmospheric CO2 → simultaneous warming and acidification → reduced oceanic CO2 absorption → accelerated global warming → coral reef degradation.
  • Lithosphere dynamics
    »
    Mining activities → soil degradation → increased land instability/increased landslide occurrences;
    »
    Urban sprawl → habitat fragmentation → erosion acceleration.
  • Cryosphere
    Amplifying feedback (degradation cycles)
    »
    Ice-albedo feedback loop: global warming → ice melt → darker surfaces (albedo reduction of 0.6→0.1) [69,70] → increased heat absorption → accelerated melting.
  • Atmosphere dynamics: anthroposphere driver
    »
    Fossil fuel combustion → CO2 accumulation (424.6 ppm in 2024) [31] → enhanced radiative forcing.
By implications, the Vicious Cycle Atlas of Fragility ultimately exposes the entropy paradox in action: anthropospheric ‘order’ systematically converts low-entropy stocks into high-entropy waste as quantified in Figure 5’s feedback loops. This irreversible transformation—documented across biosphere collapse (61.56 Mha deforestation), cryosphere destabilization (28 trillion tonnes ice loss), and lithospheric degradation (5.57 Mha mining impacts)—overwhelms Earth’s dissipative capacity, locking industrial–urban systems into a trajectory of cascading fragility.

7.2. Industrial–Urban Metabolic Crisis

Industrial–urban systems operate through a linear metabolic process that extracts low-entropy resources and transforms them into goods and services, while generating high-entropy waste. This model remains fundamentally dependent on nature’s life-support systems: 55% of global GDP (USD 58 trillion) and 85% of S&P 1200 corporations exhibit moderate-to-high reliance on ecological services—including freshwater provision, climate regulation, and nutrient cycling [102,103]. Yet, these same systems drive unprecedented biophysical degradation. Between 1970 and 2017, about 2.5 trillion tons of materials were extracted from Earth’s crust and biosphere—fueling economic growth while overwhelming planetary sinks and accelerating entropy-driven feedback loops [104].
As shown in Figure 7, global industrial–urban systems exhibit unsustainable metabolic scaling, as evidenced by resource extraction surging from 27 billion tons in 1970 to over 100 billion tons annually today—a near quadrupling in five decades. This demand is dominated by non-renewable sources: non-metallic minerals (42.8 billion tons, primarily construction sand/gravel), fossil fuels (15.5 billion tons), metallic ores (9.4 billion tons), and biomass (24.9 billion tons, including timber/crops). Critically, only 8.6 billion tons (8.6%) derive from recycled inputs, revealing severe circularity deficits (Figure 7). Should current trajectories persist, annual extraction could reach 190 billion tons by 2060—exceeding Earth’s regenerative capacity by 75% and risking irreversible biospheric disruption [105,106].
Industrial–urban metabolism binds economic activity to ecological degradation through three self-reinforcing footprints that amplify planetary stress:
  • Carbon footprint: Construction alone generates 39% of global CO2 emissions (11% from material production), while urban residents exhibit stark disparities—4.8 tCO2/yr per capita (global average) versus 16.2 tCO2/yr in high-consumption economies like the United States [107,108].
  • Material footprint: Construction consumes 40–50% of global raw materials driving agricultural and mining expansion that converted 420 million hectares (Mha) of forests into cropland/mining sites (1990–2020) [107,108,109,110,111]. This land transformation is accelerating, with mining output surging 56% (2000–2020) to 17.2 B tons/year [112] and cropland expanding 9% (2003–2019)—most rapidly in Africa and South America [113].
  • Ecological overshoot: Collectively, these pressures exceed Earth’s regenerative capacity by 75%—equivalent to consuming resources of 1.75 Earths [114].
This metabolic triad locks industrial–urban systems into an unsustainable trajectory. Without systemic transformation, a projected material demand of 190 billion tons/year by 2060 will irreversibly sever the connection between economic development and planetary health [2].

7.3. Entrenched Fossil Dominance and Climate Feedback Cycles

Fossil fuels maintain an iron grip on global energy systems, constituting 87% of primary energy consumption between 1970 and 2023 (Figure 8). In 2023 alone, oil consumption reached 196 exajoules (a 2.48% increase from 2022), natural gas rose to 164 exajoules (up 1.54%), and coal remained a persistently high-pollution energy source. This demand is driven by heavy industries—such as steel, cement, and chemicals—as well as hyper-resource-intensive food production, creating a self-reinforcing feedback loop: rising emissions intensify global warming, which in turn increases cooling demand, deepens reliance on fossil fuels, and accelerates planetary destabilization.
Rapid urbanization further entrenches this cycle through three carbon-locking mechanisms:
  • Transport networks (road, rail, and air) expand fossil dependence alongside urban sprawl.
  • Hard infrastructure (dams, power grids) relies on CO2-intensive steel and concrete, contributing to 39% of global carbon footprint [107,109,110].
  • Hard infrastructure (dams, grids) relies on CO2-intensive steel/concrete, contributing to 39% global carbon footprints.
  • Urban energy systems still meet 87% of electricity demand through fossil fuels.
This linear “extract–produce–discard” economic model accelerates resource depletion, underscoring the urgent need for circular alternatives (reuse, repair, recycling).
The climate crisis has already breached irreversible thresholds, as starkly illustrated in Figure 9.
With reference to Figure 8, energy-related CO2 emissions surged by 120.83% between 1970 and 2024, soaring from 14 to 37 billion tons annually [29]. Total GHG emissions (CO2, CH4, N2O, and fluorinated gases) rose even more dramatically—by 161.4%, from 24 to 53 billion tons over the same period [29]. Cumulatively, human activity has released 2.4 trillion tons of CO2 since 1850, with 950 billion tons remaining in the atmosphere today [115]. Atmospheric CO2 concentrations have climbed 30.38% since 1970, from 324.68 ppm to 424.61 ppm in 2024 [31]—far exceeding the 350 ppm safety limit [116].
Figure 9. Fossil fuel CO2 emissions and atmospheric CO2 concentrations. Source: [28,31,117,118]. Notes: The figure synthesizes data from authoritative sources: 2024 CO2 emissions estimates come from the Global Carbon Budget [117], while 2024 GHG emissions figures are sourced from Tiseo [118]. Historical emissions data (both GHG and energy-related CO2) derive from IEA/EDGAR [28], and atmospheric CO2 measurements reflect NOAA/ESRL’s global monitoring records (Mauna Loa trend) [31]. This compilation underscores the inextricable link between fossil fuel use, industrial activity, and atmospheric disruption—a critical baseline for assessing climate mitigation efforts. The data paints an unambiguous picture: despite decades of climate policy discourse, emissions growth remains unchecked, demanding urgent systemic intervention.
Figure 9. Fossil fuel CO2 emissions and atmospheric CO2 concentrations. Source: [28,31,117,118]. Notes: The figure synthesizes data from authoritative sources: 2024 CO2 emissions estimates come from the Global Carbon Budget [117], while 2024 GHG emissions figures are sourced from Tiseo [118]. Historical emissions data (both GHG and energy-related CO2) derive from IEA/EDGAR [28], and atmospheric CO2 measurements reflect NOAA/ESRL’s global monitoring records (Mauna Loa trend) [31]. This compilation underscores the inextricable link between fossil fuel use, industrial activity, and atmospheric disruption—a critical baseline for assessing climate mitigation efforts. The data paints an unambiguous picture: despite decades of climate policy discourse, emissions growth remains unchecked, demanding urgent systemic intervention.
Urbansci 09 00320 g009
These interdependencies fuel a vicious cycle: fossil combustion drives warming, which spikes energy demand, further amplifying emissions and hastening systemic collapse. The data is unequivocal—the crisis is not looming; it is unfolding in real time.
This fossil-driven economic model simultaneously exacerbates another crisis: the relentless expansion of resource extraction, which degrades the biosphere and lithosphere, as explored in the following section.

7.4. Resource Extraction Frontiers: Mining and Agriculture as Drivers of Biospheric Collapse

  • Mining’s multi-sphere impact
Industrial mining directly disrupts the lithosphere through surface excavation, eliminating 1.4 million ha of global tree cover (2001–2020) and releasing 36 million tons of CO2 annually. Tropical forests—which host just 7% of mining operations—bear disproportionate damage due to their high biodiversity and carbon density [119]. The cascading effects are severe:
»
Lithospheric degradation: soil erosion and compaction affect 5.57 million ha globally (2020), with impacts extending 70 km beyond mine site.
»
Biospheric fragmentation: mining disrupts one-third of global forests, exposing 70% to edge effects like invasive species and microclimate shifts [120,121,122,123].
The environmental impacts of mining are acutely concentrated in biodiverse tropical regions, where extraction activities disproportionately threaten fragile ecosystems. Indonesia stands as the epicenter of this crisis, having lost 190,100 ha of forest to mining—accounting for 58.2% of global mining-related deforestation. This devastation has pushed vulnerable species, including the critically endangered Hawksbill turtle, toward extinction [30,124]. In the Brazilian Amazon, 17 million ha have been cleared for mining operations, representing 9% of the region’s total deforestation and fragmenting vital ecosystems that sustain global climate regulation [125]. Kalimantan, Indonesia’s biodiversity-rich province, exemplifies the dual threat to natural and managed forests: 143,592 ha of natural forests and 8765 ha of plantations have been converted to mines, directly endangering IUCN-protected species such as Sun Bears and Rhinoceros Hornbills [124].
These regional case studies underscore how mining operations target Earth’s most ecologically significant landscapes, where forest loss triggers cascading collapses in biodiversity, carbon storage, and indigenous livelihoods. The concentration of damage in tropical hotspots—which represent just 7% of global mining operations but harbor the planet’s highest carbon densities and species richness—reveals a critical mismatch between economic activity and ecological preservation.
  • Agricultural expansion: the primary driver of biospheric degradation
Concurrently, agricultural expansion has surpassed mining as the primary driver of biospheric degradation. Global cropland grew by 9% (2003–2019), converting vast lithospheric surfaces—most rapidly in Africa and South America. This transformation is overwhelmingly export-oriented: 80% of land-use impacts link to agri-food exports from Latin America, Africa, and the Southeast Asia–Pacific [36]. Southeast Asia alone lost 8.2 million ha of lowland rainforests (2000–2020), including 23% within legally protected zones, primarily to oil palm and rubber plantations [36,126]. Williams et al. (2021) [127] estimate that under business-as-usual scenarios, an additional 3.35 million square kilometers of natural ecosystems (representing a 26% increase over 2010 cropland area) will be converted to agricultural use by 2050 to meet demands for food production, animal feed, fiber, and bioenergy crops.
The damage cascades across spheres. The following table (Table 4) systematically quantifies the cascading planetary impacts of anthropogenic land-use change, highlighting the interconnected crises of biosphere collapse, climate feedback, and ecosystem instability through key metrics, regional distributions, and cited evidence.
Without urgent intervention to reform land-use practices and global supply chains, these interconnected disruptions will push Earth systems toward irreversible tipping points. The time for systemic action is now—delay risks crossing thresholds from which recovery may be impossible.
  • Industrial emissions and Earth system destabilization
Covering 71% of Earth’s surface, the hydrosphere functions as the planet’s primary climate stabilizer—absorbing 93% of excess anthropogenic heat since pre-industrial times [130]. This extraordinary capacity stems from water’s specific heat, which is four times greater than air, enabling oceans to store 1000 times more thermal energy than the atmosphere for equivalent temperature changes [65].
By 2021, the global ocean had accumulated 372 zettajoules (ZJ) of excess heat—a staggering 3100% increase since 1957 (12 ZJ). To put this into perspective, this energy release is equivalent to detonating 6.2 billion Hiroshima-scale atomic bombs [32]. Concurrently, global land temperatures rose by 296%, climbing from −1.15 °C in 1850 to +2.25 °C in 2024. Even more dramatically, ocean surface temperatures experienced a 1122% increase, warming from −0.09 °C to +0.92° over the same period. When combined, land and ocean warming reached +1.33 °C by 2024, marking a 296% rise since pre-industrial levels [131]. These temperature and heat anomalies exhibit a strong correlation with rising atmospheric CO2 emissions, which totaled 1.3 trillion tons between 1970 and 2024. When accounting for all greenhouse gases (GHGs), cumulative emissions soared to 2.05 trillion tons [29].
Atmospheric CO2 concentrations reached a record 424.61 parts per million (ppm) in 2024—surpassing any level observed in at least 800,000 years [31]. This unprecedented rise in greenhouse gases has driven radiative forcing to levels that are fundamentally restructuring Earth’s climate system, triggering irreversible feedback loops such as cryospheric collapse. To emphasize, between 1994 and 2024, Earth’s ice sheets and glaciers experienced catastrophic losses totaling 28 trillion tons—equivalent to an annual discharge of 1.2 trillion tons into the oceans [34]. This melt has been distributed across critical reservoirs: mountain glaciers shed 6.54 trillion tons since 2000, with annual losses accelerating by 35% in the past decade [132]. The Greenland Ice Sheet contributed 270 billion tons per year, while Antarctica lost 150 billion tons annually, primarily due to the destabilization of West Antarctic glaciers. Collectively, these losses have driven global sea levels upward by 0.4 mm/yr from Antarctica alone [133].
The retreat of ice and snow has triggered two self-reinforcing feedback loops. First, albedo reduction—the loss of reflective surfaces—now equates to 50,000 km2 of vanished ice cover each year. This has amplified Arctic heating to three times the global average rate [134]. Second, permafrost thaw is mobilizing a carbon reservoir of 1460–1600 gigatons, double the carbon currently in the atmosphere. As frozen organic matter decomposes, it emits CO2, methane (CH4), and nitrous oxide (N2O), with thermokarst lakes acting as hotspots for methane ebullition [135]. Recent studies have confirmed that northern permafrost regions transitioned from a carbon sink to a net GHG source between 2000 and 2020 [136].
These changes are propelling Earth’s climate system toward irreversible thresholds. The Atlantic Meridional Overturning Circulation (AMOC), a critical regulator of global heat distribution, has weakened by 15% since 1950. Paleoclimate data and early-warning signals suggest it may be approaching a nonlinear collapse—an event that would disrupt monsoon systems, destabilize marine ecosystems, and trigger 20–30 cm of regional sea-level rise [137,138]. Meanwhile, atmospheric water vapor—which increases 7% per 1 °C of warming—now accounts for half of all greenhouse forcing, intensifying extreme precipitation events [139,140].
Stabilizing the climate system demands unprecedented intervention. Current emissions trajectories commit Earth to centuries of warming due to CO2’s millennial-scale persistence and feedback loops. Avoiding the collapse of oceanic buffers—which absorb 90% of excess heat and 30% of CO2—requires annual emissions cuts exceeding 7%, paired with gigaton-scale carbon removal. Incremental measures risk triggering civilizational-scale thermodynamic disruptions, underscoring the need for metabolic restructuring of industrial systems.
  • Climate feedback dynamics and hydrological cascades
    Planetary warming is now being driven primarily by self-reinforcing climate feedback mechanisms that have surpassed direct anthropogenic forcing in their impact. Two critical feedback loops are accelerating global temperature rise and hydrological disruption.
    (a)
    The Arctic is warming at triple the global average rate due to dramatic reductions in surface reflectivity. As ice and snow cover melt, albedo decreases from 0.9 to just 0.1—a 90% reduction in solar radiation reflection. This creates a vicious cycle where increased absorption of solar energy leads to further warming, additional ice loss, and accelerated permafrost thaw, with impacts exceeding previous climate model projections [17,18,141,142].
    (b)
    Water vapor–temperature feedback: Each 1 °C rise enables 7% more atmospheric water vapor (Clausius–Clapeyron scaling). Accounting for 50% of total greenhouse forcing, vapor traps additional heat, intensifying extreme precipitation and altering global hydrology [139,140]. The Earth’s climate system is approaching critical thresholds through interconnected hydrospheric destabilization processes, with three primary manifestations of growing concern:
(i)
AMOC weakening and potential collapse
The Atlantic Meridional Overturning Circulation (AMOC), which redistributes heat on our planet and has a major impact on climate, has weakened by approximately 15% since the mid-20th century [143,144]. This critical circulation system, responsible for redistributing heat across the Atlantic basin, now shows multiple indicators of approaching a tipping point [137]. Paleoclimate evidence suggests that such weakening can precede abrupt collapses, with modern observations revealing
»
Reduced deep water formation in the Labrador Sea;
»
Changing surface salinity patterns;
»
Shifting oceanic heat transport.
A potential AMOC collapse would trigger cascading hydrospheric effects with profound global consequences. The disruption of monsoon systems would dramatically alter rainfall patterns, particularly threatening food security in vulnerable regions like South Asia and the Sahel. Marine ecosystems would face catastrophic collapse as abrupt thermohaline changes devastate phytoplankton communities and commercial fisheries. Regional sea levels in the North Atlantic could rise an additional 20–30 cm by 2100 compounding existing sea-level rise from thermal expansion and ice melt [138]. Paradoxically, while global warming continues, parts of Europe might experience temporary cooling due to disrupted heat transport. These changes would also redistribute weather patterns globally, potentially causing extended droughts in some regions while intensifying storm activity in others [145]. Together, these impacts would create a complex web of climatic disruptions affecting everything from agricultural production to coastal infrastructure. This is revealed in the following sub-sections.
(ii)
Hydrospheric feedback amplification
The initial impacts of AMOC weakening would trigger several self-reinforcing feedback loops that could amplify and accelerate climate disruptions. As oceanic heat transport diminishes, paradoxical expansion of Arctic Sea ice could occur in some regions, which through increased surface albedo would further modify hemispheric energy budgets. Simultaneously, altered evaporation patterns across the Atlantic would change ocean circulation and precipitation distribution globally. Perhaps most critically, the disruption of the ocean’s biological carbon pump and reduced solubility uptake in warmer, more stratified waters would impair the marine carbon sink, leading to accelerated accumulation of atmospheric CO2 [130,146,147]. This interconnected feedback—operating through physical, chemical, and biological pathways—could compound initial perturbations to create nonlinear responses in the climate system, making subsequent recovery or stabilization increasingly difficult. The timescales of these feedback processes vary from immediate (circulation changes) to decadal (carbon cycle impacts), creating a cascade of effects that could persist long after initial AMOC weakening [148].
(iii)
Compound climate extremes and Earth system destabilization
The interplay between atmospheric and biospheric systems is accelerating climate extremes through the following synergistic feedback loops:
»
Heatwave intensification: At 1.5 C of global warming, heatwaves last 2–3 times longer due to atmospheric blocking and soil moisture depletion. Tropical megacities like Jakarta and Lagos will face 120–150 additional lethal heat days/year by 2050 [149].
»
Hydrological disruption: Amazon deforestation reduced dry-season rainfall by 30%, degrading its carbon sink by 30% [150,151,152]. Southeast Asia will see monsoon delays of 5–15 days and 22.8% more extreme wet days by 2050 [153].
»
Cryospheric collapse: Global sea-level rise now reaches 3.6 mm/yr (2.5 times faster than 1900s), with regional spikes of 15–20 cm since 1993 [100,132,133]. This stems from thermal expansion and cryospheric collapse, particularly in Greenland and Antarctica [34,132,133], which exacerbate coastal flooding through sterodynamic feedback [32,133,154].
»
Greenland’s meltwater disrupts ocean currents, amplifying coastal flooding [155,156,157,158]. Critically, meltwater from Greenland disrupts North Atlantic deep convection, amplifying sea-level rise along the U.S. East Coast [155,156], while freshwater influx reduces ocean salinity, accelerating ice-shelf melt [157,158]. These processes represent a mass transfer from ice reservoirs to oceans, with cascading geophysical impacts, including shifts in Earth’s rotational inertia.
These changes reflect irreversible climate restructuring, with CO2 persistence [134,159] and feedback (ice-albedo loss [160,161], permafrost thaw [134]) committing Earth to centuries of warming. The millennial-scale persistence of atmospheric CO2 [134,159], compounded by feedback like ice-albedo loss [134,160] and permafrost carbon release [134], commits the planet to centuries of warming. The interdependence of hydrospheric, cryospheric, and atmospheric systems underscores a non-negotiable reality: mitigation delays risk triggering irreversible tipping points in Earth’s critical regulatory processes.

8. Discussion: Earth System Destabilization and Pathways for Regeneration

Validating the M-NbRS framework introduced in Section 3, Earth functions as an interconnected systemEarth functions as an interconnected system of five dynamically coupled subsystems—the atmosphere, hydrosphere, cryosphere, biosphere, and lithosphere—regulated by biogeochemical cycles that have maintained planetary stability for millennia [18,19,20]. The M-NbRS framework’s novelty lies in its capacity to quantify cross-scale synergies—for example, urban tree canopy expansion (local) reducing energy demand, thereby lowering emissions that drive cryosphere loss (global). This is enabled by
  • Telecoupled governance: Transboundary coordination of hinterland conservation, as demonstrated by the Heart of Borneo’s 24 Mha trilateral protection scheme, creates closed-loop accountability through (i) joint policymaking between Brunei, Indonesia, and Malaysia, (ii) harmonized land-use monitoring across political borders, and (iii) reciprocal enforcement of conservation commitments. Noteworthy is the HoBI where Indonesia’s Kalimantan, the Malaysian states of Sarawak and Sabah, and Brunei Darussalam collaborate under the HoB Strategic Action Plan to establish 2.4 Mha of protected area—later expanded to 2.69 Mha—for cross border conservation and CO2 reduction This transboundary initiative directly addresses metabolic imbalances by preserving governance models that directly address metabolic imbalances by preserving low-entropy stocks such as primary rainforests while mitigating entropic waste such as carbon emissions, demonstrating institutional alignment with biospheric limits.
  • Dynamic feedback tools: The Vicious Cycle Atlas of Fragility (Figure 5) diagnoses leverage points. For example, permafrost insulation creates a positive feedback loop—slowing thaw reduces emissions, which moderates regional warming, further protecting permafrost. More specifically, permafrost insulation does not just stabilize carbon stocks (lithosphere); it also reduces methane emissions (atmosphere), protects Arctic ecosystems (biosphere), and supports Indigenous livelihoods (anthroposphere). Mapping these cascading benefits demonstrates the intervention’s high leverage potential.
However, industrial–urban metabolism now operates as a thermodynamic engine of global disruption, characterized by three destructive processes:
  • Linear resource extraction, exceeding 100 billion tons annually [2];
  • Fossil fuel combustion, driving 75% of global greenhouse gas emissions [28];
  • Rampant habitat destruction [5,6]. This system forcibly converts low-entropy natural capital into high-entropy waste [15,16,17], overriding Earth’s self-regulatory mechanisms and triggering dangerous feedback loops.
By 2024, this feedback—particularly cryospheric albedo loss amplifying Arctic warming [68,69,161]—had pushed global temperatures to 1.3 °C above pre-industrial levels, marking the hottest year in recorded history [32].
The consequences manifested as synchronized, compounding extremes across all subsystems:
  • Thermal domination: Japan recorded its hottest year (1.48 °C above baseline) [162,163], while Mali and Iran endured near −50 °C heatwaves, exposing millions to lethal conditions.
  • Hydrological chaos: record ocean heat fueled six consecutive typhoons in the Northern Philippines [164], while Pakistan received 470% above-average rainfall, triggering catastrophic floods that displaced millions [162].
  • Compound disasters: Australia faced concurrent cyclones, 49.9 °C heat, and biblical flooding [163]; South American wildfires scorched 86 million hectares, with Brazil losing 37.6 million hectares of vital ecosystems [165].
  • Biospheric collapse: fishery depletion and pollinator declines now threaten 75% of global food crops [55,166], while events like Thailand’s 44.2 °C heatwaves [155], Vietnam’s USD 3.26 billion typhoon devastation [167], and Italy’s deadly floods [162,168] exemplify systemic destabilization.
Industrial–urban metabolism has effectively hacked Earth’s operating system, breaching multiple planetary boundaries and activating irreversible feedback loops. With temperatures accelerating toward 1.5 °C and critical tipping points, for example, AMOC collapse and permafrost thaw looming, humanity faces an existential imperative: fundamentally restructure socioeconomic systems to operate within Earth’s carrying capacity or confront synchronized collapse of vital planetary systems.

8.1. PathwayRegeneration: Policy Implications for Earth System Stabilization

Validating the conceptual framework from Section 3, Table 5 and Table 6 demonstrate how Multi-Scalar Nature-Based Regenerative Solutions (M-NbRSs) counteract industrial–urban metabolic disruption through targeted interventions across five key spheres: urban systems, biosphere integrity, lithosphere rehabilitation, hydrosphere restoration, and climate regulation.
Critically, this framework positions resilience metrics as essential complements to GDP—not replacements—by quantifying how economic activities (Table 2, Column 1) translate to biophysical outcomes (Table 2, Column 4). Singapore’s biodiversity offsets exemplify this balance: developers must achieve net-positive habitat gains, for example, +30% native canopy cover (Table 6) while maintaining economic activity. This dual-accountability approach aligns with the ‘Ecological Reciprocity’ principle introduced in Section 3, ensuring GDP growth remains constrained by planetary boundary thresholds
These solutions—ranging from urban forest cooling to Indigenous wisdom in environmental management—collectively address the root causes of Earth system destabilization while delivering measurable co-benefits for biodiversity, carbon sequestration, and human resilience. Indigenous wisdom encompasses place-based knowledge systems developed over millennia, including ecological stewardship practices such as rotational farming and seed banking that enhance biodiversity and soil resilience. As demonstrated by the Heart of Borneo Initiative (Table 6), which protects 24 million hectares of biologically diverse ecosystems, Indigenous wisdom contributes significantly to enhanced forest management and ecological integrity. When integrated with scientific approaches, these time-tested practices offer powerful pathways for planetary regeneration.
Table 5. Urban nature-based regenerative solutions: climate mitigation and biodiversity revival in built environments.
Table 5. Urban nature-based regenerative solutions: climate mitigation and biodiversity revival in built environments.
Multi-Scalar Nature-Based Regenerative Solutions (M-NbRSs)—Industrial–Urban System-Based
Function/Role/MeansMechanismCase StudyOutcome
Urban greening for climate mitigation
Urban forests and street treesHeat island reductionTree transpiration and blockage of solar radiation Global assessment of tree-related cooling effectsLowers pedestrian temperature by up to 12 °C through large radiation blockage and transpiration [169].
Reduces peak monthly temperatures to below 26 °C in 83% of the cities [169].
European cities: cooling
effects of urban trees
Increasing tree coverage to 30%: cool cities by a mean of 0.4 °C [170].
European cities: urban heat island (UHI) effect Urban trees cool European cities by an average of 1.1 °C and up to 2.9 °C [171].
Global urban climate change mitigationGlobal urban trees store approximately 7.4 billion tons of carbon and sequester approximately 217 million tons of carbon (796 million CO2e/year annually [172].
Carbon sequestration (tree planting) Trees absorb CO2 through photosynthesisSequester 217 M tons CO2/year globallyMalaysia 100 million trees: capable of absorbing 22 kg of CO2 per year, contributing to local and regional warming mitigations [173,174].
China Planted 142.6 billion trees: substantial increase in carbon storage (873.1 ± 16.2 Tg C in planted forests (1.87 billion tons ± 16 million tons), contributing to mitigating local, regional and global warming [175,176,177].
Biodiversity revival in cities
Nest boxesInstalled in trees for breeding birdsSingapore Hornbill population recovery [178].
Nature corridorsNative plant routes connecting forest fragmentsSingapore Pumas, golden tamarins recolonized urban edges [178].
River cleanupPollution control and habitat restorationSingapore Smooth-coated otters returned to urban waterways [178].
Wetland conservation 400 ha network at Sungei Buloh ReserveSingapore Hosts >50% of Singapore’s bird species [179].
Global urban cooling technologies
Complementary green infrastructureGreen roofsInsulates buildings, reduces runoffBasel, Switzerland (green roofs through financial incentives and legal mandates3–7 °C surface cooling [180].
Cairo city 0.5 °C to 6 °C reduction in air temperature on rooftops [181].
Chongqing, Tokyo, and Hangzhou Reduces pedestrian temperatures by 0.5 °C, 0.1 °C, and 0.5 °C, respectively [181].
Beijing Reduces the daytime average air temperature by a maximum of 0.41 °C [181].
Cool pavementsLight-colored, high-albedo materials for multi-layer pavements or a combination of high-albedo and permeable or reflective pavements Urban climate researchReduces surface temps by 5–20 °C [182,183,184].
Cool pavement paired with trees for shade synergy City of Phoenix cool corridor5.4 °C cooler under trees and 4.4 °C cooler within building canyons compared to open areas [185,186,187].
City of Tokyo Cool pavements reduce road temperatures by up to 10 °C [185,186,187].
Cool pavement paired with tree canopyUrban microclimate simulation (Hangzhou, China)Boosts cooling effects to 10–15 °C [188].
Cool pavement paired with canopy Urban climate adaptation strategy (Madison, Wisconsin)Canopy cover exceeded 40%: greatest cooling [189].
Table 6. Earth system stabilization strategies: biosphere–lithosphere–hydrosphere interventions for planetary resilience.
Table 6. Earth system stabilization strategies: biosphere–lithosphere–hydrosphere interventions for planetary resilience.
Multi-Scalar Nature-Based Regenerative Solutions (M-NbRSs)—Earth System-Based
SolutionMechanismCase StudyOutcome
Biosphere: biodiversity and habitat restoration
Deforestation/habitat fragmentation and biodiversity conservationReforestationReconnects fragmented habitats, reduces edge effects.Reserva Biológica União, Brazil (250 m × 1000 m corridor)Pumas and golden lion tamarins returned within 5 years [190].
Wildlife corridorsEnable species migration and genetic exchange.Yellowstone-to-Yukon (3200 km)Grizzly bear populations increased by 50% since 1997 [190].
Kavango–Zambezi Transfrontier Conservation Area (5 countries)Enhanced elephant migration routes [190].
Biodiversity conservation Wildlife habitat protection/species extinctionIndonesia’s Katingan Mentaya Project (protecting and restoring 149,800 ha peat swamp forests) Offer protection for the critically endangered Bornean orangutan the endangered proboscis monkey and the endangered Bornean white-bearded gibbon [191,192].
Transboundary ecological conservationCreation of a 24 million ha totally protected area (the Heart of Borno) for sustainable biodiversity managementIndonesia–Malaysia–Brunei Heart of Borneo initiativeConserve the ecological integrity of HoB transboundary wildlife connectivity corridors to protect various endangered species such as Bornean orangutans and proboscis monkeys [193].
Mining impact mitigationOn-site restorationMicrohabitat creation (hollow logs, soil microbiome revival)Lusatian Lakes, Germany (flooded lignite mines)More than 1000 aquatic species returned; new canals support fish populations [194].
Indigenous co-managementImpact–benefit agreements (IBAs) with traditional ecological knowledgeImpact–benefit agreements (IBAs) with traditional ecological knowledgeDiamond mines northeast of Yellowknife, CanadaImproved environmental monitoring and wealth-sharing with First Nations [195,196,197].
Tropical rainforest conservation integrating Indigenous wisdomApplication of Indigenous environmental knowledge (IEK) and traditional ecological knowledge (TEK) developed over millenniaHeart of Borneo Initiative (transboundary conservation across Indonesia, Malaysia, and Brunei)Maintenance of ecological integrity across 24 million hectares of forest ecosystems [193].
Lithosphere: soil and land degradation
Mining rehabilitationSoil microbiome restorationReintroduces native microbes for nutrient cyclingWestern Australia (Jarrah Forest, Swan Coastal Plain, Pilbara)Higher organic matter decomposition and soil moisture retention [198].
Constructed wetlandsRemoves heavy metals (Pb, Cd, Fe, Ni, Cr, Cu) via phytoremediationGadoon Amazai Industrial Estate, PakistanHeavy metal reductions: Pb (50%), Cd (91.9%), Fe (74.1%), and Ni (40.9%) [199].
Sustainable land managementConservation tillageDisturbs soil minimally to retain moisture and organic matterWest of Lake Balaton in southwest HungaryConservation tillage decreased surface runoffs by 75% and soil loss by 95% [200].
Diversified crop rotationBreaks pest cycles and enhances soil fertilityNorth China PlainReduced N2O emissions by 39%, improved system’s GHG balance by 88%, stimulated soil microbial activities by including legumes in crop rotations, increased soil organic carbon stocks by 8%, and enhanced soil health [201].
Organic farmingEliminates synthetic inputs to improve soil healthIFOAM Organics Europe case study (a) Can reduce global agricultural GHG emissions by around 20%, (b) can reduce nitrous oxide and methane emissions from manure by 50% and 70%, respectively, (c) consumes around 15% less energy per unit produced compared to conventional agriculture, (d) supports rare insects and spiders, increasing their abundance by 55% and their diversity by 27% compared to conventional farming, (e) reduces soil erosion and soil loss by 22% and 26%, respectively, and (f) improves soil quality and fertility, contributing significantly to higher soil organic carbon sequestration [202].
Hydrosphere: water systems
Coastal and freshwater restorationMangrove restorationCoastal protection and carbon sequestrationIndonesia peatland The restoration of 2.3 million ha of targeted peatland ecosystems: estimated to reduce carbon emissions by 98.77–153.53 MtCO2e [203].
Indonesia mangrove restoration Mangrove swamp restoration in the coastal villages (Sriwulan, Bedono, Timbulsloko, and Surodadi) in Central Java: supports a higher macrobenthic faunal diversity and reduces wave action in highly eroded areas [204].
Coral reef restorationTransplanting heat-resistant corals to mitigate bleachingGreat Barrier Reef, AustraliaPromotes faster recovery from chemical bleaching, enhances coral’s heat tolerance, and reduces sea surface temperature rise/ocean warning [205].
Floodplain managementSustainable floodplain managementReduces flooding and improves water qualityEuropean case studyIncreases natural water retention in flood pian areas, reduces flood risks, and improves water quality, carbon sequestration, habitat, and biodiversity protection [206].
Ocean acidification and ocean warming mitigationGHG/CO2 removalMobilizing actions to limit global GHG emissions, which is critical to reducing global warming, influencing ocean heat absorption, ocean warming, and decreases in the ocean’s pHChina’s climate pledge China reduced its CO2 emissions per unit of GDP (emissions intensity) by 48.4% in 2020 compared to the 2005 level and increased its share of non-fossil fuel in primary energy consumption to 15.9% in the same year. Coal-fired power plants declined by nearly 80% in the first half of 2021, compared to the same period in 2020 [207].
Coastal erosion management Coastal protection and conservation (to protect the hydrosphere by maintaining coastal ecosystem’s health, regulate water quality, and prevent flooding)Nearshore nourishmentKlaipėda Port (Baltic Sea)Nearshore nourishment has successfully stabilized the subaerial coast with 21.1% of the nourished sand accumulating offshore and about 69.4% of the nourished sand remaining at the nourishment site [208].
Managing coastal water and marine resource sustainabilityTackling overfishing and ocean climate impacts in Britain, Portugal, Greece, Turkey, Mexico, the Philippines, and IndonesiaBuilding marine protected areas and reviving marine life Assists local communities in establishing “marine protected areas” in coastal waters. This will contribute to protecting at least 30% of the world’s oceans by the end of the decadeAssists local communities in establishing “marine protected areas” in coastal waters. This will contribute to the protection of at least 30% of the world’s oceans by the end of the decade [209,210].
Atmosphere: climate regulation
Carbon sequestrationPeatland preservationRewetting and conserving tropical peatlands, preventing CO2 release (stores 105 GtC globally)Congo Basin (29 GtC) and Indonesia (55–57 GtC) Avoids 1.9 GtCO2 emissions annually [211,212,213].
Forest conservationProtects old-growth forests (stores 170.7 GtC in biomass, 49.3 GtC in soils)Global tropical forestsCaptures 226 GtC (1/3 of industrial-era emissions) [214,215].
Cryosphere: glacier, ice, and permafrost
Albedo and methane management to enhance negative feedback to global warmingPermafrost insulationRewilding permafrost to insulate the soil from temperature rise/fluctuationsEndalen and Adventdalen, located in the vicinity of Longyearbyen, SvalbardHigh Arctic mosses and vascular vegetation slow active layer permafrost thaw, shading incoming sunlight and creating an insulating layer [216,217,218,219].
Safeguarding and prolonging the lifetime of glaciersAnthropogenic CO2 emissions reduction Reducing human-induced warming/decoupling of economic growth from GHG emissionsCopenhagen’s transition to a bicycle-friendly city42% CO2 reduction compared to 2005 levels [220].
Costa Rica’s transition to renewable energy98% of electricity from renewable sources [221].
Ørsted renewable energy transition and circular economy, with installed renewable capacity increased by 16% to 18.2 GW in 2024.(i) About 80% reduction in coal consumption since 2006; (ii) 83% carbon emissions reduction since 2006 [222,223].

8.2. Operationalizing Earth System Regeneration: Evidence from Multi-Scalar Solutions (Table 5 and Table 6)

The documented destabilization of Earth’s systems demands urgent implementation of the Multi-Scalar Nature-Based Regenerative Solution (M-NbRS) framework, as systematically cataloged in Table 5 and Table 6. These tables synthesize the peer-reviewed case studies demonstrating actionable pathways across urban to planetary scales, with measurable outcomes for climate mitigation, biodiversity revival, and socio-ecological resilience. This is briefly elucidated as follows:
  • Urban-scale efficacy (Table 5)
    Table 5 quantifies how nature-based urban design can counteract industrial–urban metabolic damage: heat mitigation—street trees reduce pedestrian temperatures by 12° C (radiation blocking) [169], while Basel’s mandated green roofs achieved 3–7 °C cooling [180].
    »
    Circular metabolism: China’s 142 billion trees now sequester 1.87 GtC, demonstrating scalable carbon-negative urbanization [175,176,177].
    »
    Biodiversity revival: Singapore’s river cleanup restored otter populations and 50% of native bird species through wetland corridors [178,179].
  • Planetary repair strategies (Table 6)
    Table 6 documents interventions to stabilize critical Earth systems:
    »
    Biosphere: Wildlife corridors increased Yellowstone grizzly bear populations by 50% [190]; organic farming cut N2O emissions by 90% while boosting soil carbon [201,202].
    »
    Hydrosphere: Java’s mangroves sequester 133 MtCO2e/year while reducing storm surges [203,204].
    »
    Cryosphere: Svalbard’s permafrost insulation slowed thaw by 30–50% [216,217,218,219].
  • Governance lever
    The tables further highlight policy enablers:
    »
    Legal frameworks: Basel’s green roof mandates [180] versus Canada’s Indigenous co-management [195,196,197];
    »
    Metrics: Biophysical indicators such as soil carbon stocks [202] outperforming GDP.

8.3. Translating M-NbRS into Governance

The successful implementation of Multi-Scalar Nature-Based Regenerative Solutions (Table 5 and Table 6) requires aligning biophysical interventions with human behavioral drivers. Empirical evidence from Akotia et al. [224] demonstrates that built environment practitioners consistently prioritize reputation enhancement and competitive advantage, motivations that can be strategically channeled through three key policy mechanisms:
  • Supply chain due diligence for deforestation-free procurement, linking municipal financing to ecological performance targets through sustainability-linked bonds;
  • Economic instruments must be redesigned to incentivize circularity, including tax rebates for industries achieving high material recovery rates and the integration of biophysical indicators alongside traditional economic metrics;
  • Ecological reciprocity mandates (requiring cities to fund transboundary conservation efforts using metropolitan taxes).
This integrated approach transforms urban–industrial systems at multiple levels: cities evolve from net environmental burdens to nodes of ecological renewal as demonstrated by Singapore’s biodiversity offset programs. This aligns human incentives with planetary boundaries, synchronizing urban–industrial metabolism with Earth’s regenerative capacity. Institutionalizing this alignment transforms cities from destabilizing forces into nodes of ecological renewal; industries from linear extractors to circular regenerators through the 3-R principles (Reduce, Reuse, Recycle); and economies from boundary-breachers to boundary-keepers operating within planetary thresholds This recalibration empirically validates the M-NbRS translation mechanism (Section 3) through policy-implementation metrics in Table 5 and Table 6. Institutionalizing this alignment, this recalibration is not idealism but survival logic: operate within planetary thresholds, or forfeit the stability enabling civilization itself.

9. Conclusions

By bridging the implementation gap between local NbSs and planetary boundaries, M-NbRSs offer an original policy-actionable framework where
  • Street trees (local) are incentivized via carbon markets tied to global ice-loss thresholds (Section 7.2);
  • Supply chain laws (global) protect forests that cool cities (regional), as demonstrated in Table 6.
These innovations are critical to counteracting the three destabilizing mechanisms analyzed in this study: (1) amplifying feedback, for example, heat islands → rising energy demand [169,180]), (2) entropic degradation (100 billion tons/year of high-entropy waste [2]), and (3) cross-scale cascades (local deforestation → global CO2 rise). While occupying just 3% of Earth’s surface, these systems function as thermodynamic engines of planetary disruption, converting low-entropy natural capital into irreversible waste while triggering biosphere collapse.
Our Vicious Cycle Atlas of Fragility reveals that escaping this spiral demands more than incremental fixes; it requires a metabolic revolution that redefines humanity’s role from extractor to regenerator within Earth’s operating systems. Industrial–urban systems must evolve from industrial extractive enclaves or urban metabolism to regenerative nodes within Earth’s systems. This transformation hinges on immediate, systemic adoption of Multi-Scalar Nature-Based Regenerative Solutions (M-NbRSs), implemented through six actionable pillars:
  • Legally mandate multi-scalar NbRSs
    • Local tier: Enact and enforce zoning laws requiring 30% urban green cover, which is projected to reduce temperatures by 0.4 °C and cool infrastructure such as green roofs and permeable pavements in all cities by 2040, modeled on Basel’s binding mandates. Link infrastructure funding to biophysical metrics. For example, cities that achieve ≥30% tree canopy cover or proven reductions in urban heat island intensity from tree canopy/green roofs qualify for additional funding.
    • Regional tier: Establish transboundary “ecological reciprocity,” requiring cities to fund hinterland watershed protection.
    • Global tier: Ratify a treaty to protect 30% of oceans and critical carbon sinks such as peatlands and mangroves by 2030, funded by reallocating USD 540 billion/year in fossil fuel subsidies.
  • Enforce supply chain accountability
    »
    Adopt EU-style extraterritorial regulations to penalize corporations sourcing from deforested lands or violating biosphere integrity boundaries.
  • Institutionalize indigenous wisdom of environmental conservation
    »
    Formalize Indigenous wisdom, that is, Indigenous environmental knowledge and traditional ecological knowledge (TEK), in co-management of biodiversity corridors such as the Canadian impact–benefit agreements and scale-up models like the Heart of Borneo Initiative to protect 24 million hectares of ecosystem diversity across three nations (Indonesia, Malaysia, and Brunei).
  • Implementation pathways for biophysical metrics integration
    »
    Urban climate mitigation: street tree coverage, for example, 12 °C cooling efficacy [169] and green roof implementation (3–7 °C reduction [180]);
    »
    Afforestation progress tracking: afforestation progress, for example, China’s 142 billion trees storing 1.87 billion tons of CO2 [175,176,177];
    »
    Biodiversity recovery: species revival metrics, for example, Singapore’s otter/hornbill populations [178].
  • Legally enforce “Rights of Nature” frameworks
    »
    Basel’s green roofs: legally required cooling infrastructure (reduces surface temperature by 3–7 °C);
    »
    Singapore’s ecological reciprocity: binding urban–hinterland conservation such as watershed protection.
  • Policy-linked fiscal instrument
    »
    Electric mobility law (Law 9518/2018): tax exemptions tied to EV adoption rates and charging infrastructure rollout as in Costa Rica;
    »
    Singapore’s biodiversity impact offset: require mitigation for approved development projects through habitat restoration [179].

9.1. Scientific Limitations and Critical Research Frontier

While this Atlas advances the mapping of urban–Earth system feedback, critical knowledge gaps undermine actionable solutions. Three interrelated frontiers demand prioritized research to avert systemic collapse:
  • Problem: a 40 to 60 year lag between urban GHG emissions and cryosphere responses, for example, ice sheet collapse or permafrost thaw, means current climate models underestimate committed warming by 0.5–1.2 °C.
  • Implication: urban decarbonization targets, for example, net-zero by 2050, may fail to prevent locked-in sea-level rise and methane releases.
  • Research imperative: develop next-generation Earth system models that integrate delayed cryosphere feedback into urban climate adaptation frameworks.

9.2. Pathways for Real-World Implementation

The “Vicious Cycle Atlas” provides a foundational framework for diagnosing cross-scale environmental feedback, but its real-world utility can be expanded through (1) dynamic modeling tools that integrate real-time data, for example, satellite-based deforestation alerts and urban GHG inventories, to visualize feedback loops with localized precision; (2) policy integration platforms that link atlas metrics to governance mechanisms, such as automated triggers for supply chain due diligence when planetary thresholds are breached; and (3) community-engaged adaptations, where Indigenous knowledge refines feedback pathways. Pilot applications in cities, like Singapore’s urban–hinterland reciprocity or the Congo Basin’s deforestation–carbon cascades, could test its capacity to bridge scientific rigor with actionable interventions.

9.3. Concluding Remarks

Our analysis demonstrates that industrial–urban systems function as thermodynamic engines of planetary disruption through three empirically validated mechanisms: (1) the quantification of metabolic impacts, including 100 billion tons/year of material demand and 61.56 million hectares of tropical deforestation; (2) documentation of cross-scale feedbacks such as urban heat islands (0.4–12 °C warming) triggering increased energy demand and CO2 emissions; and (3) threshold analysis revealing critical boundary breaches (424.61 ppm atmospheric CO2, 372 zettajoules of ocean heat accumulation). These findings necessitate urgent implementation of the Multi-Scalar Nature-Based Regenerative Solution (M-NbRS) framework supported by proven precedents ranging from Costa Rica’s near-total renewable energy transition to Singapore’s successful urban river ecosystem restoration. While the technical blueprints for transformation exist and have been rigorously validated, the critical barrier remains the mobilization of political will to implement these solutions at scale.
Through systematic analysis of biophysical thresholds, economic trade-offs, and governance models, this research delivers three essential elements to motivate and guide policy action: (1) compelling rationales grounded in the irreversible risks of delayed implementation, particularly the approaching tipping points in climate and biodiversity systems; (2) practical guidelines distilled from successful case studies across urban, regional and global scales; and (3) motivational evidence demonstrating the economic and social co-benefits of timely intervention. The frameworks and metrics presented here transform abstract ecological principles into actionable policy instruments, enabling decision-makers to confidently pursue urban regeneration while addressing legitimate concerns about economic stability and development priorities.
The choice facing policymakers is no longer between environmental protection and economic growth, but between proactive stewardship of our urban–planetary systems or costly, reactive management of their collapse. This study provides the scientific foundation, implementation tools, and ethical imperative to choose the former path—one that leads toward stable, equitable, and sustainable civilizations in balance with Earth’s life-support systems. With the M-NbRS framework, we possess both the knowledge and means to enhance this transformation; what remains is the collective will to act upon it.

Funding

This research received no external funding.

Data Availability Statement

The data presented in this study are openly available as per reference list.

Conflicts of Interest

The author declares no conflicts of interest.

References

  1. UN-Habitat. Energy and Resource Efficient Urban Neighbourhood Design Principles for Tropical Countries. In A Practitioner’s Guidebook; United Nations Human Settlements Programme (UN-Habitat): Nairobi, Kenya, 2018. [Google Scholar]
  2. Circular Economy Foundation. Circularity Gap Report 2023; Circular Economy Foundation: Brussels, Belgium, 2023. [Google Scholar]
  3. Luqman, M.; Rayner, P.J.; Gurney, K.R. On the impact of urbanisation on CO2 emissions. npj Urban Sustain. 2023, 3, 6. [Google Scholar] [CrossRef]
  4. Mukim, M.; Roberts, M. (Eds.) Thriving: Making Cities Green, Resilient, and Inclusive in a Changing Climate; World Bank: Washington, DC, USA, 2023. [Google Scholar] [CrossRef]
  5. Weisse, M.; Goldman, E.; Carter, S. Forest Pulse: The Latest on the World’s Forests; World Resource Institute: Washington, DC, USA, 2024; Available online: https://gfr.wri.org/latest-analysis-deforestation-trends (accessed on 5 April 2025).
  6. Global Forest Watch. Forest Monitoring Designed for Action; Washington, DC, USA, 2025; Available online: https://www.globalforestwatch.org/ (accessed on 5 April 2025).
  7. EEA. How Climate Change Impacts Marine Life; European Environmental Agency (EEA): Copenhagen, Denmark, 2023; Available online: https://www.eea.europa.eu/publications/how-climate-change-impacts (accessed on 8 April 2025).
  8. Ayres, R.U. Industrial metabolism: Theory and policy. In The Greening of Industrial Ecosystems; Allenby, B.R., Richards, D.J., Eds.; National Academy Press: Washington, DC, USA, 1994; pp. 23–37. [Google Scholar]
  9. Wernick, I.K. Industrial Metabolism. In International Encyclopedia of Social and Behavioral Sciences; Smelser, N.J., Paul, B., Baltes, P.B., Eds.; Pergamon: Oxford, UK, 2001; pp. 7331–7333. [Google Scholar]
  10. Martín-Gómez, A.M.; Ávila-Gutiérrez, M.J.; Lama-Ruiz, J.R.; Aguayo-González, F. Industrial metabolism: A multilevel characterization for designing sustainable manufacturing systems. Machines 2024, 12, 16. [Google Scholar] [CrossRef]
  11. Young, J.T. Is the entropy law relevant to the economics of natural resource scarcity? J. Environ. Econ. Manag. 1991, 21, 169–179. [Google Scholar] [CrossRef]
  12. Anderberg, S. Industrial metabolism and the linkages between economics, ethics and the environment. Ecol. Econ. 1998, 24, 311–320. [Google Scholar] [CrossRef]
  13. Zanjani, N.E.; Salem, Y.; Qadi, S.; Hammoud, S. The study of urban metabolism and its role in urban planning. Int. J. Urban Manag. Energy Sustain. 2021, 2, 27–42. [Google Scholar] [CrossRef]
  14. Galán-Cano, L.; Mena-Nieto, A.; Cámara-Aceituno, J.; Hermoso-Orzáez, M.J.; Terrados-Cepeda, J. Urban metabolism, sustainability and energy transition in cities: A comprehensive review. Results Eng. 2025, 25, 104278. [Google Scholar] [CrossRef]
  15. Georgescu-Roegen, N. Energy and economic myths. South. Econ. J. 1975, 41, 347–381. [Google Scholar] [CrossRef]
  16. Georgescu-Roegen, N. The Entropy Law and the Economic Process; Harvard University Press: Cambridge, MA, USA; London, UK, 1976; Volume 1408, p. 49. [Google Scholar]
  17. Swendse, R.H. Probability, Entropy, and Gibbs’ Paradox(es). Entropy 2018, 20, 450. [Google Scholar] [CrossRef]
  18. Richardson, K.; Steffen, W.; Lucht, W.; Bendtsen, J.; Cornell, S.E.; Donges, J.F.; Drüke, M.; Fetzer, I.; Bala, G.; von Bloh, W.; et al. Earth beyond six of nine planetary boundaries. Sci. Adv. 2023, 9, 37. [Google Scholar] [CrossRef]
  19. Wang, S.; Foster, A.; Lenz, E.A.; Kessler, J.D.; Stroeve, J.C.; Anderson, L.O.; Turetsky, M.; Betts, R.; Zou, S.J.; Liu, W.; et al. Mechanisms and impacts of Earth system tipping elements. Rev. Geophys. 2023, 61, e2021RG000757. [Google Scholar] [CrossRef]
  20. Liu, T.; Chen, D.; Yang, L.; Meng, J.; Wang, Z.C.; Ludescher, J.; Fan, J.F.; Yang, S.I.; Chen, D.L.; Kurths, J.; et al. Teleconnections among tipping elements in the Earth system. Nat. Clim. Change 2023, 13, 67–74. [Google Scholar] [CrossRef]
  21. Steffen, W.; Richardson, K.; Rockström, J.; Cornell, S.E.; Fetzer, I.; Bennett, E.M.; Biggs, R.; Carpenter, S.R.; de Vries, W.; de Wit, C.A.; et al. Planetary boundaries: Guiding human development on a changing planet. Science 2015, 347, 6223. [Google Scholar] [CrossRef] [PubMed]
  22. Friant, M.C.; Reid, K.; Boesler, P.; Vermeulen, W.J.V.; Salomone, R. Analyzing urban circular economy policies in Amsterdam, Glasgow, and Copenhagen. Local. Environ. 2023, 28, 1331–1369. [Google Scholar] [CrossRef]
  23. Ragazou, K.; Zournatzidou, G.; Sklavos, G.; Sariannidis, S. Integration of Circular Economy and Urban Metabolism for a Resilient Waste-Based Sustainable Urban Environment. Urban Sci. 2024, 8, 175. [Google Scholar] [CrossRef]
  24. Cialani, C.; Mortazavi, R.; Sarcinella, F. Material Footprint and Circular Economy for a Sustainable Consumption and Production Pattern. Circ. Econ. Sustain. 2025, 1–20. [Google Scholar] [CrossRef]
  25. World Bank. A Catalogue of Nature-Based Solutions for Urban Resilience; World Bank Group: Washington, DC, USA, 2021. [Google Scholar]
  26. Chowdhury, K.; Basu, S.; Pramanik, M.; Plieninger, T. Blue infrastructure as nature-based solutions for urban sustainability: Evaluating local perceptions from four Indian megacities. Nat.-Based Solut. 2025, 7, 100211. [Google Scholar] [CrossRef]
  27. McPhearson, T.; Cook, E.M.; Berbés-Blázquez, M.; Cheng, C.W.; Grimm, N.B.; Andersson, A.; Barbosa, O.; Chandler, D.G.; Chang, H.J.; Chester, M.V.; et al. A social-ecological-technological systems framework for urban ecosystem services. One Earth 2022, 5, 505–518. [Google Scholar] [CrossRef]
  28. Crippa, M.; Guizzardi, D.; Pagani, F.; Banja, M.; Muntean, M.; Schaaf, E.; Monforti-Ferrario, F.; Becker, W.E.; Quadrelli, R.; Martin, R.; et al. GHG Emissions of All World Countries—JRC/IEA; Publications Office of the European Union: Luxembourg, 2024; JRC138862. [Google Scholar] [CrossRef]
  29. Energy Institute. Statistical Review of World Energy 2024; Energy Institute: London, UK, 2024. [Google Scholar]
  30. Giljum, S.; Maus, V.; Kuschnig, N.; Luckeneder, S.; Tost, M.; Sonter, L.J.; Bebbington, A.J. A pantropical assessment of deforestation caused by industrial mining. Proc. Natl. Acad. Sci. USA 2022, 119, e2118273119. [Google Scholar] [CrossRef]
  31. Lan, X.; Keeling, R. Trends in Atmospheric Carbon Dioxide (CO2); U.S. Department of Commerce, National Oceanic & Atmospheric Administration: Washington, DC, USA, 2025. Available online: https://gml.noaa.gov/ccgg/trends/ (accessed on 10 April 2025).
  32. NASA. Ocean Warming; The National Aeronautics and Space Administration (NASA): Washington, DC, USA, 2024. Available online: https://climate.nasa.gov/vital-signs/ocean-warming/?intent=121 (accessed on 10 April 2025).
  33. Slater, T.; Lawrence, I.R.; Otosaka, I.N.; Shepherd, A.; Gourmelen, N.; Jakob, L.; Tepes, P.; Gilbert, L.; Nienow, P. Review article: Earth’s ice imbalance. Cryosphere 2021, 15, 233–246. [Google Scholar] [CrossRef]
  34. Harvey, C. Earth Has Lost 28 Trillion Tons of Ice Since the Mid-1990s. Scientific American. 2021. Available online: https://www.scientificamerican.com/article/earth-has-lost-28-trillion-tons-of-ice-since-the-mid-1990s (accessed on 10 April 2025).
  35. IPBES. Summary for Policymakers of the Global Assessment Report on Biodiversity and Ecosystem Services of the Intergovernmental Science-Policy Platform on Biodiversity and Ecosystem Services; International Platform for Biodiversity and Ecosystem Services (IPBES): Bonn, Germany, 2019. [Google Scholar]
  36. Cabernard, L.; Pfister, S.; Hellweg, S. Biodiversity impacts of recent land-use change driven by increases in agri-food imports. Nat. Sustain. 2024, 7, 1512–1524. [Google Scholar] [CrossRef]
  37. Ford, D.N. A system dynamics glossary. Syst. Dyn. Rev. 2019, 35, 369–379. [Google Scholar] [CrossRef]
  38. Checkland, P. Systems Thinking, Systems Practice; John Wiley & Sons: New York, NY, USA, 1993. [Google Scholar]
  39. Fischer, P.; Sembritzki, K.; Maier, A. System Theory. In Medical Imaging Systems. An Introductory Guide; Maier, A., Steidl, S., Christlein, V., Hornegger, J., Eds.; Spinger Open: Cham, Switzerland, 2018; pp. 13–36. [Google Scholar]
  40. Lawrencea, J.; Blackettb, P.; Cradock-Henry, N.A. Cascading climate change impacts and implications. Clim. Risk Manag. 2020, 29, 100234. [Google Scholar] [CrossRef]
  41. Schulz, K.G.; Bach, L.T.; Bellerby, R.G.J.; Bermúdez, R.; Büdenbender, J.; Boxhammer, T.; Czerny, J.; Engel, A.; Ludwig, A.; Meyerhöfer, M.; et al. Phytoplankton Blooms at Increasing Levels of Atmospheric Carbon Dioxide: Experimental Evidence for Negative Effects on Prymnesiophytes and Positive on Small Picoeukaryotes. Front. Mar. Sci. 2017, 4, 64. [Google Scholar] [CrossRef]
  42. Gowdy, J.; Susan Mesner, S. The Evolution of Georgescu-Roegen’s Bioeconomics. Rev. Soc. Econ. 1998, 56, 136–156. [Google Scholar] [CrossRef]
  43. Talwar, P.; Verma, N.; Khatri, H.; Ahire, P.D.; Chaudhary, G.; Lindenberger, C.; Vivekanand, V. A systematic review of photovoltaic-green roof systems in different climatic conditions focusing on sustainable cities and societies. Sustain. Cities Soc. 2023, 98, 104813. [Google Scholar] [CrossRef]
  44. Persson, L.; Carney Almroth, B.M.; Collins, C.D.; Cornell, S.; de Wit, C.A.; Diamond, M.L.; Fantke, P.; Hassellöv, M.; MacLeod, M.; Ryberg, M.W.; et al. Outside the Safe Operating Space of the Planetary Boundary for Novel Entities. Environ. Sci. Technol. 2022, 56, 1510–1521. [Google Scholar] [CrossRef]
  45. Briffa, J.; Sinagra, E.; Blundell, R. Heavy metal pollution in the environment and their toxicological effects on humans. Heliyon 2020, 6, e04691. [Google Scholar] [CrossRef] [PubMed]
  46. Nyiramigisha, P.; Komariah; Sajidan. Harmful Impacts of Heavy Metal Contamination in the Soil and Crops Grown Around Dumpsites. Rev. Agric. Sci. 2021, 9, 271–282. [Google Scholar] [CrossRef]
  47. El-Sharkawy, M.; Alotaibi, M.O.; Li, J.; Du, D.; Mahmoud, E. Heavy Metal Pollution in Coastal Environments: Ecological Implications and Management Strategies: A Review. Sustainability 2025, 17, 701. [Google Scholar] [CrossRef]
  48. Choy, Y.K.; Onuma, A.; Lee, K.E. The Nexus of Industrial–Urban Sustainability, the Circular Economy, and Climate–Ecosystem Resilience: A Synthesis. Sustainability 2025, 17, 2620. [Google Scholar] [CrossRef]
  49. Joyce, S. The dead zones: Oxygen-starved coastal waters. Env. Health Perspect. 2000, 108, A120–A125. [Google Scholar] [CrossRef] [PubMed]
  50. Evich, M.G.; Davis, M.J.B.; McCord, J.P.; Acrey, B.; Awkerman, J.A.; Knappe, D.R.U.; Lindstrom, A.; Speth, T.F.; Tebes-Stevens, C.; Strynar, M.J.; et al. Per- and polyfluoroalkyl substances in the environment. Science 2022, 375, 6580. [Google Scholar] [CrossRef] [PubMed]
  51. Haba, B.; Djellali, S.; Abdelouahed, Y.; Boudjelida, S.; Faleschini, F.; Carraro, M. Transforming Plastic Waste into Value: A Review of Management Strategies and Innovative Applications in Sustainable Construction. Polymers 2025, 17, 881. [Google Scholar] [CrossRef]
  52. Steffen, W.; Richardson, K.; Rockström, J.; Schellnhuber, H.J.; Dube, O.P.; Dutreuil, S.M.; Lenton, T.M.; Lubchenco, J. The emergence and evolution of Earth System Science. Nat. Rev. Earth Environ. 2020, 1, 54–63. [Google Scholar] [CrossRef]
  53. Baatz, R.; Hendricks Franssen, H.J.; Euskirchen, E.; Sihi, D.; Dietze, M.; Ciavatta, S.; Fennel, K.; Beck, H.; Lannoy, G.D. Reanalysis in Earth system science: Toward terrestrial ecosystem reanalysis. Rev. Geophys. 2021, 59, e2020RG000715. [Google Scholar] [CrossRef]
  54. Rumjaun, A.B.; Borde, B.; Guilyardi, E.; Lescarmontier, L.; Matthews, R.; Niewöhner, C.; Pen Point, S.; Schlüpmann, J.; Wilgenbus, D. IPCC Special Report. Global Warming of 1.5 °C—Summary for Teacher; Office for Climate Education (OCE): Paris, France, 2018. [Google Scholar]
  55. Ritchie, H. What Is the Ozone Layer, and Why is it Important? OurWorldinData.org. 2023. Available online: https://ourworldindata.org/ozone-layer-context (accessed on 29 March 2025).
  56. Friedlingstein, P.; O’Sullivan, M.; Jones, M.W.; Andrew, R.M.; Gregor, L.; Hauck, J.; Le Quéré, C.; Luijkx, I.T.; Olsen, A.; Peters, G.P.; et al. Global Carbon Budget 2022. Earth Syst. Sci. Data 2022, 14, 4811–4900. [Google Scholar] [CrossRef]
  57. Farquhar, G.D.; Fasham, M.J.R.; Goulden, M.L.; Heimann, M.; Jaramillo, V.J.; Kheshgi, H.S.; Le Quéré, C.; Scholes, R.J.; Wallace, D.W.R.; Prentice, I.C.; et al. The Carbon Cycle and Atmospheric Carbon Dioxide. In Climate Change 2001: The Scientific Basis. Contribution of Working Group I to the Third Assessment Report of the Intergovernmental Panel on Climate Change (IPCC); Houghton, J.T., Ding, Y., Griggs, D.J., Noguer, M., Linden, P.J.V.D., Dai, X., Maskell, K., Johnson, C.A., Eds.; Cambridge University Press: Cambridge, UK, 2001; pp. 183–237. [Google Scholar]
  58. Bollmann, M.; Bosch, T.; Colijn, F.; Ebinghaus, R.; Froese, R.; Güssow, K.; Khalilian, S.; Krastel, S.; Körtzinger, A.; Langenbuch, M.; et al. Living with the Oceans. A Report on the State of the World’s Oceans; World Ocean Review; Mare: Hamburg, Germany, 2010. [Google Scholar]
  59. Bopp, L.; Bowler, C.; Guidi, L.; Karsenti, E.; de Varga, C. The Ocean: A Carbon Pump. In Ocean and Climate, Scientific Notes; 2015; pp. 13–18. Available online: www.ocean-climate.org (accessed on 30 March 2025).
  60. Wu, Y.X.; Hain, M.P.; Humphreys, M.P.; Hartman, S.; Toby Tyrrell, T. What drives the latitudinal gradient in open-ocean surface dissolved inorganic carbon concentration? Biogeosciences 2019, 16, 2661–2681. [Google Scholar] [CrossRef]
  61. Hansell, D.A.; Carlson, C.A.; Repeta, D.J.; Schlitzer, R. Dissolved organic matter in the ocean: A controversy stimulates new insights. Oceanography 2009, 22, 202–211. [Google Scholar] [CrossRef]
  62. Lindsey, R.; Scott, M. What are Phytoplankton? The Earth Observatory, EOS Project Science Office, NASA Goddard Space Flight Center: Glenn Dale, MD, USA, 2010. Available online: https://earthobservatory.nasa.gov/features/Phytoplankton (accessed on 30 March 2025).
  63. Bressac, M.; Laurenceau-Cornec, E.C.; Kennedy, F.; Santoro, A.E.; Paul, N.L.; Briggs, N.; Carvalho, F.; Boyd, P.W. Decoding drivers of carbon flux attenuation in the oceanic biological pump. Nature 2024, 633, 587–593. [Google Scholar] [CrossRef]
  64. Yamaguchi, R.; Kouketsu, S.; Kosugi, N.; Ishii, M. Global upper ocean dissolved oxygen budget for constraining the biological carbon pump. Commun. Earth Environ. 2024, 5, 732. [Google Scholar] [CrossRef]
  65. Blümel, M.; Fisch, K.; Franke, D.; Frey, T.; Froese, R.; Greinert, J.; Gutow, L.; Gutt, J.; Hain, S.; Haroon, A. The Ocean, Guarantor of Life—Sustainable Use, Effective Protection; World Ocean Review; Mare: Hamburg, Germany, 2021. [Google Scholar]
  66. Miner, K.R.; Turetsky, M.R.; Malina, E.; Bartsch, A.; Tamminen, J.; McGuire, A.D.; Fix, A.; Sweeney, C.; Elder, C.D.; Miller, C.E. Permafrost carbon emissions in a changing Arctic. Nat. Rev. Earth Environ. 2022, 3, 55–67. [Google Scholar] [CrossRef]
  67. NOAA. Trends in CO2, CH4, N2O, SF6; Global Monitoring Laboratory, U.S. Department of Commerce, National Oceanic & Atmospheric Administration (NOAA): Washington, DC, USA, 2025. Available online: https://gml.noaa.gov/ccgg/trends/ (accessed on 30 March 2025).
  68. Flanagan, E. The Global Carbon Budget and Permafrost Feedback Loops in the Arctic; The Arctic Institute: Oslo, Norway, 2021; Available online: https://www.thearcticinstitute.org/global-carbon-budget-permafrost-feedback-loops-arctic/ (accessed on 30 March 2025).
  69. CIS. Energy: The Driver of Climate. Climate Science Investigation (CIS); National Aeronautics and Space Administration (NASA): Washington, DC, USA, 2023. Available online: https://www.ces.fau.edu/nasa/module-2/earth-energy-balance.php (accessed on 1 April 2025).
  70. Serreze, M. Why Is the Arctic So Sensitive to Climate Change and Why Do We Care? NOAA Pacific Marine Environmental Laboratory, National Oceanic and Atmospheric Administration (NOAA): Newport, OR, USA. Available online: https://www.pmel.noaa.gov/arctic-zone/essay_serreze.html (accessed on 1 April 2025).
  71. National Ocean Service. What Is the Cryosphere? Department of Commerce, National Oceanic and Atmospheric Administration (NOAA): Washington, DC, USA, 2024. Available online: https://oceanservice.noaa.gov/facts/cryosphere.html#:~:text=The%20components%20of%20the%20cryosphere,to%20regulate%20our%20planet’s%20temperature (accessed on 1 April 2025).
  72. Bitz, C.M.; Gent, P.R.; Woodgate, R.A.; Holland, M.M.; Lindsay, R. The Influence of Sea Ice on Ocean Heat Uptake in Response to Increasing CO2. J. Clim. 2006, 19, 2437–2450. [Google Scholar] [CrossRef]
  73. Josey, S.A.; Meijers, A.J.S.; Blaker, A.T.P.; Grist, J.P.; Mecking, J.; Ayres, H.C. Record-low Antarctic Sea ice in 2023 increased ocean heat loss and storms. Nature 2024, 636, 635–639. [Google Scholar] [CrossRef] [PubMed]
  74. NASA. Thawing Permafrost Could Leach Microbes, Chemicals into Environment; Jet Propulsion Laboratory, California Institute of Technology, National Aeronautics and Space Administration (NASA): Washington, DC, USA, 2022. Available online: https://www.jpl.nasa.gov/news/thawing-permafrost-could-leach-microbes-chemicals-into-environment/ (accessed on 1 April 2025).
  75. Kim, D.K.; Kim, M.C.; Woo, S.H.; Nam, S.J.; Myeong, N.R.; Kim, E.B.; Lee, Y.M. Potential risks of bacterial plant pathogens from thawing permafrost in the Alaskan tundra. Ecotoxicol. Environ. Saf. 2025, 289, 117531. [Google Scholar] [CrossRef]
  76. Ruben, M.; Marchant, H.; Wietz, M.; Gentz, T.; Strauss, J.; Koch, B.P.; Mollenhauer, G. Microbial communities degrade ancient permafrost-derived organic matter in Arctic seawater. J. Geophys. Res. Biogeosci. 2024, 129, e2023JG007936. [Google Scholar] [CrossRef]
  77. Riebeek, H. The Carbon Cycle; The Earth Observatory, EOS Project Science Office, NASA Goddard Space Flight Center: Glenn Dale, MD, USA, 2011. Available online: https://earthobservatory.nasa.gov/features/CarbonCycle (accessed on 1 April 2025).
  78. Kevin Boyce, C.; Ibarra, D.E.; Antonio, M.P.D. What we talk about when we talk about the long-term carbon cycle. New Phytol. 2023, 237, 1550–1557. [Google Scholar] [CrossRef]
  79. Brantley, S.L.; Shaughnessy, A.; Lebedeva, L.; Balashov, V.N. How temperature-dependent silicate weathering acts as Earth’s geological thermostat. Science 2023, 379, 382–389. [Google Scholar] [CrossRef]
  80. Earth Observatory. The Slow Carbon Cycle; EOS Project Science Office, NASA Goddard Space Flight Center, National Aeronautics and Space Administration (NASA): Washington, DC, USA, 2011. Available online: https://earthobservatory.nasa.gov/features/CarbonCycle/page2.php (accessed on 1 April 2025).
  81. Dias Rodrigues, C.I.; Brito, L.M.; Nunes, L.J.R. Soil Carbon Sequestration in the Context of Climate Change Mitigation: A Review. Soil Syst. 2023, 7, 64. [Google Scholar] [CrossRef]
  82. Srivastava, R.K.; Yetgin, A. An overall review on influence of root architecture on soil carbon sequestration potential. Theor. Exp. Plant Physiol. 2024, 36, 165–178. [Google Scholar] [CrossRef]
  83. Royle, S. Ancient carbon released. Nat. Clim. Change 2024, 14, 896. [Google Scholar] [CrossRef]
  84. Callahan, C.W.; Mankin, J.S. Carbon majors and the scientific case for climate liability. Nature 2025, 640, 893–901. [Google Scholar] [CrossRef] [PubMed]
  85. Hatab, A.A.; Bustamante, M.; Clark, H.; Havlík, P.; House, J.; Mbow, C.; Ninan, K.N.; Popp, A.; Roe, S.; Sohngen, B.; et al. Agriculture, Forestry and Other Land Uses 1 (AFOLU), in AR6 Synthesis Report. Climate Change 2023; Intergovernmental Panel on Climate Change (IPCC): Geneva, Switzerland, 2023. [Google Scholar]
  86. Secretariat of the Convention on Wetlands. Restoring Drained Peatlands: A Necessary Step to Achieve Global Climate Goals; Ramsar Policy Brief; The Convention on Wetlands (Ramsar Convention): Gland, Switzerland, 2021. [Google Scholar]
  87. Chandra, D.; Chatterton, P.; Chico, G.; Duhigg, G.; Gibson, K.; Hart, S.; Hickson, E.; Hill, M.; Kerkvliet-Hermans, R.; Kopansky, D. Investing in Peatlands; Climate Catalyst: London, UK; Landscape Finance Lab: Wien, Austria; Water Land: Amsterdam, The Netherlands, 2024. [Google Scholar]
  88. Dargie, G.C.; Lewis, S.L.; Lawson, I.T.; Mitchard, E.T.A.; Page, S.E.; Bocko, Y.E.; Ifo, S.A. Age, extent and carbon storage of the central Congo Basin peatland complex. Nature 2017, 542, 86–90. [Google Scholar] [CrossRef] [PubMed]
  89. FAO. The State of the World’s Forests 2022. In Forest Pathways for Green Recovery and Building Inclusive, Resilient and Sustainable Economies; Food and Agriculture Organization of the United Nations (FAO): Italy, Rome, 2022. [Google Scholar] [CrossRef]
  90. Streiff, L. NASA Satellites Help Quantify Forests’ Impacts on Global Carbon Budget; National Aeronautics and Space Administration (NASA): Washington, DC, USA, 2021. Available online: https://www.nasa.gov/science-research/earth-science/nasa-satellites-help-quantify-forests-impacts-on-global-carbon-budget/ (accessed on 1 April 2025).
  91. UNDP. Forests Can Help Us Limit Climate Change—Here Is How; United Nations Development Programme (UNDP): New York, NY, USA, 2023. [Google Scholar]
  92. Nancy, L.; Harris, N.L.; Gibbs, D.A.; Baccini, A.; Birdsey, R.A.; de Bruin, S.; Farina, M.; Fatoyinbo, L.; Hansen, M.C.; Herold, M.; et al. Global Maps Twenty-First Century For. Carbon fluxes. Nat. Clim. Change 2021, 11, 234–240. [Google Scholar] [CrossRef]
  93. UN-ECA. Mapping Carbon and Forest Certification, and Related Economic Benefits to Investors in the Transition to Net Zero: Examples from the Countries Participating in the Congo Basin Blue Fund; United Nations, Economic Commission for Africa: Addis Ababa, Ethiopia, 2024; Available online: https://hdl.handle.net/10855/50096 (accessed on 1 April 2025).
  94. Gachet, B.; Naranjo, L.G.; Román, L.; Surkin, J.; Tamayo, V. Living Amazon Report 2022; World Wildlife Fund for Nature (WWF): Quito, Ecuador, 2022. [Google Scholar]
  95. Skrzypczak, D.; Gorazda, K.; Mikula, K.; Mironiuk, M.; Kominko, H.; Sawska, K.; Evrard, D.; Trzaska, K.; Moustakas, K.; Chojnack, K. Towards carbon neutrality: Enhancing CO2 sequestration by plants to reduce carbon footprint. Sci. Total Environ. 2025, 966, 178763. [Google Scholar] [CrossRef]
  96. Ahmed, A.A.Q.; Odelade, K.A.; Babalola, O.O. Microbial Inoculants for Improving Carbon Sequestration in Agroecosystems to Mitigate Climate Change. In Handbook of Climate Change Resilience; Leal Filho, W., Ed.; Springer: Berlin/Heidelberg, Germany, 2019; pp. 1–20. [Google Scholar]
  97. Hannula, S.E.; Morriën, E. Will fungi solve the carbon dilemma? Geoderma 2022, 413, 115767. [Google Scholar] [CrossRef]
  98. UN-Habitat. World Cities Report 2022; United Nations Human Settlements Programme (UN-Habitat): Nairobi, Kenya, 2022; Available online: https://unhabitat.org/sites/default/files/2022/06/wcr_2022.pdf (accessed on 20 April 2025).
  99. World Bank. Urban Development; World Bank: Washington, DC, USA, 2023; Available online: https://www.worldbank.org/en/topic/urbandevelopment/overview (accessed on 20 April 2025).
  100. Lindsey, R. Climate Change: Global Sea Level. In Science and Information for a Climate-Smart Nation; National Oceanic and Atmospheric Administration NOAA: Washington, DC, USA, 2022. Available online: https://www.climate.gov/news-features/understanding-climate/climate-change-global-sea-level (accessed on 13 April 2025).
  101. Laffoley, D.; Baxter, J.M. (Eds.) Ocean Deoxygenation: Everyone’s Problem—Causes, Impacts, Consequences and Solutions; Full report; IUCN: Gland, Switzerland, 2019. [Google Scholar]
  102. Evison, W.; Low, L.P.; O’Brie, D. Managing Nature Risks: From Understanding to Action; PricewaterhouseCoopers (PwC): Aberdeen, UK, 2023. [Google Scholar]
  103. Whieldon, E.; Yap, S.; Raikwar, L.; Desme, G. How the World’s Largest Companies Depend on Nature and Biodiversity; S&P Global Sustainable: New York, NY, USA, 2023. [Google Scholar]
  104. Hickel, J. High-Income Countries Are Responsible for 74 Percent of Excess Resource Use Causing Ecological Breakdown; London School of Economics and Political Science: London, UK, 2022. [Google Scholar]
  105. Circular Economy Foundation. Circularity Gap Report 2020; Circular Economy Foundation: Brussels, Belgium, 2020. [Google Scholar]
  106. OECD. Global Material Resources Outlook to 2060: Economic Drivers and Environmental Consequences; OECD Publishing: Paris, France, 2019. [Google Scholar] [CrossRef]
  107. Dsilva, J.; Zarmukhambetova, S.; Locke, J. Assessment of building materials in the construction sector: A case study using life cycle assessment approach to achieve the circular economy. Heliyon 2023, 9, e20404. [Google Scholar] [CrossRef]
  108. Sameer, H.; Bringezu, S. Life cycle input indicators of material resource use for enhancing sustainability assessment schemes of buildings. J. Build. Eng. 2019, 21, 230–242. [Google Scholar] [CrossRef]
  109. Hunziker, R.; Schweizer, C. Transforming the Built Environment; World Business Council for Sustainable Development: Geneva, Switzerland, 2025; Available online: https://www.wbcsd.org/actions/transforming-the-built-environment/ (accessed on 15 April 2025).
  110. World Green Building Council. Bringing Embodied Carbon Upfront; World Green Building Council: London, UK, 2025; Available online: https://worldgbc.org/climate-action/embodied-carbon/ (accessed on 2 April 2025).
  111. Rondinel-Oviedo, D.R.; Naomi Keena, N. Embodied Carbon: A call to the building industry 2022. IOP Conf. Ser. Earth Environ. Sci. 2022, 1122, 012042. [Google Scholar] [CrossRef]
  112. Reichl, C.; Schatz, M. World Mining Data 2022; Federal Ministry of Agriculture, Regions and Tourism: Vienna, Austria, 2022; Volume 37.
  113. Potapov, P.; Turubanova, S.; Hansen, M.C.; Tyukavina, A.; Zalles, V.; Khan, A.; Song, X.P.; Pickens, A.; Shen, Q.; Cortez, J. Global maps of cropland extent and change show accelerated cropland expansion in the twenty-first century. Nat. Food 2022, 3, 19–28. [Google Scholar]
  114. Global Footprint Network. Earth Overshoot Day 2024 Approaching; Global Footprint Network: Oakland, CA, USA, 2024. Available online: https://www.footprintnetwork.org/2024/07/21/earth_overshoot_day_2024/ (accessed on 3 April 2025).
  115. The World Counts. Tonnes of CO2 Emitted into the Atmosphere. The World Count. 2025. Available online: https://www.theworldcounts.com/challenges/global-warming/global-co2-emissions (accessed on 3 April 2025).
  116. Hansen, J.; Kharecha, P.; Sato, M.; Masson-Delmotte, V.; Ackerman, F.; Beerling, D.J.; Hearty, P.J.; Hoegh-Guldberg, O.; Hsu, S.-L.; Parmesan, C. Assessing ‘‘Dangerous Climate Change’’: Required Reduction of Carbon Emissions to Protect Young People, Future Generations and Nature. PLoS ONE 2013, 8, e81648. [Google Scholar] [CrossRef]
  117. Global Carbon Budget. Fossil Fuel CO2 Emissions Increase Again in 2024; Global Carbon Project; University of Exeter’s Global Systems Institute: Exeter, UK, 2024; Available online: https://globalcarbonbudget.org/fossil-fuel-co2-emissions-increase-again-in-2024/ (accessed on 16 April 2025).
  118. Tiseo, I. Greenhouse Gas Emissions Worldwide: Statistics and Facts: Statista. 2025. Available online: https://www.statista.com/topics/5770/global-greenhouse-gas-emissions/#topicOverview (accessed on 16 April 2025).
  119. World Bank. Making Mining Forest-Smart Executive Summary Report of the World Bank; World Bank: Washington, DC, USA, 2019. [Google Scholar]
  120. Young, R.E.; Gann, G.D.; Walder, B.; Liu, J.G.; Cui, W.H.; Newton, V.; Nelson, C.R.; Tashe, N.; Jasper, D.; Silveira, F.A.O.; et al. International principles and standards for the ecological restoration and recovery of mine sites. Restor. Ecol. 2022, 30, e13771. [Google Scholar] [CrossRef]
  121. Kramer, M.; Kind-Rieper, T.; Munayer, R.; Giljum, S.; Masselink, R.; Ackern, P.-V.; Maus, V.; Sebastian Luckeneder, S.; Kuschnig, N.; Felipe Costa, F.; et al. Extracted Forests: Unearthing the Role of Mining-Related Deforestation as a Driver of Global Deforestation; WWF Deutschland: Berlin, Germany, 2023. [Google Scholar]
  122. Haddad, N.M.; Brudvig, L.A.; Clobert, J.; Davies, K.F.; Gonzalez, A.; Holt, R.D.; Lovejoy, T.; Sexton, J.O.; Austin, M.P.; Collins, C.D. Habitat fragmentation and its lasting impact on Earth’s ecosystems. Sci. Adv. 2015, 1, e1500052. [Google Scholar] [CrossRef]
  123. Sun, M.; Li, W.; Zhu, L.; Guo, Z.X.; Zhao, Z.; Meng, N.; Han, M.J.; Wang, N.; Zhang, X.M. Degradation in edge forests caused by forest fragmentation. Carbon. Res. 2025, 4, 38. [Google Scholar] [CrossRef]
  124. Iqbal, P.M. The Threat of Coal Mine Expansion Towards Biodiversity in Kalimantan; Indonesia Perkumpulan Aksi Ekologi dan Emansipasi Rakyat (AEER): Jakarta, Indonesia, 2022. [Google Scholar]
  125. Sonter, L.J.; Herrera, D.; Barrett, D.J.; Galford1, G.L.; Moran, C.J.; Soares-Filho, B.S. Mining drives extensive deforestation in the Brazilian Amazon. Nat. Commun. 2017, 8, 1013. [Google Scholar] [CrossRef] [PubMed]
  126. Zeng, Z.H.; Estes, L.; Ziegler, A.D.; Chen, A.P.; Searchinger, T.; Hua, F.Y.; Guan, K.Y.; Jintrawet, A.; Wood, E.F. Highland cropland expansion and forest loss in Southeast Asia in the twenty-first century. Nat. Geosci. 2018, 11, 556–562. [Google Scholar] [CrossRef]
  127. Williams, D.R.; Clark, M.; Buchanan, G.M.; Francesco, G.; Rondinini, C.; Tilman, D. Proactive conservation to prevent habitat losses to agricultural expansion. Nat. Sustain. 2021, 4, 314–322. [Google Scholar] [CrossRef]
  128. Byers, A.K.; Wakelin, S.A.; Condron, L.; Black, A. Land Use Change Disrupts the Network Complexity and Stability of Soil Microbial Carbon Cycling Genes Across an Agricultural Mosaic Landscape. Microb. Ecol. 2024, 87, 167. [Google Scholar] [CrossRef]
  129. Vilhar, U.; Kermavnar, J.; Kozamernik, E.; Petrič, M.; Ravbar, N. The effects of large-scale forest disturbances on hydrology—An overview with special emphasis on karst aquifer systems. Earth Sci. Rev. 2022, 235, 104243. [Google Scholar] [CrossRef]
  130. Hoegh-Guldberg, O.; Cai, R.; Poloczanska, E.S.; Brewer, P.G.; Sundby, S.; Hilmi, K.; Fabry, V.J.; Jung, S. The Ocean. In Climate Change 2014: Impacts, Adaptation, and Vulnerability, in, Part B: Regional Aspects. Contribution of Working Group II to the Fifth Assessment Report of the Intergovernmental Panel on Climate Change; Barros, V.R., Field, C.B., Dokken, D.J., Mastrandrea, M.D., Mach, K.J., Bilir, T.E., Chatterjee, M., Ebi, K.L., Estrada, Y.O., Genova, R.C., et al., Eds.; Cambridge University Press: Cambridge, UK; New York, NY, USA, 2014; pp. 1655–1731. [Google Scholar]
  131. NOAA. Climate at a Glance: Global Time Series. National Centers for Environmental Information; National Oceanic and Atmospheric Administration (NOAA): Silver Spring, MD, USA, 2025. Available online: https://www.ncei.noaa.gov/access/monitoring/climate-at-a-glance/global/time-series (accessed on 20 April 2025).
  132. The GlaMBIE Team. Community estimate of global glacier mass changes from 2000 to 2023. Nature 2025, 639, 382–388. [Google Scholar] [CrossRef] [PubMed]
  133. NASA. Antarctic Ice Mass Loss 2002–2023; National Aeronautics and Space Administration (NASA): Washington, DC, USA, 2023. Available online: https://science.nasa.gov/resource/video-antarctic-ice-mass-loss-2002-2023/ (accessed on 13 April 2025).
  134. Schuur, E.A.G.; Abbott, B.W.; Commane, R.; Ernakovich, J.; Euskirchen, E.; Hugelius, G.; Grosse, G.; Jones, M.; Koven, C.; Leshyk, V.; et al. Permafrost and Climate Change: Carbon Cycle Feedbacks from the Warming Arctic. Annu. Rev. Environ. Resour. 2022, 47, 343–371. [Google Scholar] [CrossRef]
  135. In ‘t Zandt, M.H.; Liebner, S.; Welte, C.U. Roles of Thermokarst Lakes in a Warming World. Trends Microbiol. 2020, 28, 769–779. [Google Scholar] [CrossRef]
  136. Hugelius, G.; Ramage, J.; Burke, E.; Chatterjee, A.; Smallman, T.L.; Aalto, T.; Bastos, A.; Biasi, C.; Canadell, J.G.; Chandra, N.; et al. Permafrost Region Greenhouse Gas Budgets Suggest a Weak CO2 Sink and CH4 and N2O Sources, But Magnitudes Differ Between Top-Down and Bottom-Up Methods. Global Biogeochem. Cycles 2024, 38, e2023GB007969. [Google Scholar] [CrossRef]
  137. Boers, N. Observation-based early-warning signals for a collapse of the Atlantic Meridional Overturning Circulation. Nat. Clim. Change 2021, 11, 680–688. [Google Scholar] [CrossRef]
  138. Sweet, W.V.; Horton, R.; Kopp, R.E.; LeGrande, A.N.; Romanou, A. Sea Level Rise. In Climate Science Special Report: Fourth National Climate Assessment; Wuebbles, D.J., Fahey, D.W., Hibbard, K.A., Dokken, D.J., Stewart, B.C., Maycock, T.K., Eds.; U.S. Global Change Research Program: Washington, DC, USA, 2017; Volume I, pp. 333–363. [Google Scholar] [CrossRef]
  139. Robinson, A.; Lehmann, J.; Barriopedro, D.; Rahmstorf, S.; Coumou, D. Increasing heat and rainfall extremes now far outside the historical climate. npj Clim. Atmos. Sci. 2021, 4, 45. [Google Scholar] [CrossRef]
  140. NASA Science Editorial Team. Steamy Relationships: How Atmospheric Water Vapor Amplifies Earth’s Greenhouse Effect; National Aeronautics and Space Administration (NASA): Washington, DC, USA. Available online: https://science.nasa.gov/earth/climate-change/steamy-relationships-how-atmospheric-water-vapor-amplifies-earths-greenhouse-effect/ (accessed on 20 April 2025).
  141. Lee, W.Y.; Gim, H.J.; Park, S.K. Parameterizations of Snow Cover, Snow Albedo and Snow Density in Land Surface Models: A Comparative Review. Asia-Pac. J. Atmos. Sci. 2024, 60, 185–210. [Google Scholar] [CrossRef]
  142. Wufu, A.; Chen, Y.; Yang, S.; Lou, H.; Wang, P.; Li, C.; Wang, J.; Ma, L. Changes in Glacial Meltwater Runoff and its Response to Climate Change in the Tianshan Region Detected Using Unmanned Aerial Vehicles (UAVs) and Satellite Remote Sensing. Water 2021, 13, 1753. [Google Scholar] [CrossRef]
  143. Thornalley, D.J.R.; Oppo, D.W.; Ortega, P.; Robson, J.I.; Brierley, C.M.; Davis, R.; Hall, I.R.; Moffa-Sanchez, P.; Rose, N.L.; Spooner, P.; et al. Anomalously weak Labrador Sea convection and Atlantic overturning during the past 150 years. Nature 2018, 556, 227–230. [Google Scholar] [CrossRef]
  144. Caesar, L.; Rahmstorf, S.; Robinson, A.; Feulner, G.; Saba, V. Observed fingerprint of a weakening Atlantic Ocean overturning circulation. Nature 2018, 556, 7700. [Google Scholar] [CrossRef] [PubMed]
  145. Met Office. Factsheet: The Atlantic Meridional Overturning Circulation (AMOC); Met Office: Devon, UK, 2025.
  146. Lønborg, C.; Carreira, C.; Jickells, T.; Álvarez-Salgado, X.A. Impacts of Global Change on Ocean Dissolved Organic Carbon (DOC) Cycling. Front. Mar. Sci. 2020, 7, 466. [Google Scholar] [CrossRef]
  147. Tim DeVries, T.; Le Quéré, C.; Andrews, O.; Séférian, R. Decadal trends in the ocean carbon sink. Proc. Natl. Acad. Sci. USA 2019, 116, 11646–11651. [Google Scholar] [CrossRef]
  148. Armstrong McKay, D.I.; Staal, A.; Abrams, J.F.; Winkelmann, R.; Sakschewski, B.; Loriani, S.; Fetzer, I.; Cornell, S.E.; Rockström, J.; Lenton, T.M. Exceeding 1.5 °C global warming could trigger multiple climate tipping points. Science 2022, 377, 6611. [Google Scholar] [CrossRef]
  149. Wong, T.; Campos, R.; Mackres, E.; Staedicke, S.; Doust, M. What Would Cities Look Like with 3 Degrees C of Warming vs. 1.5? Far More Hazardous and Vastly Unequal; World Resource Institute: Washington, DC, USA, 2024. [Google Scholar]
  150. Spracklen, D.V.; Baker, J.C.A.; Garcia-Carreras, L.; Marsham, J.H. The Effects of Tropical Vegetation on Rainfall. Annu. Rev. Environ. Resour. 2018, 43, 193–218. [Google Scholar] [CrossRef]
  151. Fu, R.; Lei, Y.N.; Lib, W.H.; Ariasc, P.A.; Dickinsona, R.E.; Lei, H.G.; Chakrabortya, S.; Fernandesd, K.; Liebmanne, B.; Fisherf, R.; et al. Increased dry-season length over southern Amazonia in recent decades and its implication for future climate projection. Proc. Natl. Acad. Sci. USA 2013, 110, 18110–18115. [Google Scholar] [CrossRef]
  152. Brienen, R.; Phillips, O.; Feldpausch, T.; Gloor, E.; Baker, T.R.; Lloyd, L.; Lopez-Gonzalez, G.; Monteagudo-Mendoza, A.; Malhi, Y.; Lewis, S.L.; et al. Long-term decline of the Amazon carbon sink. Nature 2015, 519, 344–348. [Google Scholar] [CrossRef]
  153. Try, S.; Qin, X.S. Evaluation of Future Changes in Climate Extremes over Southeast Asia Using Downscaled CMIP6 GCM Projections. Water 2024, 16, 2207. [Google Scholar] [CrossRef]
  154. Liang, T.-Y.; Hsiao, S.-C.; Wu, H.-L.; Chen, H.-G.; Chen, W.-B. Accelerated Ocean thermal expansion and its contribution to Global Sea-level rise. J. Sea Res. 2025, 205, 102589. [Google Scholar] [CrossRef]
  155. Böning, C.; Behrens, E.; Biastoch, A.; Getzlaff, K.; Bamber, J.L. Emerging impact of Greenland meltwater on deepwater formation in the North Atlantic Ocean. Nat. Geosci. 2016, 9, 523–527. [Google Scholar] [CrossRef]
  156. Brennan, P. Greenland Melt Speeds East Coast Sea Level Rise; NASA Science Editorial Team; National Aeronautics and Space Administration (NASA): Washington, DC, USA, 2024. Available online: https://science.nasa.gov/earth/climate-change/greenland-melt-speeds-east-coast-sea-level-rise/ (accessed on 20 June 2025).
  157. Kanzow, T.; Humbert, A.; Mölg, T.; Scheinert, M.; Braun, M.; Burchard, H.; Doglioni, F.; Hochreuther, P.; Horwath, M.; Huhn, O.; et al. The system of atmosphere, land, ice and ocean in the region near the 79N Glacier in northeast Greenland: Synthesis and key findings from the Greenland Ice Sheet–Ocean Interaction (GROCE) experiment. Cryosphere 2025, 19, 1789–1824. [Google Scholar] [CrossRef]
  158. Prakash, A.; Zhou, Q.; Hattermann, T.; Kirchner, N. Enhanced subglacial discharge amplifies Petermann Ice Shelf melting when ocean thermal forcing saturates. Nat. Commun. 2025, 16, 4213. [Google Scholar] [CrossRef] [PubMed]
  159. Yu, J.; Anderson, R.F.; Jin, Z.D.; Ji, X.; Thornalley, D.J.R.; Wu, L.; Thouveny, N.; Cai, Y.; Tan, L.; Zhang, F.; et al. Millennial atmospheric CO2 changes linked to ocean ventilation modes over past 150,000 years. Nat. Geosci. 2023, 16, 1166–1173. [Google Scholar] [CrossRef]
  160. Kashiwase, H.; Ohshima, K.I.; Nihashi, S.; Eicken, H. Evidence for ice-ocean albedo feedback in the Arctic Ocean shifting to a seasonal ice zone. Sci. Rep. 2017, 7, 8170. [Google Scholar] [CrossRef]
  161. Kristina Pistone, K.; Eisenman, I.; Ramanathan, V. Observational determination of albedo decrease caused by vanishing Arctic sea ice. Proc. Natl. Acad. Sci. USA 2014, 111, 3322–3326. [Google Scholar] [CrossRef] [PubMed]
  162. WMO. State of the Global Climate 2024; World Meteorological Organization: Geneva, Switzerland, 2025. [Google Scholar]
  163. Carrington, D. More than 150 ‘Unprecedented’ Climate Disasters Struck World in 2024, Says UN. The Guardian. 2025. Available online: https://amp.theguardian.com/environment/2025/mar/19/unprecedented-climate-disasters-extreme-weather-un-report (accessed on 22 April 2025).
  164. WWA. Climate Change Supercharged Late Typhoon Season in the Philippines, Highlighting the Need for Resilience to Consecutive Events; World Weather Attribution (WWA), Imperial College London and the Met Office: London, UK, 2024; Available online: https://www.worldweatherattribution.org/climate-change-supercharged-late-typhoon-season-in-the-philippines-highlighting-the-need-for-resilience-to-consecutive-events/ (accessed on 18 April 2025).
  165. CTIF. Record-Breaking Heatwaves Affect South America in 2025—New Large Wildfires in Chile and Argentina; International Association of Fire and Rescue Services: Ljubljana, Slovenia, 2025; Available online: https://ctif.org/news/record-breaking-heatwaves-affect-south-america-2025-new-large-wildfires-chile-and-argentina (accessed on 20 April 2025).
  166. Brunet, J.; Fragoso, F.P. What are the main reasons for the worldwide decline in pollinator populations? CABI Rev. 2024, 19, 1. [Google Scholar] [CrossRef]
  167. United Nations. Vietnam Multi-sector Assessment (VMSA) Report for Typhoon Yagi Recovery; United Nations: Ha Noi, Vietnam, 2024. [Google Scholar]
  168. Euronews. Man Swept Away in Floodwater as Heavy Rainstorms Batter Parts of Northern Italy. Euronews. 2024. Available online: https://www.euronews.com/my-europe/2024/09/06/man-swept-away-in-floodwater-as-heavy-rainstorms-batter-parts-of-northern-italy (accessed on 18 April 2025).
  169. Li, H.; Zhao, Y.; Wang, C.; Ürge-Vorsatz, D.; Carmeliet, J.; Bardhan, R. Cooling efficacy of trees across cities is determined by background climate, urban morphology, and tree trait. Commun. Earth Environ. 2024, 5, 754. [Google Scholar] [CrossRef]
  170. Iungman, T.; Cirach, M.; Marando, F.; Barboza, E.P.; Khomenko, S.; Masselot, P.; Quijal-Zamorano, M.; Mueller, N.; Gasparrini, A.; Urquiza, J.; et al. Cooling cities through urban green infrastructure: A health impact assessment of European cities. Lancet 2023, 401 (Suppl. S1), 577–589. [Google Scholar] [CrossRef] [PubMed]
  171. Marando, F.; Heris, M.P.; Zulian, G.; Udías, A.; Mentaschi, L.; Chrysoulakis, N.; Parastatidis, D.; Maes, J. Urban heat island mitigation by green infrastructure in European functional urban areas. Sustain. Cities Soc. 2022, 77, 103564. [Google Scholar] [CrossRef]
  172. Lwasa, S.; Seto, K.C.; Bai, X.; Blanco, H.; Gurney, K.R.; Kılkış, S.; Lucon, O.; Murakami, J.; Pan, J.; Sharifi, A.; et al. Urban systems and other settlements. In IPCC, 2022: Climate Change 2022: Mitigation of Climate Change. Contribution of Working Group III to the Sixth Assessment Report of the Intergovernmental Panel on Climate Change; Intergovernmental Panel on Climate Change (IPCC): Geneva, Switzerland, 2022; pp. 861–952. [Google Scholar]
  173. Sarip, S. 100 Million Tree Planting Campaign Program Organized by JPNP and UTM—The Power of Tree Planting as an Effective Carbon Sarip, S. offset Strategy; Universiti Teknologi Malaysia: Johor Darul Ta’zim, Malaysia, 2023; Available online: https://news.utm.my/2023/08/program-100-million-tree-planting-campaign-organized-by-jpnp-and-utm-the-power-of-tree-planting-as-an-effective-carbon-offset-strategy/ (accessed on 22 April 2025).
  174. Gimino, G.; Rahim, R.; Vethasalam, R. M’sia Hits Target of Planting 100 Million Trees Ahead of Schedule. The Star. 2024. Available online: https://www.thestar.com.my/news/nation/2024/12/10/m039sia-hits-target-of-planting-100-million-trees-ahead-of-schedule#:~:text=KUALA%20LUMPUR:%20Malaysia%20has%20achieved,on%20the%20Parliament%20complex%20grounds (accessed on 22 April 2025).
  175. Xinhua. China’s Sustained Afforestation Efforts Green the World; The State Council Information Office of the People’s Republic of China: Beijing, China, 2023. Available online: http://english.scio.gov.cn/chinavoices/2023-04/06/content_85213960.htm (accessed on 20 April 2025).
  176. Cheng, K.; Yang, H.; Tao, S.; Su, Y.; Guan, H.; Ren, Y.; Hu, T.; Li, W.; Xu, G.; Chen, M.; et al. Carbon storage through China’s planted forest expansion. Nat. Commun. 2024, 15, 4106. [Google Scholar] [CrossRef] [PubMed] [PubMed Central]
  177. Graham, D. How Many Trees Are There in China? Scientists Just Counted Them and Here’s What They Found? BBC Wildlife: London, UK, 2025; Available online: https://www.discoverwildlife.com/plant-facts/trees/how-many-trees-are-there-in-china (accessed on 20 April 2025).
  178. Chew, H.M. How Urbanized Singapore Is Learning to Live with Its Wildlife. CAN. 2020. Available online: https://www.channelnewsasia.com/singapore/singapore-wildlife-conservation-city-in-nature-reserves-hornbill-1339231 (accessed on 20 April 2025).
  179. Tan, M.; Ng, D. Sungai Buloh Wetland Reserve Celebrates 30 Years with Activities and Guided Walks to Previously Inaccessible areas. CAN. 2023. Available online: https://www.channelnewsasia.com/singapore/sungei-buloh-wetland-reserve-celebrates-30-years-activities-and-guided-walks-previously-inaccessible-areas-3948276 (accessed on 20 April 2025).
  180. Brenneisen, S.; Baumann, N. Green Roofs in Basel, Switzerland: Combining Mitigation and Adaptation Measures; Climate Adapt; European Environmental Agency: Copenhagen, Denmark, 2020; Available online: https://climate-adapt.eea.europa.eu/en/metadata/case-studies/green-roofs-in-basel-switzerland-combining-mitigation-and-adaptation-measures-1 (accessed on 20 April 2025).
  181. Fang, Y.; Du, X.T.; Zhao, H.Y.; Hu, M.; Xu, X.M. Assessment of green roofs’ potential to improve the urban thermal environment: The case of Beijing. Environ. Res. 2023, 237 Pt 1, 116857. [Google Scholar] [CrossRef]
  182. Amani, M.J.; Tanzadeh, R.; Moghadas Nejad, F.; Kabiri Nasrabad, M.M.; Chalabii, J.; Movahedi Rad, M. Urban Sustainability Through Pavement Technologies: Reducing Urban Heat Islands with Cool Pavements. Buildings 2025, 15, 504. [Google Scholar] [CrossRef]
  183. García-Melgar, P.; Delgado, M.G.; Romero García, C.; Ramos, J.S.; Domínguez, S.A. Nature-based cool pavements for urban overheating mitigation: Experimental proof of concept. Build. Environ. 2025, 267 Pt B, 112184. [Google Scholar] [CrossRef]
  184. Kousis, I.; Pisello, A.L. Evaluating the performance of cool pavements for urban heat island mitigation under realistic conditions: A systematic review and meta-analysis. Urban. Clim. 2023, 49, 101470. [Google Scholar] [CrossRef]
  185. Buo, I.; Sagris, V.; Jaagus, J. High-resolution thermal exposure and shade maps for cool corridor planning. Sustain. Cities Soc. 2023, 93, 104499. [Google Scholar] [CrossRef]
  186. Ariane Middel Cilek, A.; Unal, M.; Middel, A. The effects of 2-D and 3-D urban landscape metrics on mean radiant temperature in hot-arid Phoenix and Tempe, Arizona, USA. Sustain. Cities Soc. 2024, 101, 105116. [Google Scholar] [CrossRef]
  187. The Sankei Shimbun. Tokyo Introduces Cool Pavements to Combat Heat Island Effect; The Sankei Shimbun: Tokyo, Japan, 2024; Available online: https://featured.japan-forward.com/japan2earth/2024/07/7818/ (accessed on 22 April 2025).
  188. Yan, S.; Zhang, T.; Wu, Y.; Lv, C.; Qi, F.; Chen, Y.; Wu, X.; Shen, Y. Cooling Effect of Trees with Different Attributes and Layouts on the Surface Heat Island of Urban Street Canyons in Summer. Atmosphere 2023, 14, 857. [Google Scholar] [CrossRef]
  189. Ziter, C.D.; Pedersen, E.J.; Kucharik, C.J.; Turner, M.G. Scale-dependent interactions between tree canopy cover and impervious surfaces reduce daytime urban heat during summer. Proc. Natl. Acad. Sci. USA 2019, 116, 7575–7580. [Google Scholar] [CrossRef] [PubMed]
  190. Pimm, S.L.; Willigan, E.; Kolarova, A.; Huang, R. Reconnecting nature. Curr. Biol. 2021, 31, R1141–R1224. [Google Scholar] [CrossRef]
  191. Campanales, S. Protecting the Endangered Wildlife of Borneo’s Last Peatlands. EcoAct. 2023. Available online: https://eco-act.com/blog/protecting-endangered-wildlife/ (accessed on 22 April 2025).
  192. Indriatmoko, Y.; Atmadja, S.S.; Utomo, N.A.; Ekaputri, A.D.; Komalasari, M. Katingan Peatland Restoration and Conservation Project, Central Kalimantan, Indonesia, in, REDD+ on the Ground: A Case Book of Subnational Initiatives Across the Globe; Center for International Forestry Research (CIFOR): Bogor, Indonesia, 2014; pp. 309–328. [Google Scholar]
  193. Keong, C.Y.; Onuma, A. Transboundary Ecological Conservation, Environmental Value, and Environmental Sustainability: Lessons from the Heart of Borneo. Sustainability 2021, 13, 9727. [Google Scholar] [CrossRef]
  194. Mellgard, P. Life After Lignite: How Lusatia Has Returned to Nature. The Guardian. 2014. Available online: https://www.theguardian.com/environment/2014/sep/10/lusatia-lignite-mining-germany-lake-district (accessed on 22 April 2025).
  195. Sosa, I.; Keenan, K. Impact Benefit Agreements Between Aboriginal Communities and Mining Companies Their Use in Canada. Indigenous Rights and Resource Governance Research Group (IRRG). 2001. Available online: https://fpic.info/en/resources/impact-benefit-agreements-between-aboriginal-commu/ (accessed on 22 April 2025).
  196. Meerveld, D. Assessing Value: A Comprehensive Study of Impact Benefit Agreements on Indigenous Communities of Canada; Graduate School of Public and International Affairs, University of Ottawa: Ottawa, ON, Canada, 2016; Available online: http://hdl.handle.net/10393/34816 (accessed on 25 April 2025).
  197. Prno, J.; Bradshaw, B.; Lapierre, D. Impact and Benefit Agreements: Are They Working? Indigenous People Study Portal, University of Saskatchewan: Saskatoon, SK, Canada, 2010; Available online: https://iportal.usask.ca/record/28190 (accessed on 22 April 2025).
  198. van der Heyde, M.; Bunce, M.; Dixon, K.; Wardell-Johnson, G.; White, N.E.; Nevill, P. Changes in soil microbial communities in post-mine ecological restoration: Implications for monitoring using high throughput DNA sequencing. Sci. Total Environ. 2020, 749, 142262. [Google Scholar] [CrossRef] [PubMed]
  199. Khan, S.; Ahmad, I.; Tahir Shah, M.; Rehman, S.; Khaliq, A. Use of constructed wetland for the removal of heavy metals from industrial wastewater. J. Environ. Manag. 2009, 90, 3451–3457. [Google Scholar] [CrossRef] [PubMed]
  200. Madarász, B.; Jakab, G.; Szalai, Z.; Juhos, K.; Kotroczó, Z.; Tóth, A.; Ladányi, M. Long-term effects of conservation tillage on soil erosion in Central Europe: A random forest-based approach. Soil. Tillage Res. 2021, 209, 104959. [Google Scholar] [CrossRef]
  201. Yang, X.L.; Xiong, J.R.; Du, T.S.; Ju, X.T.; Gan, Y.T.; Li, S.E.; Xia, L.L.; Shen, Y.J.; Pacenka, S.; Steenhuis, T.S.; et al. Diversifying crop rotation increases food production, reduces net greenhouse gas emissions and improves soil health. Nat. Commun. 2024, 15, 198. [Google Scholar] [CrossRef]
  202. IFOAM Organics Europe. Organic Agriculture and Its Benefits for Climate and Biodiversity; IFOAM Organics Europe: Brussels, Belgium, 2022. [Google Scholar]
  203. Gunawan, H.; Tryanto, D.H.; Mizuno, K.; Kozan, O. Toward Climate Change Mitigation: Restoration of the Indonesian Peat Swamp. In Vulnerability and Transformation of Indonesian Peatlands; Mizuno, K., Kozan, O., Gunawan, H., Eds.; Global Environmental Studies; Springer: Singapore, 2023; pp. 141–157. [Google Scholar] [CrossRef]
  204. Damastuti, E.; de Groot, R.; Debrot, A.O.; Silvius, M.J. Effectiveness of community-based mangrove management for biodiversity conservation: A case study from Central Java, Indonesia. Trees For. People 2022, 7, 100202. [Google Scholar] [CrossRef]
  205. Chan, W.Y.; Meyers, L.; Rudd, D.; Topa, S.H.; van Oppen, M.J.H. Heat-evolved algal symbionts enhance bleaching tolerance of adult corals without trade-off against growth. Glob. Change Biol. 2023, 29, 6945–6968. [Google Scholar] [CrossRef] [PubMed]
  206. Christiansen, M.A.; Ivits-Wasser, V. Floodplains: A Natural System to Preserve and Restore; European Environmental Agency (EEA): Copenhagen, Denmark, 2020. [Google Scholar]
  207. Liu, H.Q.; Evans, S.; Zhang, Z.Z.; Song, W.Y.; You, X.Y. The Carbon Brief Profile: China. Carbon Brief. 2023. Available online: https://interactive.carbonbrief.org/the-carbon-brief-profile-china/index.html (accessed on 23 April 2025).
  208. Žilinskas, G.; Janušaitė, R.; Jarmalavičius, D.; Jukna, L.; Pupienis, D. Possibilities for Mitigating Coastal Erosion in the Downdrift Zone of Port Jetties Using Nearshore Nourishment: A Case Study of Klaipeda Port, Lithuania. J. Mar. Sci. Eng. 2025, 13, 628. [Google Scholar] [CrossRef]
  209. Stanway, D. ‘Revive Our Ocean’ Initiative Launched to Protect Coastal Waters. Reuters. 2025. Available online: https://www.reuters.com/sustainability/climate-energy/revive-our-ocean-initiative-launched-protect-coastal-waters-2025-04-24/ (accessed on 23 April 2025).
  210. Jaynes, C.H. Revive Our Ocean Initiative Aims to Increase Marine Protected Areas to Cover 30% of World’s Oceans by 2030. EcoWatch. 2025. Available online: https://www.ecowatch.com/revive-our-ocean-david-attenborough.html (accessed on 23 April 2025).
  211. IUCN. Peatlands and Climate Change; Issue Brief; International Union for Conservation of Nature (IUCN): Gland, Switzerland, 2021. [Google Scholar]
  212. FAO. FAO Launches Inaugural Field Project in Congo Basin to Advance Sustainable Peatland Management for Climate Action; Food and Agriculture of the United Nations (FAO): Rome, Italy, 2024; Available online: https://www.fao.org/forestry/newsroom/news-detail/fao-launches-inaugural-field-project-in-congo-basin-to-advance-sustainable-peatland-management-for-climate-action/en (accessed on 23 April 2025).
  213. Kiely, L.; Spracklen, D.V.; Arnold, S.R.; Papargyropoulou, E.; Conibear, L.; Wiedinmyer, C.; Knote, C.; Adrianto, H.A. Assessing costs of Indonesian fires and the benefits of restoring peatland. Nat. Commun. 2021, 12, 7044. [Google Scholar] [CrossRef]
  214. Mo, L.; Zohner, C.M.; Reich, P.B.; Liang, J.J.; de Miguel, S.; Nabuure, C.-J.; Renner, S.S.; van den Hoogen, J.; Araza, A.; Herold, M.; et al. Integrated global assessment of the natural forest carbon potential. Nature 2023, 624, 92–101. [Google Scholar] [CrossRef]
  215. Heinrich, V.H.A.; Vancutsem, C.; Dalagno, R.; Rosan, T.M.; Fawcett, D.; Silva-Junior, C.H.L.; Cassol, H.L.G.; Achard, F.; Jucker, T.; Silva, C.A. The carbon sink of secondary and degraded humid tropical forests. Nature 2023, 615, 436–442. [Google Scholar] [CrossRef]
  216. Bokhorst, S.; Bjerke, J.W.; Phoenix, G.K.; Jaakola, L.; Mæhre, H.K.; Tømmervik, H. Sub-arctic mosses and lichens show idiosyncratic responses to combinations of winter heatwaves, freezing and nitrogen deposition. Physiol. Plant. 2023, 175, e13882. [Google Scholar] [CrossRef]
  217. Schuuring, S.; Halvorsen, R.; Eidesen, P.B.; Niittynen, P.; Kemppinen, J.; Lang, S.I. High Arctic Vegetation Communities with a Thick Moss Layer Slow Active Layer Thaw. High Arctic vegetation communities with a thick moss layer slow active layer thaw. J. Geophys. Res. Biogeosci. 2024, 129, e2023JG007880. [Google Scholar] [CrossRef]
  218. Holmgren, M.; Groten, F.; Carracedo, M.R.; Vink, S.; Limpens, J. Rewilding Risks for Peatland Permafrost. Ecosystems 2023, 26, 1806–1818. [Google Scholar] [CrossRef]
  219. Beer, C.; Zimov, N.; Olofsson, J.; Porada, P.; Zimov, S. Protection of Permafrost Soils from Thawing by Increasing Herbivore Density. Sci. Rep. 2020, 10, 4170. [Google Scholar] [CrossRef] [PubMed]
  220. Robertson, D. Inside Copenhagen’s Race to Be the First Carbon-Neutral City. The Guardian. 2019. Available online: https://www.theguardian.com/cities/2019/oct/11/inside-copenhagens-race-to-be-the-first-carbon-neutral-city (accessed on 23 April 2025).
  221. Skowron, A.; van Rie, R.; Martinez, A. Policy Roadmap for 100% Renewable Energy for Costa Rica: Achieving a Fully Decarbonized Energy System; World Future Council: Hamburg, Germany, 2020. [Google Scholar]
  222. Poulsen, H. Taking Action to Stay Within 1.5 °C; Ørsted: Fredericia, Denmark, 2019; Available online: https://orsted.com/en/what-we-do/insights/white-papers/taking-action (accessed on 24 April 2025).
  223. Ørsted. Ørsted Annual Report 2024; Ørsted: Fredericia, Denmark, 2024. [Google Scholar]
  224. Akotia, J.; Manu, E.; Opoku, A.; Sackey, E. The role of built environment practitioners in driving the socio-economic sustainability aspects of sustainable regeneration. Constr. Econ. Build. 2020, 20, 89–108. [Google Scholar] [CrossRef]
Figure 1. Interdependence of core concepts in industrial–urban impact.
Figure 1. Interdependence of core concepts in industrial–urban impact.
Urbansci 09 00320 g001
Figure 4. Thermodynamic paradox of industrial metabolism. Notes: This schematic illustrates the thermodynamic paradox of industrial metabolism: Systems consume finite low-entropy inputs such as fossil fuels, minerals and biomass to produce valuable low-entropy outputs such as goods and machinery. However, these processes inevitably generate high-entropy waste (GHGs, PM2.5, heavy metals, waste heat) that exceeds Earth’s dissipation capacity. The “low-entropy paradox” emerges because even useful products accelerate environmental degradation, for example, urban heat islands from waste heat and CO2 emissions-induced warming effects. This reveals a fundamental tension between economic production and biospheric stability governed by the Second Law of Thermodynamics.
Figure 4. Thermodynamic paradox of industrial metabolism. Notes: This schematic illustrates the thermodynamic paradox of industrial metabolism: Systems consume finite low-entropy inputs such as fossil fuels, minerals and biomass to produce valuable low-entropy outputs such as goods and machinery. However, these processes inevitably generate high-entropy waste (GHGs, PM2.5, heavy metals, waste heat) that exceeds Earth’s dissipation capacity. The “low-entropy paradox” emerges because even useful products accelerate environmental degradation, for example, urban heat islands from waste heat and CO2 emissions-induced warming effects. This reveals a fundamental tension between economic production and biospheric stability governed by the Second Law of Thermodynamics.
Urbansci 09 00320 g004
Figure 5. Urban metabolism and low-entropy paradox. Notes: This diagram illustrates the fundamental low-entropy paradox inherent to industrial–urban metabolism. Systems continuously consume finite low-entropy resources (fossil fuels, minerals, goods/services) to sustain economic functions yet inherently generate destabilizing high-entropy waste streams through thermodynamic processes governed by the Second Law. These waste outputs—including municipal solid waste (plastics), airborne pollutants (CO2, NOX, SOX, PM2.5), and contaminated wastewater—represent dispersed disorder that accumulates beyond Earth’s dissipation capacity. The paradox emerges because even valuable low-entropy outputs (machinery, electronics) ultimately degrade into waste, accelerating global entropy while providing only temporary local order. This irreconcilable tension between resource consumption and waste generation exposes the thermodynamic unsustainability of linear industrial systems within planetary boundaries.
Figure 5. Urban metabolism and low-entropy paradox. Notes: This diagram illustrates the fundamental low-entropy paradox inherent to industrial–urban metabolism. Systems continuously consume finite low-entropy resources (fossil fuels, minerals, goods/services) to sustain economic functions yet inherently generate destabilizing high-entropy waste streams through thermodynamic processes governed by the Second Law. These waste outputs—including municipal solid waste (plastics), airborne pollutants (CO2, NOX, SOX, PM2.5), and contaminated wastewater—represent dispersed disorder that accumulates beyond Earth’s dissipation capacity. The paradox emerges because even valuable low-entropy outputs (machinery, electronics) ultimately degrade into waste, accelerating global entropy while providing only temporary local order. This irreconcilable tension between resource consumption and waste generation exposes the thermodynamic unsustainability of linear industrial systems within planetary boundaries.
Urbansci 09 00320 g005
Figure 6. Vicious Cycle Atlas of Fragility: cascading environmental feedback of industrial–urban metabolism.
Figure 6. Vicious Cycle Atlas of Fragility: cascading environmental feedback of industrial–urban metabolism.
Urbansci 09 00320 g006
Figure 7. The global industrial–urban material economy: resource extraction, consumption, and transformation.
Figure 7. The global industrial–urban material economy: resource extraction, consumption, and transformation.
Urbansci 09 00320 g007
Figure 8. Fossil fuels’ share of world energy mix (oil, natural gas, and coal). Source: [29].
Figure 8. Fossil fuels’ share of world energy mix (oil, natural gas, and coal). Source: [29].
Urbansci 09 00320 g008
Table 1. Planetary boundary stressors by Earth system domain.
Table 1. Planetary boundary stressors by Earth system domain.
DomainKey StressorsData Sources
Atmosphere26 Gt C/yr fossil CO2 (1970–2024)JRC/Crippa et al., 2024 [28]; Energy Institute, 2024 [29].
Atmospheric CO2 concentrations of 424.61 ppm (2024)Lan, X. et al., 2025 [31].
Hydrosphere372 Zetajouls (ZJ) heat (2021; +3000% since 1957)NASA, 2024 [32].
2.8 Ωarag acidificationRichardson et al., 2023 [18].
Cryosphere28 trillion tons ice loss (1994–2017)—a major driver of Arctic amplificationSlater et al., 2021; Harvey, C., 2021 [33,34].
Biosphere61.56 million ha (Mha) deforestation (2001–2023)World Resource Institute/Weisse, M. et al., 2024 [5]; Global Forest Watch, 2025 [6].
>one million species at riskIPBES, 2019 [35].
Lithosphere5.57 Mha impacts (2020)Giljum et al., 2022 [30].
8.2 Mha of land loss from agricultural conversion in Southeast AsiaCabernard et al., 2024 [36].
Table 2. Resource flow translation to systemic threshold based on Table 1.
Table 2. Resource flow translation to systemic threshold based on Table 1.
ScaleResource Flow (Input)Systemic Threshold Breach (Output)QuantificationFeedback Mechanism
GlobalFossil fuel combustionCryosphere destabilization (planetary boundary: ice mass stability)28 trillion tones of ice loss (1994–2024)Ice-albedo feedback → accelerated warming
GlobalMaterial extractionLithosphere degradation (planetary boundary: land system change)5.57 Mha mining footprint (2020)Habitat fragmentation → biodiversity loss
RegionalAgricultural expansionBiosphere degradation (planetary boundary: biosphere integrity)8.2 Mha SE Asian rainforest loss (2000–2020)Reduced evapotranspiration → regional drying
LocalUrban–industrial wasteHydrospheric contamination (planetary boundary: novel entities)Microplastic accumulation: 14 Mt/yrSoil/water toxicity → ecosystem collapse
Note: All solutions are expanded upon in Tables 5 and 6 (Section 8) with case-specific evidence.
Table 3. Quantified stressors and M-NbRS stabilization benefits.
Table 3. Quantified stressors and M-NbRS stabilization benefits.
DomainStressors (Industrial–Urban Impacts)Quantified Stressor EffectStabilizing Solution (M-NbRS)Quantified Benefit
Material FlowsGlobal material extraction100 billion tons/yr (8.6% circularity, 2020–2023 average)Urban circular metabolismPotential to increase circularity beyond 8.6% baseline
AtmosphereFossil CO2 emissions37 billion tons/yr (2024); 424.61 ppm (↑30.38% since 1970)Afforestation (China’s 142 billion trees)873.1 ± 16.2 Tg C stored (1.87 billion tons of CO2)
HydrosphereOcean heat accumulation372 Zettajoules (ZJ) (+3000% since 1957)Peatland restoration98–153 Mt CO2e/yr sequestered (global peatlands)
BiosphereDeforestation (agricultural/mining demand)61.56 million hectares (Mha) lost (2001–2023); 420 Mha to cropland (1990–2020)Heart of Borneo Initiative (transboundary forest ecosystem conservation)
»
24 Mha of forest ecosystems protected
»
Critically endangered species such as Bornean orangutan safeguarded
LithosphereMining soil degradation5.57 Mha of forested land impacted (2020); 36 Mt CO2/yrIndigenous co-management (Canada IBAs)More than 1000 aquatic species revived (Lusatian Lakes)
CryospherePermafrost thaw1400–1700 GtC at riskPermafrost insulation (Svalbard)Conservation of permafrost carbon sink
Table 4. Cascading impacts of anthropogenic land-use change on biospheric stability and climate feedback.
Table 4. Cascading impacts of anthropogenic land-use change on biospheric stability and climate feedback.
Earth System ImpactKey MechanismsQuantitative MetricsGeographic ConcentrationSources
Biosphere collapseHabitat loss and fragmentation
  • 1 million species threatened
Southeast Asia (38%)[35,36]
  • 10% genetic diversity loss (since 1870)
Latin America (36%)
  • 1.5% of global species committed to extinction (since 1995)
Africa (23%)
Climate feedbackReduced carbon sequestration and edge effects
  • 18.7 GtC released from forest fragmentation
Global (tropics predominant)[123]
  • 89.27 billion tons of carbon released (mining 2001–2023)
Biospheric instabilityLithospheric transformation
  • Nutrient cycling impairment
Global (agricultural zones)[128,129]
  • Water regulation loss
  • Pollination service decline
Cross-system couplingDeforestation → atmospheric CO2 release
  • 36 million tons CO2/yr (mining)
Tropical forests[18,31]
  • 30.38% ↑ atmospheric CO2 (1970–2024)
Note: The above table synthesizes peer-reviewed evidence of land-use change impacts across Earth’s systems, quantifying biodiversity loss, climate feedback, and ecosystem service disruptions. Data reflect anthropogenic drivers (mining, agriculture) with geographic hotspots and systemic couplings. All metrics are derived from the cited sources.
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

Keong, C.Y. The Vicious Cycle Atlas of Fragility: Mapping the Feedback Loops Between Industrial–Urban Metabolism and Earth System Collapse. Urban Sci. 2025, 9, 320. https://doi.org/10.3390/urbansci9080320

AMA Style

Keong CY. The Vicious Cycle Atlas of Fragility: Mapping the Feedback Loops Between Industrial–Urban Metabolism and Earth System Collapse. Urban Science. 2025; 9(8):320. https://doi.org/10.3390/urbansci9080320

Chicago/Turabian Style

Keong, Choy Yee. 2025. "The Vicious Cycle Atlas of Fragility: Mapping the Feedback Loops Between Industrial–Urban Metabolism and Earth System Collapse" Urban Science 9, no. 8: 320. https://doi.org/10.3390/urbansci9080320

APA Style

Keong, C. Y. (2025). The Vicious Cycle Atlas of Fragility: Mapping the Feedback Loops Between Industrial–Urban Metabolism and Earth System Collapse. Urban Science, 9(8), 320. https://doi.org/10.3390/urbansci9080320

Article Metrics

Back to TopTop