Next Article in Journal
Women’s Land Ownership and Decision-Making Power in West Sumatra
Previous Article in Journal
Reshaping Urban Innovation Landscapes for Green Growth: The Role of Smart City Policies in Digital Transformation
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

The Tropical Peatlands in Indonesia and Global Environmental Change: A Multi-Dimensional System-Based Analysis and Policy Implications

1
Faculty of Economics, Keio University, Tokyo 108-8345, Japan
2
Institute for Environment and Development (LESTARI), National University of Malaysia, Selangor 43600, Malaysia
*
Author to whom correspondence should be addressed.
Reg. Sci. Environ. Econ. 2025, 2(3), 17; https://doi.org/10.3390/rsee2030017
Submission received: 17 April 2025 / Revised: 19 June 2025 / Accepted: 29 June 2025 / Published: 1 July 2025

Abstract

Tropical peatlands store approximately 105 gigatons of carbon (GtC), serving as vital long-term carbon sinks, yet remain critically underrepresented in climate policy. Indonesia peatlands contain 57GtC—the largest tropical peatland carbon stock in the Asia–Pacific. However, decades of drainage, fires, and lax enforcement practices have degraded vast peatland areas, turning them from carbon sinks into emission sources—as evidenced by the 1997 and 2015 peatland fires which emitted 2.57 Gt CO2eq and 1.75 Gt CO2eq, respectively. Using system theory validated against historical data (1997–2023), we develop a causal loop model revealing three interconnected feedback loops driving irreversible collapse: (1) drainage–desiccation–oxidation, where water table below −40 cm triggers peat oxidation (2–5 cm subsistence) and fires; (2) fire–climate–permafrost, wherein emissions intensify radiative forcing, destabilizing monsoons and accelerating Arctic permafrost thaw (+15% since 2000); and (2) economy–governance failure, perpetuated by palm oil’s economic dominance and slack regulatory oversight. To break these vicious cycles, we propose a precautionary framework featuring IoT-enforced water table (≤40 cm), reducing emissions by 34%, legally protected “Global Climate Stabilization Zones” for peat domes (>3 m depth), safeguarding 57 GtC, and ASEAN transboundary enforcement funded by a 1–3% palm oil levy. Without intervention, annual emissions may reach 2.869 GtCO2e by 2030 (Nationally Determined Contribution’s business-as-usual scenario). Conversely, rewetting 590 km2/year aligns with Indonesia’s FOLU Net Sink 2030 target (−140 Mt CO2e) and mitigates 1.4–1.6 MtCO2 annually. We conclude that integrating peatlands as irreplaceable climate infrastructure into global policy is essential for achieving Paris Agreement goals and SDGs 13–15.

1. Introduction

Tropical peatlands are among the most carbon-dense ecosystems on Earth, storing approximately 105 gigatons of carbon (GtC) [1,2,3]. These ecosystems play a critical role in the global carbon cycle, sequestering carbon over millennia—far exceeding the decadal cycles of forests [4]. Their capacity to act as long-term carbon sinks makes them indispensable for achieving the Paris Agreement’s goal of limiting global warming to below 2 °C and advancing Sustainable Development Goal (SDG) 13 (climate action) [5]. However, human-driven degradation—through agricultural expansion, logging, and drainage—is rapidly converting these vital carbon reservoirs into significant emission sources [6].
Indonesia, hosting the Asia–Pacific’s largest tropical peatlands, epitomizes the global peatland crisis. Over 50% of its 15 million hectares of peat has been degraded, emitting an average of 0.90 gigatonnes of CO2 annually (GtCO2/yr) between 2000 and 2006. Under business-as-usual trajectories, annual emissions are projected to surge to 1.4 GtCO2 by 2025 [7]. Indonesia’s National Determined Contribution (NDC) estimated that emissions may reach 2.869 GtCO2e by 2030 without intervention [8]. These emissions culminate in catastrophic nonlinear events, exemplified by the 2015 peat fires—the most severe since 1997—which released 1.75 GtCO2 through combustion [9] and contributed 71% of Indonesia’s total PM2.5 emissions during September–October 2015 [10]. This pollution exacerbates regional haze crises, biodiversity collapse, and Arctic permafrost destabilization [11,12].
Critically, peatland fires inflict irreversible biodiversity collapse at scales unmatched by illegal logging. Where selective logging gradually fragments habitats (typically disturbing <10% of canopy cover annually), fires trigger near-instantaneous ecosystem obliteration [9]. During the 2002 Mega Rice Project in Central Kalimantan, drainage-enabled fires consumed 27–33 cm of peat [13,14]—erasing over 500 years of carbon accumulation. This instantaneous loss had the following effects:
  • It decimated keystone species: Bornean orangutan populations plummeted >50% in fire zones due to habitat incineration, including vital seed banks necessary for forest regeneration.
  • It induced permanent hydrological collapse: it altered watershed hydrology across 2.7 million hectares, converting peat domes into fire-prone wastelands.
  • It amplified carbon feedback: burned peatlands emit >3× more CO2e/ha/year than degraded forests [15] creating self-reinforcing emission cycles.
The resulting landscape-scale disruption triggers transboundary haze events (affecting 75 million people) [16] and economic losses exceeding USD 16.1 billion [17]—directly contravening SDG 15 (terrestrial ecosystems), SDG 3 (health), and SDG 13 (climate action). This represents an irreversible tipping point: once ignited, peatland fires convert millennia-old carbon sinks into permanent emission sources, accelerating global biogeochemical feedback loops.
Current Dynamic Global Vegetation Models (DGVMs) such as the Lund–Potsdam–Jena General Ecosystem Simulator (LPJ-GUESS) and the Terrestrial Ecosystem Model (TEM) inadequately capture nonlinear peatland-climate feedback due to oversimplified hydrological dynamics [18,19]. For instance, these models fail to represent critical thresholds such as irreversible peat drying below 40 cm water tables, which trigger subsidence (about 5 cm/year) and oxidative carbon loss [20]. Field studies in Indonesia demonstrate that restoring water tables to ≤40 cm reduces CO2 emissions by 20–34% per 10 cm rise [21], highlighting gaps in model parameterization. Perturbed peatlands exhibit nonlinear transitions, (for example, drainage → fire → haze → Arctic feedback loops), creating cascading impacts that destabilize regional hydrology (e.g., 15–20% rainfall decline in Jambi Province) and accelerate global permafrost thaw [15,22].
To address these limitations, this study employs systems theory to evaluate the reciprocal relationships between peatland degradation and global environmental change. Systems theory provides the foundational lens to conceptualize peatlands as complex adaptive systems where ecological, hydrological, and socio-economic components interact through feedback loops across scales (Section 6). This framework recognizes the following [23,24]:
  • Nonlinear Dynamics: Nonlinear dynamics occur when small perturbations trigger disproportionately large impacts due to thresholds, for example, irreversible peat drying at <100% moisture.
  • Feedback Loops: Describe circular cause–effect relationships where outputs of a system reinforce (positive feedback) or counteract (negative feedback) initial changes. For example, drainage lowers peatland water tables, increasing flammability and CO2 emissions, which further desiccate peat soils—a self-reinforcing cycle (see, Section 6).
  • Emergence: System-wide behaviors such as peat collapse arise from close interactions of system components.
  • Multi-scale Coupling: Local actions cascade through subsystems via teleconnections.
In this study, systems theory is employed to analyze how interactions between ecological, socio-economic, and institutional elements drive system behavior, enabling holistic insights into challenges like peatland degradation. This approach captures interdependencies often overlooked in siloed analyses.
The primary objectives of this study are twofold:
  • To develop a systems causality model capturing multi-scale feedback loops between local peatland degradation, such as drainage and fires, and global environmental change, integrating ecological, hydrological, and socio-economic drivers through the lens of thresholds and emergent properties.
  • To propose actionable governance strategies prioritizing proactive protection over reactive mitigation, aligned with the SDGs and Paris Agreement.
While our model does not employ quantitative equations, its design and validation align with best practices for qualitative systems analysis in environmental policy research. As mentioned above, this study employs a qualitative systems causality model designed to map relationships between peatland degradation drivers and global change impacts such as CO2 emissions and biodiversity loss. The model prioritizes conceptual clarity over quantitative simulations, as its goal is to identify feedback loops and policy leverage points rather than generate numerical predictions.

1.1. Indonesia as a Critical Case Study

Indonesia hosts 23–24 million hectares (36%) of the world’s tropical peatlands [25,26]. By analyzing its peatland dynamics, this research offers a blueprint for peat-rich regions like the Amazon and Congo Basin to enhance sustainable peat resource use and management.

1.2. Study Proposals

(i)
Precautionary Governance:
  • Preemptive governance of peatland thresholds through IoT enforcement, (water tables ≤ 40 cm) and corporate liability for violations.
  • Adaptive 50-year protection transitioning to permanent safeguards.
(ii)
Global Climate Stabilization Zones:
  • Legally designate deep peat domes (>3 m) as non-negotiable reserves.
(iii)
Precision Monitoring:
  • IoT/LIDAR networks for real-time fire and subsidence detection.
  • Blockchain-tracked deforestation ban.
(iv)
ASEAN Transboundary Enforcement:
  • Regional firefighting protocols and palm oil levy (1–3%) funding rewetting, community programs, and technology upgrades.
These strategies align with SDGs 13–15 and the Paris Agreement, prioritizing peatlands as irreplaceable climate infrastructure.
This paper is structured as follows:
  • Section 2: Literature review on peatland–climate dynamics and policy gaps.
  • Section 3: Methodology integrating systems theory and causal loop modeling.
  • Section 4: Results on emissions, biodiversity loss, and hydrological impacts.
  • Section 5: Discussion of Indonesia’s peatland degradation drivers.
  • Section 6: Systems theory framework for multi-scale feedback analysis.
  • Section 7: Discussion.
  • Section 8: Policy implications for proactive governance and transboundary enforcement.
  • Section 9: Conclusions advocating peatlands as global climate infrastructure.

2. Literature Review

Tropical peatlands have emerged as critical ecosystems in global climate and biodiversity debates. Recent research has explored their degradation drivers, restoration strategies, and socio-ecological impacts through diverse disciplinary lenses. Mishra et al. (2021) [27], for instance, synthesize the biophysical dynamics of peatland degradation in Southeast Asia, emphasizing drainage, agricultural expansion, and fires as primary drivers of ecosystem collapse. Grounded in ecosystem management theory, their work advocates for restoration strategies like rewetting and revegetation but narrowly focuses on ecological recovery, sidelining socio-political complexities such as land tenure conflicts. Similarly, Sasmito et al. (2025) [28] employ quantitative modeling to estimate the carbon mitigation potential of peatland conservation, highlighting the role of policy interventions. While their land-use science approach underscores the urgency of reducing emissions, it neglects the institutional fragmentation that often undermines policy implementation in regions like Kalimantan.
Climate-centric studies further illustrate the limitations of siloed frameworks. Hapsari et al. (2022) [23] drawing on paleoecology and climate science, trace the long-term impacts of sea-level rise on Indonesian peatlands and propose adaptive measures like fire management. However, their analysis lacks actionable pathways for integrating these measures into national climate policies. Harenda et al. (2018) [24], rooted in carbon cycle theory, position peatlands as vital carbon sinks but adopt a linear cause–effect model, for example, drainage → emissions → climate impacts, overlooking feedback loops such as climate-induced droughts exacerbating degradation. This linearity persists even in socio-ecological studies: Budiningsih et al. (2024) [29], for example, stress multi-stakeholder collaboration in restoration efforts but treat ecological recovery and community livelihoods as parallel processes rather than interconnected systems.
A notable exception is Osaki et al. (2024) [30], whose transdisciplinary "Peatlogy" framework bridges natural sciences, social sciences, and policy-making. By analyzing circular interactions between peatland hydrology, land-use decisions, and governance, they challenge reductionist paradigms. Yet, their work stops short of developing a causality model to quantify these relationships, leaving a critical gap in tools for systemic analysis.
Collectively, these studies reveal three unresolved challenges. First, reductionist approaches dominate, with most works isolating ecological, carbon, or socio-economic dimensions rather than addressing their interdependencies. Second, linear frameworks—such as “drainage → emissions → climate change”—ignore feedback loops like climate feedback intensifying degradation. Third, scale disconnects persist: local impacts such as biodiversity loss in Sumatra are rarely linked to regional ones such as transboundary haze or global consequences, for example, methane emissions under warming scenarios.
Similarly, in Indonesia, current policies such as Indonesia’s PP 57/2016 focus on reactive mitigation, for example, post-fire restoration, but fail to address nonlinear degradation thresholds such as irreversible peat drying or transboundary feedback loops, for instance, haze-driven permafrost thaw. This study fills this gap by advocating for precautionary governance that prioritizes indefinite protection over costly post hoc fixes.
To address these gaps, this study introduces a systems causality model that integrates ecological, socio-economic, and political dimensions into a unified framework. Unlike prior studies, the model captures circular causality—for instance, drainage-driven peat drying increases fire risk, which accelerates carbon emissions and climate feedback loops, further destabilizing peatlands. It also bridges scale disconnects by linking local drivers, for example, oil palm expansion in Riau, to regional impacts such as hydrological disruption across Southeast Asia and global outcomes, for example, carbon emission contributions to atmospheric CO2 levels. By synthesizing transdisciplinary insights, the model advances beyond siloed approaches, offering a holistic tool to analyze degradation and design policies that balance ecological resilience with human needs.

3. Materials and Methods

3.1. Study Area

This study focused on Indonesia’s tropical peatlands, with primary emphasis on Kalimantan (Section 5), home to one of the country’s largest peatland areas and the region most severely affected by recurrent fires and with the highest rates of burning [31].

3.2. Data Collection

This study synthesized remote sensing data including Landsat-derived peat-fire carbon emissions data spanning major fire events (1997–2023) from two key sources:
(i)
The peer-reviewed literature, focused on drivers of peatland degradation such as drainage and fires and associated impacts including CO2 emissions, biodiversity loss.
(ii)
Policy documents from Indonesian governmental agencies, including but not limited to the following:
  • Ministry of Environment and Forestry (KLHK/MoEF):
    »
    Rencana Strategis Tahun 2020–2024 (Revisi 2021) [Strategic Plan 2020–2024 (2021 Revision)] (KLHK, 2021) [32].
    »
    FOLU NET SINK: Indonesia’s Climate Actions Towards 2030 [33].
    »
    The State of Indonesia’s Forests 2024: Towards Sustainability of Forest Ecosystems in Indonesia (MoEF, 2024) [34].
  • Legal enactments, such as
    »
    Government Regulation 71/2014 and Government Regulation No. 57/2016 and on Peat Protection;
    »
    Presidential Instruction No. 1/2016 on the Establishment of the Peatland Restoration Agency (BRG).
  • IUCN biodiversity data.

3.3. Systems Causality Modeling Approach

Figure 1 and Table 1 are foundational to our causal modeling approach, providing the visual and operational framework for analyzing peatland degradation dynamics. We define a “system” as a set of interconnected components such as peat hydrology, land-use decisions, and climate feedback loops that collectively produce emergent behaviors not evident from individual parts alone. Following this, peatland systems are conceptualized as complex socio-ecological systems interacting with five Earth subsystems (Section 7):
  • Biosphere (biodiversity loss);
  • Atmosphere (CO2 emissions);
  • Hydrosphere (water table collapse);
  • Lithosphere (subsidence);
  • Cryosphere (ice melt acceleration).

3.4. Variable Selection and Operationalization Followed a Two-Tier Framework

  • Table 1 defines all model components (drivers, thresholds, solutions) with descriptive categories. Components were defined using peer-reviewed Indonesian case studies. For example, a water table depth <40 cm defined irreversible drying [19].
  • Table 2 isolates empirically validated thresholds from Table 1 used for
    »
    Historical validation, for example, a water table < 40 cm predicting fires;
    »
    Scenario projections, for example, a 34% emission reduction from rewetting.
Demonstrably, rewetting policies (Table 1) were quantified using Table 2’s 34% emission reduction threshold to model restoration impacts.
As shown in Figure 1, feedback loops were constructed using Table 1’s relationships. It shows the causal structure of feedback mechanisms (R1/R2) and visually maps feedback loops (R1/R2) and cross-system interactions validated against Indonesian case studies.
»
Reinforcing loop (+) (R1):
Drainage (+) → peat drying (+) → fires (+) → emissions (+) → climate extremes (+) drought (+) → drying (+)
»
Validated by 2015 fires where drainage-triggered drought amplified emissions. Balancing loop (−) (R2)
Rewetting (+) → water > 4 cm (+) → fire reduction (−) → emissions (−) → climate stabilization (−)
was supported by a 34% emission reduction in rewetted areas.

4. Results

4.1. Multi-Dimensional Impacts of Peatland Degradation

Our analysis revealed three interconnected scenarios of peatland degradation impact (Section 7).
  • Local:
    »
    Loss of biodiversity, for example, a 15% decline in peat swamp forest species in Sumatra.
    »
    Increased peat subsidence (2–5 cm/year in drained areas).
  • Regional:
    »
    Transboundary haze from fires, affecting air quality in Southeast Asia.
    »
    Hydrological disruption, reducing water availability for 2 million people in Kalimantan.
  • Global:
    »
    Annual CO2 emissions of 0.8–1.2 Gt from Indonesian peatlands (5–10% of global peatland emissions).
    »
    Feedback loops accelerating climate change, for instance, methane release from rewetting attempts.

4.2. Systems Causality Model Outcomes

The model identified three critical leverage points for conservation:
»
Strict hydrological thresholds legally enforce water table depths ≤ 40 cm (PP 57/2016) to prevent oxidation and subsidence (2–5 cm/year).
»
Rewetting imperative: hydrological restoration of 590 km2/year contribute to mitigates 1.4–1.6 Mt CO2 annually.
»
Legally enshrine core peat domes (>3 m depth) as non-negotiable “Global Climate Stabilization Zones.”
»
Policy integration: aligning peatland conservation with SDGs 13, 14, and 15 enhances funding and enforcement.
Policy-ready framework: a precautionary governance framework designed to address irreversible peatland degradation.

4.3. Key Findings

Our system analysis revealed how localized drainage (a <40 cm water table) triggered planetary scale feedback through three empirical advances. Our study establishes three critical advances in understanding peatland dynamics.
We identified critical biophysical thresholds (a <40 cm water table depth, <4 mm/day rainfall) that serve as primary activation points for irreversible degradation. Historical events demonstrate that when these thresholds are breached, they consistently enable catastrophic outcomes through validated mechanisms:
Hydrological collapse: water tables below −40 cm trigger peat oxidation (2–5 cm/year subsidence) and create permanently flammable conditions.
Validation by major events:
  • The 1997 fires, where water tables dropped to −120 cm in Central Kalimantan’s Mega Rice Project.
  • The 2015 catastrophe, where rainfall fell to 1.0 mm/day during peak burning.
  • The 2023 escalation, where extreme drought (attributed to threshold conditions) enabled 1.16 Mha to be burned.
Despite robust regulatory frameworks like Indonesia’s PP 57/2016 mandating critical hydrological safeguards (≤40 cm water tables), persistent implementation barriers undermine peatland protection. Institutional fragmentation across ministries, for example, between the Ministry of Environment and Forestry (KLHK), Peatland Restoration Agency (BRG), and local governance, dilutes enforcement accountability, while economic pressures from palm oil—representing 4.5% of national GDP—create financial incentives that outweigh existing penalties. Our proposed “Global Climate Stabilization Zones” directly counter these structural gaps through IoT-driven accountability systems automating real-time violation fines and ASEAN transboundary governance integrating regional fire-suppression protocols with blockchain-tracked deforestation bans funded by a 1–3% palm oil levy.

5. Peatlands: An Overview

Peatlands are wetland ecosystems defined by a surface layer of partially decomposed organic material (peat) at least 30 cm thick, with organic content exceeding 30% [35,36,37]. These landscapes form under waterlogged, anaerobic conditions, accumulating organic matter at an exceptionally slow rate of 0.5–1 mm annually—a process requiring millennia to develop significant depth. Globally distributed across 175 countries, peatlands cover approximately 4 million hectares (3% of Earth’s land area) and serve as critical long-term carbon sinks [38]. Unlike forests, which cycle carbon through seasonal growth and decomposition, intact peatlands sequester carbon almost permanently, storing it in waterlogged peat layers. However, they also emit methane (CH4), a potent greenhouse gas with a global warming potential 28–84 times greater than CO2 over 100- and 20-year timescales, respectively [37,39].
Tropical peatlands, distinct from their moss-dominated boreal counterparts, are characterized by dense peat swamps or mangrove forests where organic matter accumulates under waterlogged conditions [40]. When disturbed by drainage or fire—common in land conversion practices—these ecosystems shift from carbon sinks to catastrophic emitters, releasing more than 2 Gt CO2eq annually (5% of global anthropogenic emissions) [37,39].
Peat fires pose unique risks: they smolder underground, resist suppression, and emit mercury at rates 15 times higher than forest fires, exacerbating air pollution and climate feedback loops [41]. Beyond their climatic role, tropical peatlands support unparalleled biodiversity and provide vital ecosystem services, including flood mitigation and freshwater regulation, making their preservation a global priority.

5.1. Peatlands in Indonesia

Peatlands are critical global carbon reservoirs, storing over 550 gigatonnes (Gt) of carbon—equivalent to 42% of all soil carbon and surpassing the carbon stored in all the world’s forests combined [37,38,42]. Tropical peatlands alone account for 152–288 Gt, with Southeast Asia holding 68.5 Gt, of which 57.4 Gt (84%) is concentrated in Indonesia’s peatlands—the world’s second-largest tropical peatland area, spanning 22.5 million hectares across Sumatra, Kalimantan, and Papua [43,44].
Indonesia’s peat soils store 20 times more carbon per hectare than its above-ground vegetation, representing 74% of the nation’s soil carbon pool [43,44]. When undisturbed, these ancient ecosystems (formed over 15,000 years) act as fire-resistant carbon sinks, underscoring their pivotal role in mitigating global warming and advancing climate resilience. Indonesia’s peatlands are concentrated in Sumatra, Kalimantan, and Papua (Figure 2).

5.2. Socio-Ecological Importance

  • Social Importance
Indonesian peatlands are deeply intertwined with regional livelihoods, biodiversity, and cultural heritage. Economically, they sustain approximately 40 million rural and Indigenous people through commercial agriculture such as oil palm and rubber, subsistence farming, and resource extraction (timber, coal [46]. These activities underpin national economic growth while supporting local communities. Beyond their economic value, peatlands provide irreplaceable ecological and cultural benefits. Their regulating services—such as erosion control, water purification, and carbon sequestration—mitigate floods and droughts, while cultural practices tied to these ecosystems sustain Indigenous identities.
For instance, the Dayak Ngaju communities of Kalimantan rely on peatlands for spiritual rituals and utilize the forest for timber, non-timber products such as rattan and bamboo, medicinal plants, and food. Hunting, fishing, and gathering are integral traditions of the Ngaju people in peat swamp forests. Key resources include unidentified local species from peatland waterways and various medical plants such as (a) Bajakah (Spatholobus littoralis), a woody vine traditionally used for medicinal purposes; Asam Limpasu (Baccaurea lanceolata), a fruit-bearing tree used for food and traditional remedies; (c) Akar Kuning (Arcangelisia flava), a medicinal root known for its anti-inflammatory properties; and (d) Sungkai (Peronema canescens), whose leaves are used to treat fevers and malaria [47].
  • Ecological Significance
Indonesian peatlands are biodiversity hotspots, hosting 927 plant species (300 endemic) and critical habitat for endangered fauna like orangutans, Sumatran tigers, and Sunda clouded leopards. Ecologically, Indonesian peatlands are biodiversity hotspots, hosting 927 plant species with more than 300 endemic species and providing critical habitat for endangered fauna such as Bornean orangutans, Sunda clouded leopards, and Bornean elephants [47,48]. Kalimantan, in particular, is renowned for its rich biodiversity, as illustrated in Table 3.
Coastal peatlands sustain unique blackwater habitats, supporting endemic fish species such as Channa sp., Wallago leeri, Anabas testudineus, Trichogaster pectoralis, Trichogaster trochopterus, and the endangered Scleropagus Formosus [51]. Undisturbed peatlands also act as reservoirs of germ plasm, offering genetic resources to enhance disease resistance and agricultural productivity [52]. Undisturbed peatlands are vital for sustaining biodiversity, as their ecological integrity supports specialized species through unique hydrological and vegetative conditions. Intact peatlands, characterized by their water-saturated soils, slow organic decomposition processes, and specialized vegetation adapted to acidic, nutrient-poor conditions, are indispensable for sustaining the critical ecosystem services shown in Table 4.

Indonesia’s Peatlands: A Linchpin for Climate Resilience and Biodiversity Conservation

Peatlands possess unique physical characteristics, including irreversible drying upon drainage, high porosity, and low vertical hydraulic conductivity. When dried, peat becomes highly erodible, flammable, and water-repellent, complicating rehydration and restoration efforts. Drainage disrupts their hydrological functions, accelerating degradation. Anthropogenic activities, particularly drainage for agriculture or plantations, trigger peat oxidation and subsidence (2–5 cm annually), while fire-prone conditions release stored carbon as CO2 and toxic haze. For instance, burning just one centimeter of peat emits ~5 Mg of carbon per hectare (18 Mg CO2-equivalent), intensifying global warming [53]. Regions such as Papua, Central Kalimantan, and Riau—home to 50% of Indonesia’s peatlands—are disproportionately affected. Papua alone accounts for 6.3 million hectares (28% of the national total) [45].
Degradation of these ecosystems accelerates global warming and disrupts regional hydrology, jeopardizing water access for millions. Indonesia’s peatlands thus epitomize the urgent need for conservation strategies that prioritize avoiding degradation over costly post hoc restoration. Protecting Indonesia’s peatlands is not only vital for safeguarding biodiversity, such as orangutans, clouded leopards, and endemic fish, and ecological integrity but also for fostering sustainable development and meeting global climate goals. Their conservation represents a critical nexus between environmental stewardship and climate action.

5.3. Conservation and Policy Integration

Indonesia offers a pivotal opportunity to combat climate change through natural climate solutions (NCSs), including peatland protection, improved management, and restoration to enhance carbon sequestration. The country has prioritized peatland conservation in its climate agenda, such as the following:
  • FOLU Net Sink 2030, targeting emissions of −140 Mt CO2e by 2030 and −304 Mt CO2e by 2050 [34].
Enhanced NDC, aiming for a 31.89% emissions reduction through national efforts, rising to 43.20% with international support by 2030 [54].

Indonesian Peatland Conservation: Policy and Regulatory Framework

The conservation of Indonesia’s peatlands is anchored in a robust legal and policy framework that integrates constitutional mandates, national legislation, and targeted presidential directives. While the 1945 Constitution of Indonesia does not explicitly mention peatlands, its Article 28H(1) guarantees the right to a healthy environment—a provision interpreted to encompass peatland protection as part of sustainable development. This constitutional foundation has informed a multi-layered governance structure, including flagship policies such as the National Action Plan on Climate Change (2007), which aims to rehabilitate 53.9 million hectares of degraded peatlands and forests by 2050, prioritizing fire prevention, carbon sink enhancement, and biodiversity conservation. Subsequent laws and regulations balance ecological preservation with socio-economic needs, ranging from Government Regulation No. 57/2016 (enforcing ≤40 cm water tables) to Presidential Instruction No. 1/2016, which established the Peatland Restoration Agency (BRG) to coordinate large-scale rewetting and fire suppression. These efforts align with international commitments like the Paris Agreement and Sustainable Development Goals (SDGs), underscoring peatlands’ role in climate resilience and biodiversity protection. Table 5 summarizes this comprehensive governance framework, detailing key legislative instruments, their objectives, and alignment with national and global sustainability targets.
Recent initiatives, such as the Peatland Restoration Agency (BRG), established in 2016, have improved coordination. The BRG has rewetted 3.6 million hectares of peatlands as of 2023, reducing CO2 emissions by 7.8 million tonnes annually. However, conflicting economic priorities—such as palm oil production (59% of global output in 2023)—often overshadow conservation goals. Indonesia’s peatland policies also align with international commitments, including UNEP and the Roundtable on Sustainable Palm Oil (RSPO).

5.4. Indonesia’s Peatland Dynamics: A Comprehensive Analysis

Indonesia’s peatlands, spanning over 15 million hectares, represent one of the world’s most vital carbon sinks, yet decades of mismanagement and exploitation have transformed these ecosystems into significant sources of greenhouse gas emissions. The degradation stems from a complex interplay of weak policy enforcement, agricultural expansion, and recurrent fires, with profound implications for global climate stability and regional health.

5.4.1. Regulatory Efforts and Shortcomings

Despite the inclusion of environmental protection in Indonesia’s 1945 Constitution (Article 28H) and the establishment of a robust legal framework, peatland degradation has persisted, driven primarily by agricultural expansion, commercial development, and weak enforcement. This degradation accelerated in the 1970s with state-led large-scale conversions of peatlands for rice cultivation, culminating in the 1995 Mega Rice Project—a disastrous initiative that drained approximately one million hectares of Central Kalimantan’s peat swamp forests [55]. The project’s extensive drainage infrastructure destabilized peat hydrology, triggering subsidence (3–5 cm/year) and oxidation, which released millennia-old carbon stores and increased fire vulnerability [56,57].
Peat degradation or subsidence is driven by (i) compaction, the compression of aerated peat layers, (ii) consolidation, the loss of buoyancy in drained peat, and (iii) oxidation, the aerobic decomposition of organic matter to CO2 [56,57]. Human activities exacerbate these processes through (i) drainage for agriculture (oil palm, rice), which lowers water tables, drying peat; (ii) land conversion which destroys peat-forming vegetation; and (iii) fires, used for land clearance, which ignite underground peat layers.

5.4.2. Historical Fire Episodes and Impact: Key Events

The 1982–1983 El Niño–Southern Oscillation (ENSO) event caused severe drought conditions in Indonesia, which intensified wildfires. Approximately 5 million hectares of land burned during that period, including 3.5 million hectares in East Kalimantan alone [58,59]. While limited monitoring technology at the time obscured precise quantification of emissions, the fires released substantial amounts of CO2 and other greenhouse gases, contributing to regional air pollution and global carbon feedback loops.
The 1994 “Great Fire of Borneo,” centered in Central Kalimantan, Indonesia, burned 5.11 million hectares of forests, brush, and grasslands, driven by widespread land-clearing practices and exacerbated by drought conditions [58]. The fires produced a dense, transboundary haze that blanketed Indonesia, Malaysia, and Singapore from August to October 1994, severely degrading air quality. Elevated levels of carbon monoxide (CO), nitrogen dioxide (NO2), and particulate matter (PM10/PM2.5) triggered a surge in respiratory emergencies, particularly among children, with asthma cases rising sharply across the region [60]. In Indonesia’s extensive peat swamp forests, falling water tables dried out organic peat, enabling deep smoldering combustion that destroyed standing biomass and ignited subsurface layers. These underground fires persisted for weeks, releasing toxic compounds and contributing to a tropospheric ozone spike recorded at Watukosek, Indonesia—a phenomenon linked to large-scale biomass burning [61]. Tropospheric ozone is an important air pollutant and greenhouse gas, which plays a key role in atmospheric chemistry. The haze’s bioclimatic impacts disrupted ecosystems, reduced visibility, and strained public health systems, underscoring the vulnerability of peatlands to human-induced degradation and climate variability.
The 1997–1998 peatland fires, intensified by severe El Niño drought conditions, devastated 3.4 million hectares of land in Indonesia, including 750,000 hectares of peatlands in Kalimantan, and 300,000 and 400,000 ha in Sumatra and West Papua, respectively [62,63]. These fires released 0.81–2.57 gigatonnes (Gt) of carbon—equivalent to 13–40% of global fossil fuel emissions at the time—primarily from the combustion of ancient peat deposits [63]. This catastrophic event temporarily elevated Indonesia to the world’s third-largest greenhouse gas (GHG) emitter in 1997, surpassing industrialized nations like Japan and Germany [64]. Tropospheric ozone (O3), was significantly elevated during the 1997 Indonesian peatland fires due to photochemical reactions between nitrogen oxides (NOx) and volatile organic compounds (VOCs) emitted by burning vegetation and peat. This ozone spike underscored the complex atmospheric consequences of large-scale biomass burning [61]. Approximately 20 million people in Indonesia suffered from respiratory diseases such as bronchial asthma and acute respiratory infections [65].
The economic losses from the 1997/98 peatland fires—encompassing damage to agriculture, forestry, CO2 emissions, ecosystem services (flood protection, erosion control, and carbon sink loss), and health impacts from toxic haze—were estimated at USD 8.5 to USD 9.4 billion. The breakdown by sector is detailed in Table 6.
In 2006, Indonesia experienced its most severe forest fires since 1998, concentrated in Kalimantan (Borneo) and Sumatra. Driven largely by land-clearing practices, these fires produced a hazardous haze that blanketed Southeast Asia, severely degrading air quality and causing widespread health crises. During the 2006 El Niño event, emissions from a 2.79 Mha study area in Central Kalimantan alone released an estimated 49.15 ± 26.81 million tonnes (Mt) of carbon—equivalent to 10–33% of the European Union’s total transport sector emissions that year [67].
The 2015 Indonesian fire crisis marked one of the most catastrophic environmental disasters in the nation’s history, surpassing the intensity of previous major episodes like 2006 and 1997. Driven by a potent combination of El Niño-induced drought and widespread land-clearing practices on degraded peatlands, the fires released an estimated 884 million tonnes of CO2—97% originating from Indonesian peatlands—equivalent to Germany’s annual fossil fuel emissions [68].
Satellite data revealed unprecedented smoke pollution, with PM2.5 concentrations exceeding WHO hazard thresholds by 10–20 times, blanketing Southeast Asia in toxic haze. This pollution caused over 100,000 premature deaths (91,600 in Indonesia, 6500 in Malaysia, and 2200 in Singapore) and 500,000+ severe respiratory cases, disproportionately affecting children, the elderly, and low-income communities [69]. Economically, the fires inflicted USD 16 billion in losses—double the value of Indonesia’s 2014 palm oil exports (USD 8 billion) and exceeding the entire palm oil industry’s annual economic contribution (USD 12 billion). The crisis overwhelmed healthcare systems, paralyzed transportation networks, and decimated agriculture and tourism, compounding long-term socio-economic vulnerabilities [17]. Peat-swamp forest (PSF) fires cause extreme tree mortality and significantly reduce tree species richness and non-tree flora abundance. Although few fauna species are exclusively restricted to PSFs, the unprecedented scale of 2015 peatland and lowland forest burning implied catastrophic habitat loss. This event likely inflicted the following:
  • Widespread negative impacts on most PSF-dependent species;
  • Severe population declines for many taxa;
  • Local extinctions of vulnerable species in heavily affected areas.
Mechanistically, decades of peatland drainage for palm oil and pulpwood plantations amplified flammability, while weak governance allowed slash-and-burn practices to persist. An analysis of rainfall patterns shows degraded peatlands in Kalimantan now exhibit heightened sensitivity to drought (<4 mm/day rainfall), with fire risk escalating even under historically wetter conditions [9].
The 2015 crisis underscored the dire consequences of prioritizing short-term economic gains over ecological resilience. Emissions from peat fires accelerated global warming, destabilized regional climate systems, and locked Indonesia into a vicious cycle where environmental degradation perpetuates human suffering. To avert future disasters, experts have urged systemic reforms: banning peatland drainage, enforcing fire-free land management, and scaling restoration efforts like rewetting and paludiculture. Without such measures, Indonesia risks entrenching a feedback loop that jeopardizes local livelihoods, regional stability, and global climate targets.
Following the catastrophic 2015 fires, Indonesia implemented moratoriums on peatland conversion and strengthened the enforcement of environmental laws like No. 32/2009 (environmental protection) and No. 41/1999 (forestry). These measures initially reduced burned areas by 60% between 2016 and 2017, with only 38,363 ha and 165,528 ha burned in those years, respectively, emitting 96.7 million tonnes and 12.5 million tonnes of CO2 [70,71]. Despite strengthened peatland protections, enforcement gaps and corporate noncompliance persist. Dozens of companies have violated peatland regulations—exploiting weak monitoring systems and negligible penalties—undermining conservation progress. By 2018, burned areas surged to 529,267 ha—a threefold increase from 2017—as drainage and illegal burning resumed [72].
The 2019 fires marked a devastating resurgence, burning 1.64 million hectares (including 494,000 hectares of peatlands) and releasing 624–708 million tonnes of CO2eq, depending on measurement methodologies These fires damaged critical biodiversity hotspots like Berbak-Sembilang National Park and cost USD 5.2 billion in economic losses, including healthcare burdens, agricultural disruption, and tourism declines [73]. Independent studies highlighted discrepancies in emissions’ estimates, underscoring the difficulty of quantifying underground peat fires—a persistent challenge in Indonesian fire management.
In 2023, Indonesia faced a severe escalation in wildfires, with 1.16 million hectares burned—a fivefold increase from 2022. This spike was driven by extreme drought conditions linked to El Niño, which prolonged dry spells and intensified fire susceptibility. Kalimantan, a peatland-rich region, bore the brunt of the crisis, recording the highest fire frequency and most recurrent incidents, as degraded peat soils and land-use practices amplified flammability [74]. Overall, recurrent fires between 1997 and 2023 burned 1.16–2.6 million ha annually, reflecting systemic vulnerabilities in land-use governance and peatland conservation [75].
Indonesia’s peatland degradation persists despite comprehensive policies, rooted in a critical gap in recognizing peatlands’ role in stabilizing the Earth system and the cascading consequences of their destabilization across interconnected planetary spheres—biosphere, hydrosphere, atmosphere, lithosphere, and cryosphere. This systemic blind spot erodes political will, perpetuates fragmented enforcement, and prioritizes short-term economic gains such as oil palm plantation expansion over long-term resilience. Without a holistic understanding of feedback loops—such as drainage-induced emissions amplifying climate change, which in turn intensifies fire risks—policies remain siloed, enforcement lax, and profit-driven exploitation dominant.
A system dynamics approach, which maps these interdependencies and quantifies trade-offs between conservation and degradation, can reframe peatlands from crisis zones to planetary stabilizers. By illuminating how peatland health underpins global climate stability, biodiversity, and human well-being, this approach can foster the political and societal urgency needed for genuine conservation. The following discussion introduces systems theory as a tool to conceptualize Earth system interconnectedness, equipping policymakers, industries, and the public to advance systems-literate governance—where ecological resilience and sustainable development are mutually reinforcing priorities.

6. Systems Theory: A Framework for Unlocking Complexity

Indonesia’s peatland conservation impasse epitomizes the Anthropocene dilemma: peatland ecosystems critical to climate resilience and global environmental stability are being sacrificed for short-term growth. Yet, this crisis also presents an opportunity—to pioneer systems-literate governance that reconciles economic growth with ecological resilience and human well-being. To achieve this, drawing on systems theory, we develop a causal loop analysis to unlock peatlands’ complex dynamics from the local and global context. Systems theory provides a holistic lens to analyze peatlands as interconnected socio-ecological systems.
Systems theory is an interdisciplinary framework that provides a set of principles for analyzing the relationships, interactions, and interdependencies among components within complex systems. It views systems—such as peatland ecosystems, biological systems, or climate systems—as irreducibly integrated wholes. This perspective emphasizes the importance of understanding the system as a whole, rather than focusing solely on its individual parts [76]. Systems theory is particularly valuable for analyzing complex systems that cannot be fully understood, rationalized, or predicted using traditional linear approaches. By focusing on the interactions between components, systems theory reveals emergent properties of system interactions.

6.1. Core Concepts of Systems Theory and Dynamics [76,77,78]

  • System: A network of interconnected and interdependent components (e.g., peat soil, water, vegetation, wildlife) that function collectively to achieve a unified purpose. The integrity of a system depends on these relationships; isolated components cannot sustain functionality.
  • Systems Thinking: A holistic approach that examines interactions between components rather than isolating individual elements. It identifies emergent patterns, such as feedback loops, to predict system behavior over time.
  • System Behavior: The dynamic and often nonlinear interactions among components in response to external stimuli. For example, reduced rainfall in peatlands lowers water tables, increasing flammability and biodiversity loss.
  • Thresholds: Tipping points. A threshold is not a maximum limit but rather a critical point at which a significant change or new action is initiated. A well-known example of a threshold is the 1.5 °C global temperature increase target established by the Paris Agreement in 2015, which signifies a critical tipping point for climate systems or peat moisture <100%, which triggers irreversible drying.
  • Emergent Properties: Emergent properties are properties that arise only when components interact as a whole. Emergent properties are collective properties of the overall system. The resultant emergent behaviors bring together a new combination of capabilities, which can be either beneficially stabilizing or potentially harmful. For example, the resilience of an ecosystem is an emergent property that results from the interactions between species, nutrient cycles, and environmental conditions (for example, outcomes like transboundary haze or biodiversity loss arising from component interactions).

6.2. System Dynamics

System dynamics, grounded in systems thinking and systems theory, is a methodology for modeling and analyzing the time-varying behavior of interconnected system components. It provides tools for understanding how systems evolve over time and how their components interact to produce dynamic outcomes. System dynamics is particularly useful for capturing the complexity and emergent properties of systems.
As shown in Figure 3 below, the primary elements of system dynamics include the following:
  • Stocks: accumulations of units within a system such as accumulation of water or carbon in a peatland.
  • Flows: the movement of units across system boundaries over time (e.g., inflow and outflow of water in a peatland, carbon emissions).
  • System signals: Systems process input signals denoted as x(t) and produce output signals denoted as y(t). These signals are typically represented by stocks and flows (Figure 3). Signals, which convey information within a system, are often analog in nature, meaning they vary or fluctuate as a function of time, t. That is, they are dynamic and convey information to stimulate system responses (Figure 3). For example, in an ecological system, the water level (a stock) in a peatland system changes in response to rainfall (an inflow) and evaporation (an outflow). These dynamics are critical for understanding how systems respond to external stimuli and maintain their functionality.
  • In system dynamics, an excitatory input x(t) acts as a positively weighted signal that perturbs the system. This input becomes active only when its magnitude exceeds a predefined threshold, triggering a response that drives the system away from equilibrium, for example, initiating cascading effects like peat drainage or fire cycles. Conversely, an inhibitory input y(t) counterbalances the excitatory signal by applying negative feedback. This input suppresses the system’s deviation from equilibrium, restoring stability and mitigating runaway dynamics such as policy interventions to block drainage or restore hydrology.

6.3. Feedback Loops and Stability

Feedback loops are central to system dynamics, regulating stability or driving collapse.
There are two main types of feedback mechanisms:
  • Positive feedback loops: These loops amplify changes in the system, driving it away from equilibrium. They are often associated with positive feedback mechanisms, where the output reinforces the input, leading to exponential growth or decline. For example, in a peatland ecosystem, drying conditions can create a positive feedback loop, where reduced water levels lead to further drying and increased flammability. For example, peatland drainage lowers water tables, drying peat and increasing fire risk, which releases more CO2, accelerating climate change and further drying peatlands (Figure 3).
  • Negative feedback loops (balancing feedback loops): These loops counteract changes, promoting stability and equilibrium. They are associated with negative feedback mechanisms, where the output dampens the input, restoring the system to its resting state. For instance, water recharge in a peatland can create a negative feedback loop, counteracting drying conditions and restoring hydrological balance (Figure 3) For example, peatland drainage lowers water tables, drying peat and increasing fire risk, which releases more CO2, accelerating climate change and further drying peatlands.
It may be noted that the terms positive and negative do not refer to a value-based assessment of “good” or “bad,” but rather to the type of change that takes place in the system. As indicated above, positive feedback loops happen when a stimulus causes an amplification effect that reinforces change in the system after being fed back to its source (Figure 3). Positive feedback amplifies system output, resulting in growth or decline. It is often associated with system destabilization, although there are exceptions. Negative feedback dampens outputs, stabilizing the system around an equilibrium point, bringing the system back to homeostasis. A property of a homeostatic system is resilience defined as the ability of the system to return to stability in the face of perturbations. Furthermore, feedback loops are closely associated with vicious cycles and virtuous cycles:
  • Vicious cycles: A vicious cycle is a self-reinforcing chain of events driven by a positive feedback loop. In such cycles, each iteration amplifies the effects of the previous one, leading to destabilization and adverse outcomes. Vicious cycles often lack a tendency toward equilibrium or homeostasis, at least in the short term. An example of a vicious cycle/self-reinforcing loop driven by positive feedback is deforestation → soil erosion → biodiversity loss→ reduced carbon storage → increased emissions.
  • Virtuous cycles: A self-balancing loop where each event enhances the beneficial effects of the next. These cycles are often driven by negative feedback loops, which counteract destabilizing influences and promote stability. An example of a stabilizing loop driven by negative feedback is peatland conservation/rewetting → hydrological recovery→ biodiversity rebound → enhanced carbon storage → climate mitigation.
As shown in Figure 3, feedback loops control systems (causal circularity) are fundamental to system dynamics, enabling systems to maintain homeostasis, and resist external perturbations. Feedback loop may be defined as a cyclical chain of cause-and-effect relations that forms a loop, starting from the input signal which is applied at the system input to produce an output response signal. The response signal provides the stimuli to trigger the generation of an input that is returned to the system to control its behaviors.
Unlike external inputs, which originate from outside the system, feedback loops are intrinsic mechanisms that arise from within the system itself. This internal origin allows feedback loops to play a self-regulatory role, enabling the system to adapt dynamically to changing conditions. As noted by Courtney Brown (2008) [79], feedback loops are a fundamental feature of complex systems, providing a mechanism for self-correction and stability.

7. Discussion

Using causal loop diagrams (Figure 3 and Figure 4), we analyzed how Indonesia’s peatland degradation interacted with Earth’s subsystems, exposing local-to-global feedback dynamics.
Indonesia’s recurring peatland fires epitomize a socio-ecological crisis, driven by short-term economic gains conflicting with long-term sustainability. As interconnected socio-ecological systems, peatlands are embedded within Earth’s biosphere, interacting dynamically with its subsystems (Figure 3 and Figure 4), storing 57 GtC, equivalent to 6–8 years of global fossil fuel emissions.
In their intact state, peatlands sustain self-regulating cycles through negative feedback loops (Figure 3), maintaining equilibrium across Earth’s subsystems [80]. These mechanisms ensure resilience and stability through the following processes:
  • Ecological interaction
    • Flora and fauna synergy (Table 3)
      »
      Plants such as Nepenthes pitcher plants provide habitat and food for the fauna, while animals including orangutans and pollinators enable seed dispersal, nutrient cycling, and pest control.
      »
      Example: Orangutans disperse seeds across peat swamp forests, fostering vegetation regrowth and carbon sequestration.
  • Hydrological regulation
    • Water recharge
      »
      Peat soils act as sponges, absorbing rainfall and stabilizing groundwater levels. This buffers against droughts and floods while maintaining moisture levels critical for peat integrity.
      »
      Carbon sequestration: Waterlogged conditions inhibit decomposition, preserving 57 GtC stored in Indonesian peatlands and preventing CO2 release.
  • Biocapacity and ecosystem services
    • Clean water filtration: Peatlands filter pollutants, providing clean water for surrounding ecosystems and human communities.
    • Microclimate stabilization: Evapotranspiration cools local climates, mitigating heat extremes.
    • Biodiversity sanctuaries host endangered species (e.g., Sumatran tiger, false gharial) and unique flora, sustaining ecological networks.
  • Peatland ecosystem functions (Table 4):
    • Climate regulation: As Earth’s largest terrestrial carbon sink, peatlands capture and store carbon, offsetting ~6–8 years of global fossil fuel emissions.
    • Climate adaptation regulates water flow during extreme weather, reducing flood risks and recharging aquifers.
Intact peatlands exemplify nature’s balancing act, where negative feedback loops sustain carbon storage, hydrological stability, and biodiversity. Their degradation disrupts these cycles, underscoring the urgency of conservation to preserve their irreplaceable role in global climate and ecological resilience (Figure 4).
  • Legend
    • Rsee 02 00017 i001 Red arrows: Red arrows: Positive feedback (reinforcing) loops (vicious cycle) Positive feedback (reinforcing) loops → Vicious cycles, for example, “Drainage → Peat drying → More drainage.
    • Rsee 02 00017 i002 Green arrows: Negative feedback (balancing) loops → Virtuous cycles/solutions, for example, “Rewetting → Fire suppression → Less degradation.”
    • (+) = Amplifying effect, for example, drainage → (amplifying) peat drying
    • (−) = Inhibiting effect, for example, rewetting→ (inhibiting) fire risk
    • ↑ Increases/enhances:
    • » Degradation context: “Peat drying ↑ Fire risk” (Red +, ↑).
    • » Solution context: “Rewetting ↑ Water tables” (Green −, ↑).
    • ↓ Decreases/suppresses:
    • » Degradation context: “Biodiversity loss ↓ Ecosystem functions” (Red +, ↓).
    • » Solution context: “Rewetting ↓ Biodiversity loss” (Green −, ↓).
    • → Neutral directional link, for example, “Agricultural expansion → Drainage”
  • Key loop identification:
    • R (reinforcing): Drainage → (+) peat drying → (+) fires → (+) emissions
    • B (balancing): Rewetting →(−) fire Risk → (−) emissions → (+) climate stability
    • pIOD used in the diagram refers to Positive Indian Ocean Dipole events
  • Narration: This diagram illustrates cascading impacts from Indonesian peatland degradation. The interpretation of the linkages is as follows:
    Starting at local scale (left), human drivers (agricultural expansion, weak enforcement) initiate degradation, emitting GHGs that intensify local impacts (biodiversity loss, haze) and amplify regional climate extremes (heatwaves, abnormal rainfall).
    Regional consequences (center): Atmospheric pollution crosses borders, for example, producing haze in Malaysia and Singapore. Climate disruptions fuel socioeconomic/health crises.
    Global cascades (right): GHG accumulation → cryosphere collapse (glacier retreat, albedo loss). Ocean warming and sea-level rise create transboundary risks.
For example:
  • El Niño/pIOD amplification is located below “Weak enforcement” and parallel to “Human-induced degradation” and positioned as drought intensifiers:
    »
    Local warming impact: Heat intensifications and lowering water tables, for example, 2015 fires at 1.0 mm/day rainfall.
    »
    Regional extremes: Deadly heatwaves and drought and transboundary pollution/prolonged haze.
  • Zoonotic-pandemic pathways are positioned almost parallel to “Atmospheric GHG concentrations”:
    »
    Habitat fragmentation from peatland degradation displaced fruit bats, which led to the outbreak of Nipah virus through the following:
    .
    Spillover forces human-wildlife contact → local Nipah virus outbreak from displaced fruits bats in Central and East Kalimantan → spreads to Malaysia and Singapore across the South China Sea (regional impact).
    .
    Global health impacts: Ancient pathogen release via global warming-induced permafrost thaw risks global pandemic potential.
  • The rewetting-stabilization/policy intervention loop is a balancing loop where rewetting enforces water table ≥−40 cm, restores hydrology and reduces fire risk, leading to lower emissions and the stabilization of the climate system.
Rewetting policies water table ≥−40 cm (IoT-enforced water tables/threshold restoration) → hydrological recovery → fire reduction → lower CO2 emissions → climate stabilization.
Scientific linkages from peatland emissions to global feedback loops (shown in thick red arrows).
Core emission pathway:
  • Human-induced peatland degradation → atmospheric GHG concentrations
    »
    Albedo feedback loop: Atmospheric CO2 emissions/concentrations → global warming → glacier retreat → exposed dark surface → reduced albedo → increased heat absorption → (leads to further) global warming.
    »
    Glacier/permafrost thaw → exposed darker surfaces (land/water) → reduced solar reflectivity → increased heat absorption → amplified warming.
  • Acidification pathway: Atmospheric CO2 emissions/concentrations → CO2 dissolution in oceans → lowered pH→ carbonate dissolution → marine ecosystem collapse.
These pathways demonstrate how Indonesia’s peat degradation initiates planetary-scale crises through well-established geochemical cascades.
  • Visual guide to diagram connections:
    • Start: Human-induced degradation (top left).
    • Critical junction: Atmospheric CO2 concentrations (center top):
      • Albedo branch:
        »
        Atmospheric CO2 emissions/concentrations → glacier retreat → albedo effect → ⇧ self-reinforcing warming (⇧ = reinforcing).
        »
        Glacier/permafrost thaw → exposed darker surfaces (land/water) → reduced solar reflectivity → increased heat absorption → amplified warming.
      • Acidification branch
        »
        Atmospheric CO2 emissions/concentrations → ocean acidification → global/social health impacts.
      • Policy implications: Breaking the emission pathway via rewetting and peat protection simultaneously mitigates both feedback loops, conserving Arctic ice and marine ecosystems.
These interdependencies reveal planetary-scale risks: peatland degradation can initiate self-reinforcing cycles, such as carbon emissions intensifying global warming, which accelerates polar ice melt and sea-level rise. Local land-use practices thus carry global consequences, demanding urgent, coordinated governance to prevent irreversible ecological tipping points and meet climate goals. A systems-level analysis—integrating feedback loops, critical thresholds, and subsystem linkages—is essential for designing strategies that harmonize ecological resilience, climate mitigation, and human livelihoods. The following sections dissect these dynamics to advance actionable pathways for conservation and sustainable development.

7.1. Interconnected Feedback Loops and Degradation Mechanisms

Before dissecting the feedback dynamics driving peatland collapse, it is essential to define threshold levels—critical tipping points beyond which degradation becomes irreversible and triggers emergent planetary risks. In complex systems like peatlands, thresholds demarcate domains of systemic stability from cascading failure. For Indonesia’s peat ecosystems, three non-negotiable thresholds govern resilience:
  • A hydrological threshold (<−40 cm water table depth) triggers irreversible peat oxidation and subsidence (2–5 cm/year).
  • A climatic threshold (<4 mm/day rainfall) locks in permanent fire vulnerability independent of historical baselines.
  • A carbon sink threshold (<3 m peat depth) converts millennial carbon reservoirs into permanent emission sources.
Breaching these boundaries initiates self-reinforcing feedback loops that propagate degradation across scales. For instance, transgressing the hydrological threshold (<−40 cm) activates the drainage–desiccation–oxidation loop: drainage → peat drying → fire risk → CO2 emissions → radiative forcing. This cascade transforms local threshold breaches into planetary risks, such as Arctic permafrost thaw via climate teleconnections. The following analysis reveals how anthropogenic actions cross these thresholds, triggering irreversible Earth subsystem destabilization.

7.2. Thresholds and Emergent Properties: The Tipping Points of Planetary Destabilization

Peatland ecosystems function as complex adaptive systems where breaching critical thresholds triggers nonlinear, irreversible shifts with cascading planetary consequences. Our analysis identified three cardinal tipping points governing their stability:
  • Hydrological collapse (<−40 cm water table):
    »
    Induces irreversible peat oxidation, subsidence (2–5 cm/year), and permanent loss of land utility.
  • Combustion threshold (<4 mm/day rainfall):
    »
    Locks in year-round fire vulnerability independent of historical baselines, elevating fire probability by 85% in degraded peatlands.
  • Carbon sink collapse (<3 m peat depth):
    »
    Converts millennial carbon reservoirs (>500 tons C/ha) into permanent emission sources.
When surpassed, these thresholds generate emergent properties—unpredictable system-level outcomes propagating across the Earth’s subsystems. The interaction of these thresholds across planetary subsystems (Figure 4) is epitomized in Indonesia’s peatland feedback model (Figure 5): breaching the lithospheric hydrological threshold (<−40 cm water tables) initiates cascades where drainage emissions force atmospheric–cryospheric teleconnections, accelerating Arctic thaw at more than 2–4 times global rates, while fire-induced iron-ash runoff triggers hydrospheric collapse via algal blooms. Simultaneously, biospheric fragmentation elevates zoonotic risks, completing a vicious cycle of planetary destabilization that scales from local degradation to global sea-level rise.
For instance, breaching the lithospheric hydrological threshold (<−40 cm water tables) triggers marine ecosystem collapse through iron-ash runoff, inducing algal blooms that cause 100% coral mortality in the Mentawai Islands—a cascading failure collapsing fisheries (40–60% stock decline). Simultaneously, drainage-driven peat emissions (0.8–1.5 GtC/year) generate climate teleconnections, accelerating Arctic permafrost thaw at 2–4× the global rate and destabilizing ancient carbon stocks [81]. Critically, breached climatic thresholds—exemplified by rainfall reduction to 1.0–1.5 mm/day during Kalimantan’s 2015 fires [82]—lock in permanent fire vulnerability, disrupting local hydrology and triggering agricultural losses exceeding USD 4.8 billion [17].

7.3. Local Level (The Cause of Peatland Fires and Carbon Emissions)

Having established these thresholds, we now quantify their propagated impacts across Earth’s subsystems. At the local level, human-induced peatland degradation initiates self-reinforcing positive feedback loops that drive ecological collapse (Figure 5).
  • Agricultural expansion and drainage
    • Drainage–desiccation cycle:
      »
      Drainage lowers water tables by 1–3 m, drying peat soils and increasing flammability.
      »
      Example: The 2015 peat fires emitted 1.75 Gt CO2eq, worsening El Niño-driven droughts and prolonging fire seasons [68].
    • Subterranean fires:
      »
      Smoldering peat fires in Indonesia burn at an average depth of 0.51 ± 0.05 m [63], often evading detection by satellites. Current remote sensing technologies, including Landsat, lack reliable methods to identify these subsurface fires, as smoldering emits minimal thermal signals detectable by orbiting sensors [83].
    • Weak governance:
      »
      Only around 10% compliance with restoration laws, such as Law No. 32/200, sustains illegal land conversion and peat collapse (2–5 cm/year subsidence) [84]. Here, economic incentives such as palm oil production and weak policy enforcement such as lax sanctions under Law No. 41/1999, drive peatland degradation. For instance, the 2023 fires burned 1.16 million ha due to El Niño and agricultural expansion. In addition, the removal of the 3 m peat exploitation limit in 2021 prioritized short-term economic gain. This effectively endangered 70% of non-dome peatlands, exposing them to further exploitation. Meanwhile, communities lacking alternatives resort to slash-and-burn farming, signifying how socio-economic inequalities gaps compound natural hazards and amplify environmental degradation.
      »
      Threshold Effects: Peatlands within the Mega Rice Project in Central Kalimantan have lost their capacity to absorb and retain rainwater, maintaining drier-than-natural conditions year-round. This irreversible hydrological dysfunction heightens fire vulnerability, as seen in recurrent severe peat fires [85].
  • Ecological and socio-economic impacts
    • Local scale
      »
      Hydrological disruption:
      .
      Peat fires and prolonged drought destabilize agriculture.
      .
      Toxic haze and environmental health:
      »
      Toxic haze: Peat fires emit dense smoke laden with PM2.5, carbon monoxide, nitrogen dioxide, and mercury, blanketing Indonesia and causing respiratory illnesses as reflected by more than 900,000 cases in 2019. Another case in point is in Palangka Raya City (Central Kalimantan), where PM10 concentrations during the 2015 fires exceeded 2500 μg/m3—over 12 times the WHO’s 24 h guideline (200 μg/m3)—compared to <2000 μg/m3 in 2002 and 2006. This extreme pollution directly correlated with peat loss from fires in the Kalampangan area, exacerbating public health crises [13].
      »
      Marine collapse:
      • Polluted runoff containing oxides of carbon and nitrogen, ammonia, and polynuclear aromatic hydrocarbons elevates coral bleaching risks by increasing zooxanthellae density in coral tissue. These pollutants, combined with stressors like ocean acidification, raise mortality rates for marine organisms and devastate fisheries [86].
      »
      Economic losses:
      .
      The 2015 peatland fires in Indonesia caused USD 16 billion in economic losses [17] surpassing the USD 8 billion value of the country’s palm oil exports in 2014 [31]. This loss dwarfed the USD 4.5 billion cost of the 1997 fires, with both crises driven by severe disruptions to agriculture, transportation, and healthcare systems [17].
      »
      Environmental and health impacts:
      • Severe droughts lower the water table, exposing more peat and increasing susceptibility to burning.
      .
      The 2015 fires destroyed 2.6 million hectares of peatlands and forests in Sumatra and Kalimantan, leading to habitat destruction and biodiversity loss, including 346,039 hectares of agricultural land [17].
      .
      The resulting toxic haze triggered public health emergencies and long-term environmental degradation, disproportionately affecting rural communities. For example, villages like Tumbang Nusa and Tanjung Taruna in Central Kalimantan endured annual smog exposure for up to three months during fire seasons from 1997 to 2015, crippling livelihoods tied to fishing and non-timber forest product harvesting [87].
      .
      Cultural erosion:
      • Indigenous communities, such as the Dayak and Banjar tribes in Kalimantan, faced cultural erosion as fires destroyed sacred forests and displaced ancestral lands. Traditional practices—including ritual ceremonies, subsistence farming, and medicinal plant harvesting—were disrupted, severing ties to cultural heritage [88].
      .
      Systemic poverty:
      • The destruction of agricultural land and health crises from haze pollution reduced productivity and deepened poverty. Smallholder farmers, already vulnerable, lost crops and income, while healthcare costs strained household budgets. The World Bank [17] emphasized that these compounding effects entrenched systemic poverty in fire-affected regions.

7.4. Global Level: Cascading Effects of Peatland Disintegrations

While the above impacts originate locally, they amplify into planetary crises through the Earth’s interconnected systems. To contextualize the global cascading effects of Indonesia’s peatland degradation, we first established local feedback loops as catalysts for planetary risks. Systems thinking revealed how seemingly localized disruptions—like drainage-induced fires or governance failures—propagated through the Earth’s subsystems (atmosphere, hydrosphere, lithosphere, biosphere, cryosphere), driving global warming, biodiversity loss, and socio-economic inequities. This foundation ensures solutions target root causes, not symptoms.
Indonesia’s peatlands, vital carbon sinks and biodiversity hubs, face severe degradation from drainage, fires, and weak governance. This disintegration, as discussed below, triggers cascading impacts across the Earth’s subsystems—atmosphere, hydrosphere, lithosphere, biosphere, and cryosphere—threatening global climate stability and sustainable development.

7.5. Global Level: Tropical Peatland Disintegrations and Threats to Global Sustainability

Peatland degradation undermines the Sustainable Development Goals (SDGs) and Paris Agreement via tropical peatland disintegration and global environmental change:
»
SDG 1 (no poverty): Livelihood losses in agriculture and fisheries exacerbate rural poverty.
»
SDG 13 (climate action): Peat fires emit 884 Mt CO2/year, rivaling Germany’s fossil fuel emissions.
»
SDG 14–15 (life below water/on land): Coastal flooding and habitat loss threaten more than 300 species including those shown in Table 3.

7.6. Tropical Peatland Disintegration and Environmental Change

Having established the drainage–subsidence–governance (DSG) nexus as the core degradation engine, we now analyze its cascading impacts across the Earth’s subsystems—beginning with local manifestations before scaling to global consequences. First, we quantify immediate biophysical and social disruptions across Indonesian peat landscapes; subsequently, we trace their propagation through atmospheric, hydrospheric, and cryospheric systems.
Peatlands function as critical climate regulators, storing immense carbon stocks over millennia yet capable of rapid carbon release when disturbed. This section examines the self-amplifying feedback loops that transform localized peatland degradation into global biogeochemical crises. At the core lies a hydrological pivot: drainage triggers peat drying, enabling fires that liberate lithospheric carbon as atmospheric CO2. These emissions induce radiative forcing, accelerating climate extremes that further desiccate peatlands—establishing a planetary-scale feedback circuit connecting land-use decisions to the Earth system destabilization. We dissect these cascading interactions across subsystems, quantifying how peatland drainage ignites a chain reaction with disproportionate climate consequences.
With reference to Figure 5, drainage-driven desiccation and weak governance enforcement synergistically ignite peatland degradation cycles—triggering cross-subsystem feedback loops that originate locally before triggering global consequences. We first isolate these local mechanisms, examining how
»
Drainage collapses water tables (<40 cm) to destabilize peat hydrology;
»
Governance voids enable illegal conversion, accelerating fire risk and biodiversity loss;
»
Community health and livelihoods bear immediate costs.
This local lens establishes the foundation for the subsequent Earth-subsystem analysis
Local drivers → global consequences
Drainage → fires → CO2 emissions:
Draining peatlands lowers water tables, igniting fires that release 1.75 Gt CO2eq (2015), equivalent to Germany’s annual emissions [68].
  • Weak governance → peat collapse:
    Illegal land conversion causes subsidence (2–5 cm/year), releasing ancient carbon stocks (57 GtC) and accelerating sea-level rise [63].
Earth’s subsystem impacts
Atmosphere
  • PM2.5 pollution:
    »
    Mechanism: 2015 fires emitted 7.33 Tg PM2.5 (more than 3.5 times higher than non-peat estimates) [10].
    »
    Impact: More than 100,000 premature deaths in Southeast Asia; USD 9B/year regional economic losses [89].
  • Methane emissions:
    »
    Mechanism: Peat decomposition released 226.2 Mt CO2e/year by 2012 [90].
    »
    Impact: Methane, which has a global warming potential 28 times greater than that of carbon dioxide over a 100-year timescale accelerates global warming [91].
Hydrosphere—disrupted hydrology:
»
Mechanism:
»
In Jambi Province, peat fires and prolonged drought during the 2019 fires reduced rainfall from 40 mm to nearly 0 mm in fire hotspots, severely disrupting local water cycles [92].
»
During the 2015 peat fires, daily rainfall plummeted to 1.0 mm (compared to the normal average of 3.9 mm), with the dry season extending to 150 days (May–October). This extreme aridity intensified fire persistence and peatland degradation [93].
Hydrological impacts on the hydrosphere
The drastic rainfall reduction caused the following:
»
Soil moisture depletion: Drying of peat soils, reducing groundwater recharge and lowering water tables.
»
Peatland degradation: Loss of water retention capacity in peat ecosystems, accelerating irreversible drainage.
»
Surface water scarcity: Rivers and reservoirs dried up, disrupting freshwater availability for agriculture and communities.
»
Increased drought feedback: Drier conditions amplified fire risk, creating a self-reinforcing cycle of aridity.
Lithosphere—peat subsidence:
»
Mechanism:
Peat subsidence occurs when drained peatlands lose water, causing the organic-rich soil to compact and oxidize. This process
»
Reduces peat volume as peat soils shrink by 2–5 cm/year due to water loss and microbial decomposition;
»
Alters soil structure as oxidation breaks down organic matter, collapsing pore spaces and permanently lowering land elevation.
Lithospheric impacts
  • Permanent land loss:
    »
    Subsidence causes the irreversible lowering of the land surface, reducing the lithosphere’s capacity to support ecosystems or infrastructure.
    »
    Example: In Indonesia, subsidence has rendered 1.2 million hectares of peatland unusable for agriculture or habitation [94].
    »
    Salination: Sinking land allows saltwater intrusion into freshwater aquifers, contaminating groundwater (a critical lithosphere–reservoir interaction).
  • Carbon loss
    »
    When drained, peat soils undergo oxidation, emitting 1.9 ± 0.3 Gt CO2/year [64]—comparable to 2–3% of global fossil fuel emissions. This shifts peatlands from long-term carbon sinks to net emission sources, degrading their climate mitigation capacity [95].
Subsidence exacerbates drainage issues, accelerating further peat degradation and land collapse—a self-reinforcing cycle.
Permafrost thaw in situations where there is no direct physical interaction
It is noteworthy that despite lacking direct physical interaction, tropical/boreal peatlands and Arctic permafrost systems are critically interconnected through global radiative forcing. Peatland degradation releases contemporary carbon stocks as CO2/CH4, amplifying atmospheric warming. This warming disproportionately accelerates in polar regions (Arctic amplification), triggering permafrost thaw that liberates ancient cryospheric carbon reserves, and thawed organic matter undergoes microbial decomposition, emitting additional CO2 and CH4, thereby establishing a self-reinforcing feedback loop: peat emissions → warming → thaw → permafrost emissions → enhanced warming. This cross-system coupling transforms locally contained carbon pools into global climate liabilities, demonstrating how geographically disjointed ecosystems interact via atmospheric teleconnections.
Permafrost thaw initiates a self-reinforcing vicious cycle driven by three interconnected mechanisms: (a) albedo reduction, (b) thermokarst formation, and (c) microbial activation.
Biosphere
Coral mortality mechanism:
»
Wildfire ash (iron-rich) washes into coastal waters, triggering harmful algal blooms (red tides). These blooms block sunlight and deplete oxygen, causing 100% mortality in coral reefs, for example, in Mentawai Islands, Indonesia [96].
Biospheric impacts
  • Collapse of marine ecosystems:
    »
    Loss of coral reefs disrupts habitat for 76% of Coral Triangle species, including fish, crustaceans, and symbiotic organisms [97].
    »
    Example: Declines in reef fish populations destabilize food webs, affecting predators like sharks and rays [95].
    »
    Fisheries Collapse: Coastal communities lose livelihoods as fish stocks decline by 40–60% in fire-affected regions [95].
  • Biodiversity loss:
    »
    Mechanism: Peat fires and deforestation fragment habitats, isolating populations of endangered species like orangutans and disrupting ecological corridors for over 300 species including Sumatran tigers and clouded leopards. Accelerating biodiversity loss disproportionately threatens keystone species—organisms with outsized ecological roles [98]. Their decline triggers cascading ecosystem dysfunction: for instance, vanishing orangutans (Pongo pygmaeus) disrupt seed dispersal for >60% of dipterocarp trees (Dipterocarpaceae), critically impairing forest regeneration in degraded peatlands.
  • Zoonotic disease risks:
    »
    Habitat loss forces wildlife into human settlements, increasing contact rates. This elevates risks of disease spillover, for example, the Nipah virus [99,100]. Critically, permafrost thaw may resurrect ancient viruses; viable Salmonella DNA and functional viral genes have been recovered from 8-million-year-old glacial ice [101,102], suggesting climate change could unlock prehistoric pandemics alongside contemporary spillover events.
Cryosphere
Mechanism:
Peat-driven warming:
»
Tropical peat fires (e.g., Indonesia, 2015) emit 0.8–1.5 GtC/year of CO2, amplifying global warming. Arctic regions warm 2–4 times faster than the global average due to polar amplification, accelerating permafrost thaw [98].
»
Peatland emissions and permafrost thaw: Indirect link via global warming peatland emissions (primarily CO2 and methane from tropical and boreal peat fires/degradation) contribute to global greenhouse gas (GHG) concentrations. This global warming indirectly accelerates Arctic permafrost thaw through Arctic amplification (the Arctic warms 2–4 times faster than the global average).
Evidently, peatlands function as planetary tipping elements—where localized drainage triggers self-amplifying feedback loops (drainage → drying → fire → CO2 → climate extremes) that cascade across the Earth’s subsystems. The data demonstrate that these degraded landscapes shift from carbon sinks to cascading emission sources. This disrupts the Earth’s interconnected systems through multiple feedback loops. In summary, peat fires and drainage release vast CO2 emissions, amplifying global warming and accelerating Arctic permafrost thaw, which further destabilizes the climate. Hydrological collapse from dried peatlands reduces rainfall, exacerbates droughts, and acidifies oceans, driving coral bleaching and fisheries collapse.
Peatland degradation drives coastal subsidence, biodiversity collapse, and toxic haze—direct consequences of weak governance and unsustainable land use. Without intervention, Indonesia’s GHG emissions are projected to reach 2.869 GtCO2e by 2030 (NDC’s business-as-usual scenario). However, targeted action could reduce emissions by 0.834 GtCO2e (29%) through domestic efforts or 1.081 GtCO2e (41%) with international support, offering a critical pathway to avert climate and ecological crises. Peatland-specific projections estimate 2030 emissions at 0.569 GtCO2e (0.510 GtCO2e without El Niño; 0.588 GtCO2e with El Niño), comprising land-cover change (0.025 GtCO2e), decomposition (0.228–0.278 GtCO2e), canals (0.021–0.024 GtCO2e), and fires (0.110 GtCO2e), with an additional 0.0108 GtCO2e annually from degradation [8].

8. Policy Implications

Given the irreversible planetary costs of disrupted peatland feedback loops—where local drainage cascades into global radiative forcing—the evidence compels urgent policy intervention. The cascading impacts of peatland degradation as elucidated above imperatively demand a paradigm shift from reactive mitigation to precautionary governance prioritizing indefinite peatland protection. This framework, anchored in strict drainage bans (<40 cm water table depth, per Indonesian PP 57/2016) and advanced fire-suppression technologies, is structured as detailed in the following.
Core Precautionary Governance Framework
  • Indefinite peatland protection
    • Non-negotiable safeguards: Prioritize perpetual protection over reactive mitigation, recognizing peatlands’ millennial-scale carbon accumulation and irreversibility of degradation.
    • Strict hydrological thresholds: Legally enforce water table depths ≤40 cm (PP 57/2016) to prevent oxidation and subsidence (2–5 cm/year). IoT sensors and automated fines ensure compliance.
  • Minimum 50-year protection baseline
    • Stabilization phase: A 50-year protection period is critical to stabilize emissions by securing Indonesia’s immense carbon stock (54.6 Gt C), equivalent to 6 years of global fossil fuel emissions. This stock remains stable only under hydrated conditions.
    • Rewetting imperative: Restoring the peatland hydrology is the only way to prevent peat oxidation and mitigate CO2 emissions. Peatland rewetting could deliver up to 13% of Indonesia’s emissions reductions from natural climate solutions (NCS) by suppressing CO2 from peat oxidation [99]. At the project scale, rewetting 590 km2 of drained peat swamp forest mitigates 1.2–1.5 Mt CO2-eq annually—consistent with empirical measurements of 17–39% emission reductions across land uses [100].
  • Adaptive milestones for long-term resilience
    • Short-term (10–20 years): Immediate drainage cessation, IoT-driven enforcement of water tables, and fire-suppression tech deployment such as LIDAR for early detection.
    • Mid-term (30–50 years): Biodiversity recovery (e.g., orangutan habitat restoration, endemic fish species like Scleropagus formosus).
    • After 50 years: Transition to permanent protection contingent on ecological thresholds (e.g., sustained water tables ≤40 cm, peat depth stability).
  • Global Climate Stabilization Zones (preemptive threshold defense)
    • Governance must preempt thresholds (e.g., enforcing ≤40 cm water tables via IoT sensors) to avoid emergent risks like Arctic methane releases. Our proposed Global Climate Stabilization Zones (>3 m peat domes) directly safeguard these tipping points, while ASEAN transboundary enforcement disrupts cross-haze feedback loops. Government must legally enshrine core peat domes (>3 m depth) as non-negotiable “Global Climate Stabilization Zones”, prohibiting commercial activity to safeguard their carbon stocks (57 GtC nationally). This entails the following:
      »
      Recognizing peatlands as critical planetary infrastructure under national and international law.
      »
      Binding hydrological safeguards: Enforce water table depth limits (≤40 cm) and subsidence thresholds (≤1 cm/year) via IoT sensors (piezometers, soil moisture detectors) that trigger automated fines for violations.
      »
      Total protection for deep peat: Designate peat domes deeper than 3 m as permanent “No-Go Zones,” prohibiting commercial activity.
      »
      Corporate criminal liability: Legally require “Peatland-System Due Diligence,” holding corporate boards liable for repeated violations. Executives face jail time, while concession holders post “peat bonds” such as USD 10,000/ha forfeited for breaches.
This Core Precautionary Governance Framework is an integrated system designed to address the irreversible risks of peatland degradation through proactive, science-driven safeguards. Its components are interlinked to balance immediate action with long-term ecological stability, ensuring peatlands function as permanent climate infrastructure. Below is how its elements connect and reinforce one another.
Supporting strategies:
  • Precision monitoring and enforcement
    • IoT networks: Real-time monitoring of water tables, methane/CO2 emissions, and subsidence via piezometers and soil moisture sensors.
    • Fire-suppression technologies: Deploy drones and Sentinel-1 radar to detect subsurface fires, closing the 60% “emission gap” in current detection systems.
  • Equity-centered livelihood transitions
    • Paludiculture: Replace drainage-dependent crops with wetland-compatible alternatives (sago, jelutong) to align livelihoods with peat hydrology.
    • Indigenous stewardship: Formalize Dayak and Kubu land rights, integrating traditional fire management, for example, tembawang agroforestry, into national strategies.
Remarks: Equity-centered restoration bridges ecological and human well-being, ensuring advanced technologies like IoT and drones complement—not replace—Indigenous stewardship.
Regional accountability mechanisms:
Local actions must align with regional and international frameworks to address the transboundary nature of peatland degradation:
  • ASEAN Haze Agreements institutionalizing transboundary enforcement
    »
    Joint firefighting protocols: Mandate shared firefighting resources, for example, aerial water bombers, regional response teams, across ASEAN member states during haze crises.
    »
    IoT-driven peat health monitoring: Establish a unified regional platform for real-time data sharing on water tables, subsidence, and fire risks, leveraging IoT sensors and Sentinel-1 radar.
    »
    Standardized penalties: Harmonize fines for cross-border violations, for example, Malaysian/Singaporean firms draining Indonesian peatlands).
  • ASEAN Haze Mitigation Fund:
    »
    Palm oil levy: Impose a 1–3% levy on regional palm oil exports to generate USD 200–500 million annually, funding the following:
    »
    Peatland rewetting and canal blocking.
    »
    Community-based fire prevention programs.
    »
    Technology upgrades (drones, LIDAR) for early fire detection.
    »
    Transparency mechanisms to publicly audit fund allocation.
This regional approach can transform Indonesia’s peatland governance into a local and regional preventive mechanism, directly advancing
  • SDG 13 (climate action) by cutting transboundary haze emissions (1.75 Gt CO2eq in 2015) and stabilizing regional climate patterns;
  • SDG 15 (life on land) by protecting biodiversity hotspots, for example, orangutans and Scleropagus Formosus in Borneo peatlands.
Indonesia’s peatlands are not merely ecosystems—they are planetary life-support systems, safeguarding humanity from climate chaos. The choice is binary:
  • Embrace precaution: Legally enshrine peatlands as Global Climate Stabilization Zones, enforcing indefinite protection through IoT-driven accountability, ASEAN transboundary governance, and adaptive milestones (50-year stabilization → perpetual stewardship). This secures the irreplaceable 57 GtC stockpile, stabilizes regional hydrology, and upholds SDGs 13–15.
  • Risk collapse: Prioritize short-term profit, triggering cascading feedback loops—2.5 Gt CO2eq/year emissions, Arctic permafrost thaw, biodiversity extinction, and haze-induced health crises—that no mitigation can undo.
Indonesia’s FOLU Net Sink 2030 targets hinge on protecting its tropical peatlands across Kalimantan, Sumatra, and Papua—regions that collectively store over 50% of the nation’s peat carbon [34,35,101]. While Kalimantan alone accounts for 30% of Indonesia’s peatland areas, Sumatra and Papua face parallel threats from drainage and fires, necessitating integrated conservation efforts. Achieving the sectoral goal of −140 Mt CO2e by 2030 requires strict adherence to Ministerial Decree SK.129/2017, which mandates IoT-enforced water tables (≤40 cm) and fire bans to mitigate emissions [20]. Without these measures, recurrent fires—such as the 1.16 million ha burned in 2023—will derail progress toward −304 Mt CO2e by 2050, jeopardizing national and global climate commitments [34].
Indonesia must lead by example, demonstrating that peatland stewardship is both ecologically vital and economically viable. By anchoring governance in indefinite safeguards such as PP 57/2016 and regional collaboration, for example, ASEAN haze agreements, Indonesia can inspire global efforts to protect tropical peatlands—from the Amazon to the Congo Basin. The science is unequivocal; the tools exist. What remains is the political will to act.

9. Conclusions

Indonesia’s tropical peatlands, which store some 57 GtC globally, are indispensable yet undervalued climate infrastructure. This study confirms that their degradation—driven by drainage, fires, and governance failures—has transitioned these millennia-old carbon sinks into catastrophic emission sources, with cascading consequences for the Earth’s systems. Three interconnected feedback loops, identified through a systems-based analysis, underpin this crisis:
  • Drainage–desiccation–oxidation: Lowering water tables beyond the −40 cm threshold triggers irreversible peat oxidation (2–5 cm/year subsidence) and fires, risking 4.3 GtCO2eq/year emissions by 2050—a return to the 1997-level catastrophe.
  • Fire–climate–permafrost: Emissions from peat combustion (18 Mg CO2eq/ha/cm burned) intensify radiative forcing, accelerating Arctic permafrost thaw (+15% since 2000) and disrupting monsoons.
  • Economic–governance failure: Palm oil’s economic dominance (4.5% of Indonesia’s GDP) perpetuates drainage, while weak enforcement (30+ annual violators) locks in degradation.
The proposed precautionary governance framework addresses these feedback loops:
  • IoT-enforced water tables (≤40 cm depth) reduce emissions by 34%, aligning with Indonesia’s FOLU Net Sink 2030 target (−140 Mt CO2e).
  • Global Climate Stabilization Zones protect peat domes (>3 m depth, storing 57 GtC) as non-negotiable reserves.
  • ASEAN transboundary enforcement, funded by a 1–3% palm oil levy, integrates blockchain-tracked deforestation bans and fire suppression.
This research achieved its dual aims by
  • Pioneering a systems-based causality model that quantifies feedback loops such as drainage–desiccation–oxidation, validated against historical thresholds:
    • Irreversible peat drying below −40 cm water tables;
    • Fire vulnerability at <4 mm/day rainfall;
    • Carbon source conversion across a <3 m peat depth.
  • Exposing planetary cascades where local degradation triggers
    • Atmospheric crises (1.75 Gt CO2eq emissions rivaling Germany’s fossil output);
    • Cryospheric destabilization (Arctic thaw +15% since 2000);
    • Biosphere collapse (50% Bornean orangutan decline).
  • Bridging policy–science gaps via SDG-aligned strategies that address enforcement failures (<10% compliance) through IoT monitoring and paludiculture transitions.
    Policy Imperative:
    To avert irreversible Earth system destabilization, these findings demand preemptive governance of critical thresholds—enforcing ≤−40 cm water tables via IoT sensors and legally safeguarding >3 m peat domes as Global Climate Stabilization Zones. This must be scaled through ASEAN transboundary enforcement, funded by a 1–3% palm oil levy, to disrupt cross-haze feedback loops and position peatlands as non-negotiable planetary stabilizers for the Paris Agreement and SDGs.
    Theoretical recommendations:
    • Political ecology and institutional theory
      »
      Conceptualize how transboundary institutions, for example, ASEAN’s haze agreements, overcome scalar mismatches in environmental governance.
    • Earth system science
      »
      Define “Anthropogenic Tipping Cascades”:
      Theorize peatlands as initiators in global tipping point sequences (e.g., Indonesian drainage → permafrost thaw → AMOC collapse).
      »
      Integrate peatlands into planetary boundary metrics:
      Propose a revised “Interference with Biogeochemical Cycles” boundary incorporating peat carbon vulnerability.
    Challenges and forward pathways:
While regional case studies such as those in Kalimantan validate the model, scalability across Indonesia’s diverse peat landscapes and other tropical regions (Amazon, Congo) requires high-resolution monitoring (e.g., LIDAR, IoT) to address subsurface fire detection gaps. Equitable implementation must also prioritize Indigenous stewardship and paludiculture transitions to reconcile ecological and human needs.
Without systemic action, annual emissions will surpass 2.5 GtCO2/year (40% above Indonesia’s Paris budget), exacerbating transboundary haze (100,000+ deaths/year) and biodiversity collapse. By enshrining peatlands as Global Climate Stabilization Zones—enforcing IoT safeguards, corporate liability (USD 10,000/ha bonds), and ASEAN collaboration—Indonesia can avert the irreversible destabilization of the Earth system. This framework, rooted in science and policy innovation, positions peatlands as linchpins of global climate resilience, demanding immediate integration into SDG and Paris Agreement agendas.
Limitations and future directions:
  • The causal loop model, while holistic, relies on regional case studies (Kalimantan), necessitating expanded validation across Indonesia’s diverse peat landscapes.
  • Subsurface fire detection gaps and equity blind spots such as Indigenous land tenure require IoT networks and participatory research.
  • The current causal loop model prioritizes conceptual clarity over numerical prediction in association with complex socio-ecological and Earth’s system interactions. Future work will implement quantitative system dynamics simulations using Vensim to
    • Calibrate stock-flow equations, for example carbon stocks and water table depth against historical data;
    • Conduct sensitivity analyses on policy levers, for example, IoT compliance rates;
    • Project long-term scenarios (2050) under business-as-usual vs. governance intervention.
Indonesia’s peatlands are planetary stabilizers, not expendable resources. By enshrining them as Global Climate Stabilization Zones and through Indigenous stewardship, the nation can align its FOLU Net Sink 2030 (−140 Mt CO2e) with global climate targets. Failure risks locking in 1997-level emissions, biodiversity collapse, and the irreversible destabilization of the Earth system.

Author Contributions

Conceptualization, Y.K.C. and A.O.; methodology, Y.K.C. and A.O. validation, Y.K.C. and A.O.; formal analysis, Y.K.C. investigation, Y.K.C. and A.O.; resources, A.O. and Y.K.C.; writing—original draft preparation, Y.K.C.; writing—review and editing, Y.K.C. and A.O.; supervision, A.O. All authors have read and agreed to the published version of the manuscript.

Funding

This research received no external funding.

Data Availability Statement

Publicly accessible.

Conflicts of Interest

The authors declare no conflicts of interest.

References

  1. 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]
  2. Warren, M.; Hergoualc’h, K.; Kauffman, J.B.; Murdiyarso, D.; Kolka, R. An appraisal of Indonesia’s immense peat carbon stock using national peatland maps: Uncertainties and potential losses from conversion. Carbon Balance Manag. 2017, 12, 12. [Google Scholar] [CrossRef]
  3. Asyhari, A.; Gangga, A.; Septiadi, P.C.A.; Ritonga, R.A.; Candra, R.A.; Anshari, G.Z.; Bowen, J.C.; Perryman, C.R.; Novita, N. Quantifying the fluxes of carbon loss from an undrained tropical peatland ecosystem in Indonesia. Sci. Rep. 2024, 14, 11459. [Google Scholar] [CrossRef]
  4. Leifeld, J.; Menichetti, L. The underappreciated potential of peatlands in global climate change mitigation strategies. Nat. Commun. 2018, 9, 1071. [Google Scholar] [CrossRef] [PubMed]
  5. Evans, M. Progress and Perils in Indonesia’s Peatland Restoration Journey; Center for International Forestry Research (CIFOR): Nairobi, Kenya; World Agroforestry (ICRAF): Bogor Barat, Indonesia, 2020. [Google Scholar]
  6. Choy, Y.K. Global Environmental Sustainability: Case Studies and Analysis of the United Nations’ JourToward Sustainable Development; Elsevier: Amsterdam, The Netherlands; Oxford, UK; Cambridge, MA, USA, 2020. [Google Scholar]
  7. BAPPENAS. Reducing Carbon Emissions from Indonesia’s Peat Lands Interim Report of a Multi-Disciplinary Study December 2009; Indonesian National Development Planning Agency (BAPPENAS): Jakarta, Indonesia, 2009. [Google Scholar]
  8. Murdiyarso, D.; Anshari, G.Z.; Manuri, S.; Wijaya, A.; Darmawan, A.; Budiman, A.; Purwanto, J.; Karim, A.; Rinuwat, T.; Kustiyo; et al. Reference Level for the Estimation of Emission Reduction from Peatland Restoration; Badan Restorasi Gambut (BRG): Jakarta, Indonesia, 2018. [Google Scholar]
  9. Fielda, R.D.; van der Werfc, G.R.; Faninc, T.; Eric, J.; Fetzerd, E.J.; Fuller, R.; Jethvae, H.; Levye, R.; Liveseyd, N.J.; Luo, M.; et al. Indonesian fire activity and smoke pollution in 2015 show persistent nonlinear sensitivity to El Niño-induced drought. Proc. Natl. Acad. Sci. USA 2016, 113, 9204–9209. [Google Scholar] [CrossRef]
  10. Kiely, L.; Spracklen, D.V.; Wiedinmyer, C.; Conibear, L.; Reddington, C.L.; Archer-Nicholls, S.; Lowe, D.; Arnold, S.R.; Knote, C.; Khan, M.F.; et al. New estimate of particulate emissions from Indonesian peat fires in 2015. Atmos. Chem. Phys. 2019, 19, 11105–11121. [Google Scholar] [CrossRef]
  11. 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]
  12. Rein, G.; Huan, X. Smouldering wildfires in peatlands, forests and the arctic: Challenges and perspectives. Curr. Opin. Environ. Sci. Health. 2021, 24, 100296. [Google Scholar] [CrossRef]
  13. Kusin, K.; Jagau, Y.; Ricardo, J.; Saman, T.N.; Aguswan, Y. Peat lost by fire in Kalampangan area, Central Kalimantan, Indonesia. IOP Conf. Ser. Earth Environ. Sci. 2020, 504, 012009. [Google Scholar] [CrossRef]
  14. Hirano, T.; Ohkubo, S.; Itoh, M.; Tsuzuki, H.; Sakabe, A.; Takahashi, H.; Kusin, K.; Osaki, M. Large variation in carbon dioxide emissions from tropical peat swamp forests due to disturbances. Commun. Earth Environ. 2024, 5, 221. [Google Scholar] [CrossRef]
  15. van der Werf, G.R.; Randerson, J.T.; Giglio, L.; van Leeuwen, T.T.; Chen, Y.; Rogers, B.M.; Kasibhatla, P.S. Global fire emissions estimates during 1997–2016. Earth Syst. Sci. 2017, 9, 697–720. [Google Scholar] [CrossRef]
  16. Choy, Y.K.; Onuma, A. Transboundary Ecological Conservation, Environmental Value, and Environmental Sustainability: Lessons from the Heart of Borneo. Sustainability 2021, 13, 9727. [Google Scholar] [CrossRef]
  17. World Bank. The Cost of Fire: An Economic Analysis of Indonesia’s 2015 Fire Crisis; World Bank Group: Washington, DC, USA, 2018; Available online: https://documents.worldbank.org (accessed on 30 May 2025).
  18. Qiu, C.J.; Zhu, D.; Ciais, P.; Guenet, B.; Krinner, G.; Peng, S.S.; Aurela, M.; Bernhofer, C.; Brümmer, C.; Bret-Harte, S.; et al. ORCHIDEE-PEAT(revision 4596), a model for northern peatland CO2, water, and energy fluxes on daily to annual scales. Geosci. Model Dev. 2018, 11, 497–519. [Google Scholar] [CrossRef]
  19. Newbold, T.; Tittensor, D.P.; Harfoot, M.B.J.; Scharlemann, J.P.W.; Purves, D.W. Non-linear changes in modelled terrestrial ecosystems subjected to perturbations. Sci. Rep. 2020, 10, 14051. [Google Scholar] [CrossRef]
  20. Hooijer, A.; Page, S.; Canadell, J.G.; Silvius, M.; Kwadijk, J.; Wösten, H.; Jauhiainen, J. Subsidence and carbon loss in drained tropical peatlands. Biogeosciences 2012, 9, 1053–1071. [Google Scholar] [CrossRef]
  21. Lam, S.K.; Goodrich, J.P.; Liang, X.; Zhang, Y.J.; Pan, B.; Schipper, L.A.; Sulaeman, Y.; Nelson, L.; Chen, D. Mitigating soil greenhouse-gas emissions from land-use change in tropical peatlands. Front. Ecol. Environ. 2022, 20, 352–360. [Google Scholar] [CrossRef]
  22. Hodgkins, S.B.; Richardson, C.J.; Dommain, R.; Wang, H.J.; Glaser, P.H.; Verbeke, B.; Winkler, B.R.; Cobb, A.R.; Rich, V.I.; Missilmani, M.; et al. Tropical peatland carbon storage linked to global latitudinal trends in peat recalcitrance. Nat. Commun. 2018, 9, 3640. [Google Scholar] [CrossRef] [PubMed]
  23. Hapsari, K.A.; Jennerjahn, T.C.; Nugroho, S.H.; Yulianto, E.; Behling, H. Sea level rise and climate change acting as interactive stressors on development and dynamics of tropical peatlands in coastal Sumatra and South Borneo since the Last Glacial Maximum. Glob. Change Biol. 2022, 28, 3459–3479. [Google Scholar] [CrossRef] [PubMed]
  24. Harenda, K.; Lamentowicz, M.; Samson, M.; Chojnicki, B.H. The Role of Peatlands and Their Carbon Storage Function in the Context of Climate Change. In Interdisciplinary Approaches for Sustainable Development Goals; Zielinski, T., Sagan, I., Surosz, W., Eds.; Springer International Publishing: Cham, Switzerland, 2018; pp. 169–187. [Google Scholar]
  25. UNOPS. Restoring Indonesian Peatland, Protecting our Planet; United Nations Office for Project Services (UNOPS): Copenhagen, Denmark, 2025; Available online: https://www.unops.org/news-and-stories/stories/restoring-indonesian-peatlands-protecting-our-planet (accessed on 30 May 2025).
  26. MoEF Decree of the Minister of Environment and Forestry Number. SK.129/MENLHK/SETJEN/PKL.0/2/2017 on Determination of Peat Hydrological Unit; Indonesian Ministry of Environment and Forestry (MoEF): Jakarta, Indonesia, 2017. [Google Scholar]
  27. Mishra, S.; Page, S.E.; Cobb, A.R.; Lee, J.S.H.; Jovani Sancho, A.J.; Sjögersten, S.; Jaya, A.; Aswandi; Wardle, D.A. Degradation of Southeast Asian tropical peatlands and integrated strategies for their better management and restoration. J. Appl. Ecol. 2021, 58, 1370–1387. [Google Scholar] [CrossRef]
  28. Sasmito, S.D.; Taillardat, P.; Adinugroho, W.C.; Krisnawati, H.; Novita, N.; Fatoyinbo, L.; Friess, D.A.; Page, S.E.; Lovelock, C.E.; Murdiyarso, D.; et al. Half of land use carbon emissions in Southeast Asia can be mitigated through peat swamp forest and mangrove conservation and restoration. Nat Commun. 2025, 16, 740. [Google Scholar] [CrossRef]
  29. Budiningsih, K.; Putera, P.B.; Nurlia, A.; Ulya, N.A.; Nurfatriani, F.; Salminah, M.; Yuniati, D.; Widarti, A. Peatland restoration research: A global overview with insights from Indonesia. J. Ecol. Environ. 2024, 48, 263–276. [Google Scholar] [CrossRef]
  30. Osaki, M.; Kato, T.; Dohong, A.; Putri, A.; Takahashi, H. Peatlogy in tropical peatland: A new transdisciplinary science. Res. Outreach. 2024. [Google Scholar] [CrossRef]
  31. UNDRR. Indonesia Wildfires, 2023—Forensic Analysis; United Nations Office for Disaster Risk Reduction (UNDRR): Geneva, Switzerland, 2024; Available online: https://www.undrr.org/resource/indonesia-wildfires-2023-forensic-analysis (accessed on 3 June 2025).
  32. KLHK. Rencana Strategis Tahun 2020–2024 (Revisi). Kementerian Lingkungan Hidup dan Kehutanan (KLHK), Jakarta, Indonesia, 2021 (Strategic Plan of the Ministry of Environment and Forestry for 2020–2024. Revised edition 2021. Available online: https://www.menlhk.go.id/cadmin/uploads/renstra_full_3_1_960074a1c4.pdf (accessed on 3 June 2025).
  33. Ministry of Environment and Forestry. FOLU NET SINK: Indonesia’s Climate Actions Towards 2030; Ministry of Environment and Forestry: Jakarta, Indonesia, 2023. [Google Scholar]
  34. MoEF. Indonesia’s Forests 2024. Towards Sustainability of Forest Ecosystems in Indonesia; Ministry of Environment and Forestry: Jakarta, Indonesia, 2024. [Google Scholar]
  35. Joosten, H.; Clarke, D. Wise Use of Mires and Peatlands–Background and Principles Including a Framework for Decision-Making; International Mire Conservation Group and International Peat Society: Jyväskylä, Finland, 2002. [Google Scholar]
  36. Rydin, H.; Jeglum, J.K. The Biology of Peatlands, 2nd ed.; Oxford University Press: Oxford, UK, 2013. [Google Scholar]
  37. IUCN. Peatlands and Climate Change; International Union for Conservation of Nature (IUCN): Gland, Switzerland, 2021. [Google Scholar]
  38. Bain, C.G.; Bonn, A.; Stoneman, R.; Chapman, S.; Coupar, A.; Evans, M.; Gearey, B.; Howat, M.; Joosten, H.; Keenleyside, C.; et al. UK Commission of Inquiry on Peatlands; IUCN UK Peatland Programme; IUCN: Edinburgh, UK, 2011. [Google Scholar]
  39. Abel, S.; Archibald, L.; Appulo, L.; Barthelmes, A.; Berghöfer, U.; Bruisch, K.; Büttner, M.; Davies, M.; Dewitz, I.; Elshehawi, S.; et al. Peatlands Atlas. Facts and Figures About Wet Climate Guardian; Bund für Umwelt und Naturschutz (BUND): Berlin, Germany, 2023. [Google Scholar]
  40. Nicholas, T.; Girkin, N.T.; Cooper, H.V.; Ledger, M.J.; O’Reilly, P.; Thornton, S.A.; Åkesson, C.M.; Cole, L.E.S.; Hapsari, K.A.; Hawthorne, D.; et al. Tropical peatlands in the Anthropocene: The present and the future. Anthropocene 2022, 40, 100354. [Google Scholar]
  41. Hu, Y.; Fernandez-Anez, N.; Smith, T.E.L.; Rein, G. Review of emissions from smouldering peat fires and their contribution to regional haze episodes. Int. J. Wildland Fire 2018, 27, 293–312. [Google Scholar] [CrossRef]
  42. Ramsar Convention Secretariat. Keep Peatlands Wet for a Better Future; The Ramsar Convention on Wetlands: Gland, Switzerland, 2015. [Google Scholar]
  43. Page, S.E.; Rieley, J.O.; Banks, C.J. Global and regional importance of the tropical peatland carbon pool. Glob. Change Biol. 2011, 17, 798–818. [Google Scholar] [CrossRef]
  44. Ribeiro, K.; Pacheco, F.S.; Ferreira, J.W.; de Sousa-Neto, E.R.; Hastie, A.; Krieger Filho, G.C.; Alvalá, P.C.; Forti, M.C.; Ometto, J.P. Tropical peatlands and their contribution to the global carbon cycle and climate change. Glob. Change Biol. 2021, 27, 489–505. [Google Scholar] [CrossRef]
  45. Tim Publikasi Katadata. Luas Gambut Indonesia Terbesar Kedua di Dunia Luas lahan gambut di Indonesia Mencapai 22, 5 Juta Hektare. Dkatadata, Jurnalisme Data, Indonesia. 2019. Available online: https://katadata.co.id/timpublikasikatadata/infografik/5e9a519433cb1/luas-gambut-indonesia-terbesar-kedua-di-dunia (accessed on 3 June 2025).
  46. UNDP. Indonesia’s Social Forestry Programme Supports Livelihoods and Climate Action; United Nations Development Programme (UNDP): New York, NY, USA, 2021. [Google Scholar]
  47. Suwito, D.; Suratman, S.; Poedjirahajoe, E. The Covid-19 pandemic impact on indigenous people livelihoods in the peat swamp forest ecosystem in Central Kalimantan Indonesia. IOP Conf. Ser. Earth Environ. Sci 2021, 894, 012023. [Google Scholar] [CrossRef]
  48. Ministry of the National Development Planning. Indonesian Biodiversity Strategy and Action Plan (IBSAP) 2015–2020; Ministry of the National Development Planning: Jakarta, Indonesia, 2016. [Google Scholar]
  49. Boyd, N.; Brousseau, J.; Collier, S.; Dow, R.; Dowds, N.; Dragiewicz, M.; Houlihan, P.; Jarrett, B.; Kulu, I.; Page, S.; et al. Biodiversity of the Sebangau tropical peat swamp forest, Indonesian Borneo. Mires Peat 2018, 22, 1–50. [Google Scholar] [CrossRef]
  50. IUCN. The IUCN Red List of Threatened Species, Version 2025-1; International Union for Conservation of Nature and Natural Resources (ICUN): Gland, Switzerland. Available online: https://www.iucnredlist.org/ (accessed on 28 May 2025).
  51. Thornton, S.A.; Dudin; Page, S.E.; Upton, C.; Harrison, M.E. Peatland fish of Sebangau, Borneo: Diversity, monitoring and conservation. Mires Peat 2018, 22, 1–25. [Google Scholar] [CrossRef]
  52. ASEAN Secretariat. Peatlands in Indonesia. ASEAN Haze Portal; The ASEAN Secretariat: Jakarta, Indonesia, 2021; Available online: https://hazeportal.asean.org/peatlands-in-sea/peatlands-in-indonesia/ (accessed on 17 May 2025).
  53. Wahyunto, W.; Dariah, A.; Agus, F. Distribution, Properties, and Carbon Stock of Indonesian Peatland. In Proceedings of the International Workshop on Evaluation and Sustainable Management of Soil Carbon Sequestration in Asian Countries, Bogor, Indonesia, 28–29 September 2010. [Google Scholar]
  54. UN-REDD Programme, Indonesia’s Climate Ambitions Advance with the Launch of Forest and Peatland Conservation and Sustainability Initiative. UN-REDD Programme. 2025. Available online: https://www.un-redd.org/post/press-release-indonesias-climate-ambitions-advance-launch-forest-and-peatland-conservation-and (accessed on 10 March 2025).
  55. Putra, E.I.; Hayasaka, H.; Takahashi, H.; Usup, A. Recent Peat Fire Activity in the Mega Rice Project Area, Central Kalimantan, Indonesia. J. Disaster Res. 2008, 3, 334–341. [Google Scholar] [CrossRef]
  56. Sumarga, E.; Hein, L.; Hooijer, A.; Vernimmen, R. Hydrological and economic effects of oil palm cultivation in Indonesian peatlands. Ecol. Soc. 2016, 21, 52. [Google Scholar] [CrossRef]
  57. Hein, L.; Sumarga, E.; Quiñones, M.; Suwarno, A. Effects of soil subsidence on plantation agriculture in Indonesian peatlands. Reg. Environ. Change 2022, 22, 121. [Google Scholar] [CrossRef]
  58. Dennis, R. A Review of Fire Projects in Indonesia (1982–1998); Center for International Forestry Research (CIFOR): Bogor, Indonesia, 1999. [Google Scholar]
  59. IAWF. Report from Indonesia: The Promise of Participation Amid the Politics of Peat Fires and Haze; International Association of Wildland Fire: Missoula, MT, USA, 2017. [Google Scholar]
  60. Nichol, J. Bioclimatic Impacts of the 1994 Smoke Haze in Southeast Asia. Atmos. Environ. 1997, 31, 1209–1219. [Google Scholar] [CrossRef]
  61. Fujiwara, M.; Kita, K.; Kawasaki, S.; Ogawa, T.; Komala, N.; Saraspriya, S.; Suripto, A. Tropospheric ozone enhancements during the Indonesian forest fire events in 1994 and in 1997 as revealed by ground-based observations. Geophys. Res. Lett. 1999, 26, 2417–2420. [Google Scholar] [CrossRef]
  62. Goldammer, J.G.; Anja, A.A. Fire Situation in Indonesia, in, Global Forest Fire Assessment 1990–2000; Food and Agriculture Organization of the United Nations (FAO): Rome, Italy, 2001; pp. 132–144. [Google Scholar]
  63. Page, S.E.; Siegert, F.; Rieley, J.O.; Boehm, H.-D.V.; Jayak, A.; Limink, S. The amount of carbon released from peat and forest fires in Indonesia during 1997. Nature 2002, 420, 61–65. [Google Scholar] [CrossRef] [PubMed]
  64. Saharjo, B.H. GHG Emissions’ Estimation from Peatland Fires in Indonesia—Review and Importance of Combustion Factor, in, Vegetation Fires and Pollution in Asia; Vadrevu, K.P., Ohara, T., Justice, C., Eds.; Springer: Cham, Switzerland, 2023; pp. 433–445. [Google Scholar]
  65. Dawud, Y. Smoke episodes and assessment of health impacts related to haze from forest fires: Indonesian experience. In Proceedings of the Health Guidelines for Vegetation Fire Events, Lima, Peru, 6–9 October 1998; Background Papers; World Health Organization (WHO): Geneva, Switzerland, 1999; pp. 313–333. [Google Scholar]
  66. Barber, C.V.; Schweithelm, J. Trial by Fire. Forest Fires and Forest Policy in Indonesia’s Era of Crisis and Reform. Wahington D.C., United States: World Resource Institute in Collaboration with WWF-Indonesia, Jakarta, Indonesia and Telapak Indonesia Foundation; Bogor, Indonesia. 2000. Available online: https://www.wri.org/research/trial-fire (accessed on 3 June 2025).
  67. Ballhorn, U.; Siegert, F.; Mason, M.; Limin, S.; Limin, S. Derivation of Burn Scar Depths and Estimation of Carbon Emissions with LIDAR in Indonesian Peatlands. Proc. Natl. Acad. Sci. USA 2009, 106, 21213–21218. [Google Scholar] [CrossRef] [PubMed]
  68. Huijnen, V.; Wooster, M.J.; Kaiser, J.W.; Gaveau, D.L.A.; Flemming, J.; Parrington, M.; Inness, A.; Murdiyarso, D.; Main, B.; van Weele, M. Fire carbon emissions over maritime southeast Asia in 2015 largest since 1997. Sci. Rep 2016, 6, 886. [Google Scholar] [CrossRef]
  69. Koplitz, S.N.; Mickley, L.J.; Marlier, M.E.; Buonocore, J.J.; Kim, P.S.; Liu, T.J.; Sulprizio, M.P.; DeFries, R.S.; Jacob, D.J.; Schwartz, J.; et al. Public health impacts of the severe haze in Equatorial Asia in September–October 2015: Demonstration of a new framework for informing fire management strategies to reduce downwind smoke exposure. Environ. Res. Lett. 2016, 11, 094023. [Google Scholar] [CrossRef]
  70. Harrison, M.K.; Capilla, B.R.; Ripoll, B.; Thornton, S.A.; Cattau, M.E.; Susan Page, S.E. Impacts of the 2015 fire season on peat-swamp forest biodiversity in Indonesian Borneo. In Proceedings of the 15th International Peat Congress 2016. Malaysia, 15-19 August 2016. [Google Scholar]
  71. MoEF. Managing Peatlands to Cope with Climate Change: Indonesia’s Experience; Ministry of Environment and Forestry (MoEF): Jakarta, Indonesia, 2018. [Google Scholar]
  72. ASEAN Learning Center. Area burned in 2019 forest fires in Indonesia Exceeds 2018. ASEAN Learning Center. Department of Local Administration, Dusit, Bangkok. Available online: https://asean.dla.go.th/public/news.do?cmd=news&category=1&nid=26424&lang=en&random=1583774366832#:~:text=That%20is%20more%20than%20the,showed%202.6%20million%20hectares%20burned (accessed on 30 May 2025).
  73. World Bank. Investing in People; The World Bank: Washington, DC, USA, 2019. [Google Scholar]
  74. Prasetyo, L.B.; Setiawan, Y.; Condro, A.A.; Kustiyo, K.; Putra, E.I.; Hayati, N.; Wijayanto, A.K.; Ramadhi, A.; Murdiyarso, D. Assessing Sumatran Peat Vulnerability to Fire under Various Condition of ENSO Phases Using Machine Learning Approaches. Forests 2022, 13, 828. [Google Scholar] [CrossRef]
  75. Mizuno, K.; Fujita, M.S.; Kozan, O.; Itoh, M.; Shiodera, S.; Naito, D.; Suzuki, H.; Gunawan, H. Introduction: The Vulnerability and Transformation of Indonesian Peatlands. In Vulnerability and Transformation of Indonesian Peatlands; Mizuno, K., Kozan, O., Gunawan, H., Eds. Springer: Berlin, Germany, 2023. [Google Scholar] [CrossRef]
  76. Laszlo, A.; Krippner, S. Systems Theories: Their Origins, Foundations, and Development. In Systems Theories and A Priori Aspects of Perception, Jordan, J.S., Ed.; Elsevier: Amsterdam, The Netherlands, 1997; pp. 47–74. [Google Scholar]
  77. Aslaksen, E.W. The System Concept and Its Application to Engineering; Springer: Berlin, Germay, 2013. [Google Scholar]
  78. Ford, D.N. A system dynamics glossary. Syst. Dyn. Rev. 2019, 35, 369–379. [Google Scholar] [CrossRef]
  79. Brown, C. Graph Algebra: Mathematical Modeling with a Systems Approach; Sage Publication: Los Angeles, CA, USA; London, UK; New Delhi, India; Singapore, 2008. [Google Scholar]
  80. Glazovsky, N. Environmental Structure and Function: Earth System. In Earth System: History and Natural Variability (Volume IV); Cilek, V., Ed.; Encyclopedia of Life Support Systems (EOLSS) Publisher: Oxford, UK, 2009; pp. 335–383. Available online: https://www.eolss.net/Sample-Chapters/C12/E1-01-08-07.pdf (accessed on 10 March 2025).
  81. Hooijer, A.; Silvius, M.; Wösten, H.; Page, S. PEAT-CO2, Assessment of CO2 emissions from drained peatlands in SE Asia; Delft Hydraulics Report Q3943; Delft Hydraulics: Rotterdamsewe, The Netherlands, 2006. [Google Scholar]
  82. Sulaiman, A.; Osaki, M.; Takahashi, H.; Yamanaka, M.D.; Susanto, R.D.; Shimada, S.; Kimura, K.; Hirano, T.; Wetadewi, R.I.; Sisva, S.; et al. Peatland groundwater level in the Indonesian maritime continent as an alert for El Niño and moderate positive Indian Ocean dipole events. Sci. Rep. 2023, 13, 939. [Google Scholar] [CrossRef] [PubMed]
  83. Elvidge, C.D.; Zhizhin, M.; Hsu, F.C.; Baugh, K.; Khomarudin, M.R.; Vetrita, Y.; Sofan, P.; Suwarsono; Hilman, D. Long-wave infrared identification of smoldering peat fires in Indonesia with nighttime Landsat data. Environ. Res. Lett. 2015, 10, 065002. [Google Scholar] [CrossRef]
  84. Zainal, A.W. Peatlands Restoration Policies in Indonesia: Success or Failure? IOP Conf. Ser. Earth Environ. Sci. 2022, 995, 012068. [Google Scholar]
  85. Putra, E.I.; Cochrane, M.A.; Vetrita, Y.; Graham, L.; Saharjo, B.H. Degraded Peatlands, Ground Water Level, and Severe Peat Fire Occurrences. In Proceedings of the 15th International Peat Congress 2016, Kuching, Sarawak, Malaysia, 15–19 August 2016. [Google Scholar]
  86. Jaafar, Z.; Loh, T.L. Linking land, air and sea: Potential impacts of biomass burning and the resultant haze on marine ecosystems of Southeast Asia. Glob. Change Biol. 2014, 20, 2701–2707. [Google Scholar] [CrossRef]
  87. Akbar, A.; Adriani, S.; Priyanto, E. The potential for peatland villages to prevent fire: Case study of Tumbang Nusa Village Central Kalimantan. IOP Conf. Ser. Earth Environ. Sci. 2021, 758, 012017. [Google Scholar] [CrossRef]
  88. Alcorn, J.B. An Introduction to the Linkages between Ecological Resilience and Governance. In Indigenous Social Movements and Ecological Resilience: Lessons from the Dayak of Indonesia; Alcorn, J.B., Royo, A.G., Eds.; Biodiversity Support Program: Washington, DC, USA, 2000; pp. 1–16. [Google Scholar]
  89. Datta, A.; Krishnamoorti, R. Understanding the Greenhouse Gas Impact of Deforestation Fires in Indonesia and Brazil in 2019 and 2020. Front. Clim. 2022, 4, 799632. [Google Scholar] [CrossRef]
  90. Masri, A.Y.; Rosehan, A.; Tosiani, A.; Rossita, A.; Darmawan, A.; Yusara, A.; Marthinus, D.; Riana, A.; Pratiwi, E.; Novitri, F.; et al. Indonesia Third Biennial Update Report Under the United Nations Framework Convention on Climate Change; Minister of Environment and Forestry: Jakarta, Indonesia, 2021. [Google Scholar]
  91. IPCC. Climate Change 2021: The Physical Science Basis. Contribution of Working Group I to the Sixth Assessment Report of the Intergovernmental Panel on Climate Change; Cambridge University Press: Cambridge, UK; New York, NY, USA, 2021. [Google Scholar] [CrossRef]
  92. Khalifah Insan Nur Rahmi, K.I.; Prasasti, I.; Nugroho, J.T.; Khomaruddin, M.R. Analyzing Rainfall and Greenness Vegetation Level onForest/Land Fire Area in Jambi and Central Kalimantan Provinces using Remote Sensing Data. IOP Conf. Ser. Earth Environ. Sci. 2021, 893, 012067. [Google Scholar] [CrossRef]
  93. Hayasaka, H.; Sepriando, A. Severe Air Pollution Due to Peat Fires During 2015 Super El Niño in Central Kalimantan, Indonesia. In Land-Atmospheric Research Applications in South and Southeast Asia; Vadrevu, K.P., Ohara, T., Justice, C., Eds.; Springer: Cham, Switzerland, 2018; pp. 129–142. [Google Scholar]
  94. Murdiyarso, D.; Frolking, S. Carbon accumulation of tropical peatlands over millennia: A modeling approach. Globa Change Biol. 2015, 21, 431–444. [Google Scholar] [CrossRef]
  95. UKCEH, Peatlands factsheet. UK Centre for Ecology & Hydrology (UKCEH), Wallingford, United Kingdom. 2019. Available online: https://www.ceh.ac.uk/sites/default/files/Peatland%20factsheet.pdf (accessed on 10 March 2025).
  96. Abram, N.J.; Gagan, M.K.; McCulloch, M.T.; Chappell, J.; Hantoro, W.S. Coral reef death during the 1997 Indian Ocean Dipole linked to Indonesian wildfires. Science 2003, 301, 952–955. [Google Scholar] [CrossRef]
  97. UNFCCC. Coral Reefs; United Nations Framework Convention on Climate Change (UNFCCC): Bonn, Germany, 2021; Available online: https://unfccc.int/news/coral-reefs (accessed on 10 March 2025).
  98. IUCN. The 2023 IUCN Situation Analysis on Ecosystems of the Yellow Sea with Particular Reference to Intertidal and Associated Coastal Habitats; The International Union for Conservation of Nature (IUCN): Bangkok, Thailand, 2023. [Google Scholar]
  99. Chua, K.B.; Chua, B.H.; Wang, C.W. Anthropogenic deforestation, El Niño and the emergence of Nipah virus in Malaysia. Malays. J. Pathol. 2002, 24, 15–21. [Google Scholar]
  100. Chua, K.B. Nipah virus outbreak in Malaysia. J. Clin. Virol. 2003, 26, 265–275. [Google Scholar] [CrossRef] [PubMed]
  101. Pikuta, E.V.; Marsic, D.; Bej, A.; Tang, J.; Krader, P.; Hoover, R.B. Carnobacterium pleistocenium sp. nov., a novel psychrotolerant, facultative anaerobe isolated from permafrost of the Fox Tunnel in Alaska. Int. J. Syst. Evol. Microbiol. 2005, 55 Pt 1, 473–478. [Google Scholar] [CrossRef] [PubMed]
  102. Bidle, K.D.; Lee, S.H.; Marchant, D.R.; Falkowski, P.G. Fossil genes and microbes in the oldest ice on Earth. Proc. Natl. Acad. Sci. USA 2007, 104, 13455–13460. [Google Scholar] [CrossRef] [PubMed]
Figure 1. Dual-pathway model of loop mechanics showing degradation cycles (red) and restoration interventions (green). Notes: This systems diagram illustrates the causal relationships in tropical peatland degradation. Oil palm expansion drives peat drainage, lowering water tables below critical thresholds (<4 cm depth). This triggers peat drying, escalating fire risk and CO2 emissions, while causing subsidence and coastal flooding. Consequences include local biodiversity loss, intensified drought, and climate extremes that amplify global radiative forcing. Rewetting policies represent an intervention to break this cycle by restoring hydrological balance.
Figure 1. Dual-pathway model of loop mechanics showing degradation cycles (red) and restoration interventions (green). Notes: This systems diagram illustrates the causal relationships in tropical peatland degradation. Oil palm expansion drives peat drainage, lowering water tables below critical thresholds (<4 cm depth). This triggers peat drying, escalating fire risk and CO2 emissions, while causing subsidence and coastal flooding. Consequences include local biodiversity loss, intensified drought, and climate extremes that amplify global radiative forcing. Rewetting policies represent an intervention to break this cycle by restoring hydrological balance.
Rsee 02 00017 g001
Figure 2. Ten provinces with the largest share of peatlands in Indonesia (in million ha). Source: [45].
Figure 2. Ten provinces with the largest share of peatlands in Indonesia (in million ha). Source: [45].
Rsee 02 00017 g002
Figure 3. Feedback loop control mechanism. Note: Feedback loops (positive/negative) govern system dynamics. Positive loops amplify deviations, for example, GHG emissions → warming → ice melt → albedo loss → accelerated warming. Negative loops stabilize, for example, carbon tax → emission cuts → reduced radiative forcing → slower warming. Stocks (system state) accumulate inflows over time.
Figure 3. Feedback loop control mechanism. Note: Feedback loops (positive/negative) govern system dynamics. Positive loops amplify deviations, for example, GHG emissions → warming → ice melt → albedo loss → accelerated warming. Negative loops stabilize, for example, carbon tax → emission cuts → reduced radiative forcing → slower warming. Stocks (system state) accumulate inflows over time.
Rsee 02 00017 g003
Figure 4. Peatlands in the earth system: self-regulating feedback loops and cascading effects. Notes: The above diagram alongside Figure 5 collectively reveal how breaching critical thresholds in Indonesia’s peatlands triggers emergent properties that cascade across Earth’s subsystems, driving planetary destabilization. The above figure establishes Earth’s structural domains (atmosphere, lithosphere, biosphere, cryosphere, hydrosphere), each governed by stability thresholds. However, in Figure 5, when human actions (agricultural expansion) breach these thresholds, unpredictable emergent properties override subsystem boundaries, leading to system collapses.
Figure 4. Peatlands in the earth system: self-regulating feedback loops and cascading effects. Notes: The above diagram alongside Figure 5 collectively reveal how breaching critical thresholds in Indonesia’s peatlands triggers emergent properties that cascade across Earth’s subsystems, driving planetary destabilization. The above figure establishes Earth’s structural domains (atmosphere, lithosphere, biosphere, cryosphere, hydrosphere), each governed by stability thresholds. However, in Figure 5, when human actions (agricultural expansion) breach these thresholds, unpredictable emergent properties override subsystem boundaries, leading to system collapses.
Rsee 02 00017 g004
Figure 5. Feedback mechanisms in tropical peatland degradation: Local to global cascading impacts.
Figure 5. Feedback mechanisms in tropical peatland degradation: Local to global cascading impacts.
Rsee 02 00017 g005
Table 1. Model component definitions.
Table 1. Model component definitions.
TermCategoryDescription
Oil palm expansionCore driverConversion of peat forests to plantations
Peat drainageProblem sourceArtificial water removal for agriculture
Water < 4 cmCritical thresholdWater depth triggering flammable conditions
Peat dryingDegradation processMoisture loss from drained peat
FiresAmplifying riskPeatland wildfires
CO2 emissionClimate driverGHG from decomposition/fire (4.3 Gt/year projected)
Global radiative forcingClimate impactEnhanced atmospheric heat trapping
Climate extremesSystemic consequenceIntensified droughts/heatwaves/storms
DroughtClimate feedbackProlonged dry periods (prolonged drought dries peatland ponds/streams, destroying aquatic habitats for fish, amphibians, and water-dependent insects)
SubsidencePhysical impactGround sinking (>2 cm/year)
Coastal floodingSecondary impactSeawater inundation of subsided land
Local biodiversity lossEcological costSpecies extinction from habitat destruction
Rewetting policiesSolution mechanismWater restoration in peatlands
Notes: Table 1 defines key components of tropical peatland degradation systems. It categorizes drivers such as oil palm expansion, processes such as peat drying, impacts such as coastal flooding, and solutions such as rewetting policies based on their functional roles in causal relationships.
Table 2. Core validation thresholds.
Table 2. Core validation thresholds.
Key VariableOperational DefinitionThreshold Source
Water table depthCritical flammability trigger<40 cm (Irreversible drying)
Rewetting policiesSolution mechanism for hydrologic restoration34% emission reduction
SubsidencePhysical collapse indicator>2 cm/year (Drained peat)
Notes: Table 2 specifies operational thresholds and mitigation metrics for critical degradation variables. It quantifies tipping points, for example, a <40 cm water table depth, solution efficacy (34% emission reduction from rewetting), and collapse indicators (>2 cm/year subsidence), with sources aligned with established biophysical mechanisms.
Table 3. Biodiversity in Kalimantan. Sources: [48,49,50].
Table 3. Biodiversity in Kalimantan. Sources: [48,49,50].
Common NameScientific NameIUCN Status
Fauna
Bornean orangutanPongo pygmaeusCritically endangered
Helmeted hornbillRhinoplax vigilCritically Endangered
White-winged duck Cairina scutulataCritically endangered
Painted terrapinBatagur borneoensisCritically endangered
Bornean terrapinOrlitia borneensisCritically endangered
Straw-headed Bulbuls Pycnonotus zeylanicusCritically endangered
Hairy-nosed otterLutra sumatranaEndangered
Proboscis monkeysNasalis larvatusEndangered
Wrinkled hornbillRhabdotorrhinus corrugatusEndangered
Asian arowanaScleropages formosusEndangered
False gharialTomistoma schlegeliiEndangered
Malayan sun bearHelarctos malayanusVulnerable
Chinese egretEgretta eulophotesVulnerable
Sunda clouded leopardsNeofelis diardiVulnerable
TomistomaTomistoma schlegeliiEndangered
Bornean white-bearded GibbonHylobates albibarbisEndangered
Müller’s gibbonHylobates muelleriEndangered
Flora
Clipeate pitcher plantNepenthes clipeataCritically endangered
Tualang treeKoompassia excelsaThreatened
Borneo kauri Agathis borneensisEndangered
Johore shoreaShorea johorensisVulnerable
Red balau or balangeranShorea balangeranVulnerable
Light red merantiShorea smithianaVulnerable
Bornean ironwoodEusideroxylon zwageriVulnerable
Table 4. Functions of the peatland ecosystem.
Table 4. Functions of the peatland ecosystem.
Function CategoryDescription
Regulating functions Climate regulation: Carbon sequestration and storage at local and global scales.
Erosion control: Vegetation stabilizes shorelines, reduces storm impacts (wind speed, waves, runoff), and traps sediments.
Water Purification: Vegetation breaks down contaminants (bacteria, pesticides, nutrients) and improves water quality.
Hydrological functionsGroundwater replenishment and water cycling (hydrology): Part of the water filters into the ground and recharges underground aquifers (groundwater reservoirs).
Flood mitigation: Acts as a sponge, absorbing excess water during wet seasons and releasing it gradually during dry seasons.
Provisioning functionsResource Supply: Provides timber, non-timber products (fuelwood, medicinal plants), and supports fish populations for local consumption
Supporting functionsBiodiversity Conservation: Coastal peatlands serve as transitional zones between aquatic and terrestrial ecosystems, hosting rare and endemic species
Unique Habitats: Acidic blackwater rivers in peat swamp forests support endemic fish species (e.g., Channa sp., Wallago leeri)
Cultural functionsGenetic Reservoir: Preserves germ plasm for agricultural and medicinal research
Table 5. Indonesian peatland governance: legal and policy framework for conservation, climate action, and biodiversity protection.
Table 5. Indonesian peatland governance: legal and policy framework for conservation, climate action, and biodiversity protection.
Document TypeNumber/YearKey ObjectivesFocus Area
Forestry LawNo. 41/1999» Prohibits land clearing in protected peat areas.Establishes state control over forests and peatlands to maximize public welfare. Focuses on biodiversity protection and ecosystem integrity.
» Mandates restoration of degraded peatlands.
» Requires permit revisions if peatland functions change.
Environmental Protection LawNo. 32/2009» Authorizes regional governments to implement environmental plans to prevent peat degradation.Ensures sustainable development through fire control, pollution prevention, and peatland restoration.
» Mandates restoration to protect carbon storage and biodiversity.
National PolicyNational Action Plan on Climate Change (2007)» Aims to rehabilitate 53.9 million hectares of degraded peatlands and forests by 2050.
» Prioritizes fire prevention, carbon sink enhancement, and biodiversity conservation.
Climate resilience, emission reduction, and large-scale ecosystem restoration.
Presidential DecreeNo. 62/2013Establishes the REDD+ Managing Agency to coordinate emission reduction efforts, including peatland conservation.Climate governance and REDD+ project implementation.
Presidential InstructionNo. 2/2007Accelerates rehabilitation of 1.45 million hectares of degraded peatlands in Central Kalimantan through rewetting and revegetation.Peatland revitalization and regional ecological recovery.
Presidential InstructionNo. 10/2011 (extended 2013, 2015, 2017)Imposes a moratorium on new licenses in primary forests/peatlands to align with REDD+ goals.Deforestation control and compliance with international climate commitments.
Presidential InstructionNo. 1/2016Establishes the Peatland Restoration Agency (BRG) to coordinate restoration, fire prevention, and biodiversity protection in 7 provinces.Peatland restoration, fire suppression, and habitat conservation.
» Rewetted 3.6 Mha by 2023.
Presidential InstructionNo. 5/2019Extends the 2011 moratorium, halting new permits in primary forests/peatlands to reduce emissions.Sustainable land-use practices and emissions reduction.
Government RegulationNo. 71/1990Promotes sustainable practices, including water table management and restrictions on monoculture plantations in sensitive peat zones.Sustainable peatland management and zoning.
Government RegulationNo. 68/1998 (amended 2011, 2015)Protects ecological integrity of conservation areas (KSA/KPA) and Essential Ecosystem Areas (KEE), including peatlands. Requires local-central collaboration for management.Conservation zoning and ecosystem protection.
Government RegulationNo. 4/2001Holds landholders accountable for fire prevention and environmental damage. Imposes fines for negligence.Forest and land fire control, environmental accountability.
Government RegulationNo. 71/2014 (amended by No. 57/2016)Protects peatlands through zoning (≥30% as protected area), bans harmful activities (drainage, burning), and mandates environmental permits.Legal enforcement, hydrological safeguards, and ecosystem protection.
» Enforces ≤ 40 cm water tables.
Government RegulationNo. 16/2017Provides technical guidelines for peatland restoration (rewetting, revegetation) to restore hydrological and ecological functions.Ecosystem restoration protocols and stakeholder roles.
Ministry RegulationNo. 10/2019Manages peat domes (hydrological units) to maintain water regulation and carbon storage.Hydrological conservation and peat dome protection.
» Mandates IoT-based water table monitoring.
Ministry RegulationNo. 7/2021Operationalizes Indonesia’s FOLU Net Sink 2030 targets, aiming for −140 Mt CO2e by 2030 through peatland protection and restoration.Climate action alignment, carbon sequestration, and emission reduction.
Ministry RegulationNo. 13/2024Implements Enhanced Nationally Determined Contribution (ENDC) targets (31.89–43.20% emissions reduction by 2030) through peatland conservation.International climate commitments and SDG alignment.
Table 6. The economic losses from the 1997/98 peatland fire. Source: [66].
Table 6. The economic losses from the 1997/98 peatland fire. Source: [66].
SectorMinimum Loss (USD Millions)Maximum Loss (USD Millions)Estimation Method
Agriculture 27502750Crop value loss: Market price valuation of destroyed crops (oil palm, rubber, rice) included opportunity costs of lost production cycles during drought
Forestry17612583Timber stock valuation (timber stock depletion): replacement cost of commercial species
NTFPs606606Market price method (non-timber forest products): income loss included subsistence gathering losses for rural communities
Flood protection397397Replacement cost approach (infrastructure repair costs): damaged drainage systems
Erosion and siltation13001300Productivity loss approach: sediment removal/water treatment facility costs
Carbon sink14461446Carbon pricing based on IPCC carbon accounting methods
Other costs including health, transportation and firefighting290320Hospitalizations and productivity loss from respiratory diseases and airline disruption costs
Total 85509402
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

Choy, Y.K.; Onuma, A. The Tropical Peatlands in Indonesia and Global Environmental Change: A Multi-Dimensional System-Based Analysis and Policy Implications. Reg. Sci. Environ. Econ. 2025, 2, 17. https://doi.org/10.3390/rsee2030017

AMA Style

Choy YK, Onuma A. The Tropical Peatlands in Indonesia and Global Environmental Change: A Multi-Dimensional System-Based Analysis and Policy Implications. Regional Science and Environmental Economics. 2025; 2(3):17. https://doi.org/10.3390/rsee2030017

Chicago/Turabian Style

Choy, Yee Keong, and Ayumi Onuma. 2025. "The Tropical Peatlands in Indonesia and Global Environmental Change: A Multi-Dimensional System-Based Analysis and Policy Implications" Regional Science and Environmental Economics 2, no. 3: 17. https://doi.org/10.3390/rsee2030017

APA Style

Choy, Y. K., & Onuma, A. (2025). The Tropical Peatlands in Indonesia and Global Environmental Change: A Multi-Dimensional System-Based Analysis and Policy Implications. Regional Science and Environmental Economics, 2(3), 17. https://doi.org/10.3390/rsee2030017

Article Metrics

Back to TopTop