57
1 Melt rate of Palcaraju glacier, Cordillera Blanca, Peru: attribution of anthropogenic influence and proposed methodology for calculating adaptation cost FHS in Geography Rupert Stuart-Smith January 2019 Word count: 11,491

Melt rate of Palcaraju glacier, Cordillera Blanca, Peru ...applicable to ‘slow-onset’ impacts, such as glacial melt, in which climate change can be shown to materially affect the

  • Upload
    others

  • View
    2

  • Download
    0

Embed Size (px)

Citation preview

Page 1: Melt rate of Palcaraju glacier, Cordillera Blanca, Peru ...applicable to ‘slow-onset’ impacts, such as glacial melt, in which climate change can be shown to materially affect the

1

Melt rate of Palcaraju glacier, Cordillera Blanca, Peru: attribution of anthropogenic influence and

proposed methodology for calculating adaptation cost

FHS in Geography

Rupert Stuart-Smith

January 2019

Word count: 11,491

Page 2: Melt rate of Palcaraju glacier, Cordillera Blanca, Peru ...applicable to ‘slow-onset’ impacts, such as glacial melt, in which climate change can be shown to materially affect the

2

Acknowledgements

The author would like to thank Professor Myles Allen for supervising this dissertation, for advice on

conceiving the study and comments on the text; Dr Sihan Li for co-supervision, providing Regional

Climate Model data, assistance with Python, and comments on the text; Dr Karsten Haustein for his

help with climate data operators; and Felix Heilmann for his translation of a court document from

German.

Contents

Acknowledgements 2

List of figures 3

1 Abstract 4

2 Summary 4

3 Introduction 5

4 Data 8

4.1 Regional Climate Model 8

4.2 Observations 9

4.3 Digital Elevation Model 10

5 Methods 11

5.1 Temperature bias correction 11

5.2 Precipitation bias correction 11

5.3 Distributed energy and mass balance (DEMB) glacier model 13

6 Results 21

6.1 Model evaluation 21

6.2 Energy fluxes 24

6.3 Meteorological data 26

6.4 Attribution of anthropogenic influence on glacier melt rate 28

7 Further modelling work 30

7.1 RCM outputs 30

7.2 Glacier model 31

8 Discussion of legal context 33

8.1 Implications for climate litigation 37

9 Conclusion 38

10 References 39

Appendix 47

Page 3: Melt rate of Palcaraju glacier, Cordillera Blanca, Peru ...applicable to ‘slow-onset’ impacts, such as glacial melt, in which climate change can be shown to materially affect the

3

List of figures

Figure 1: Satellite image of the southern Parque Nacional Huascarán. 8

Figure 2: Map of HadRM3P domain used for this study. 9

Figure 3: Palcaraju glacier outline. 10

Figure 4: Ablation time series for Shallap and Palcaraju glaciers (2007). 22

Figure 5: Mass balance time series for Shallap and Palcaraju glaciers (2007). 23

Figure 6: Energy fluxes for Shallap glacier. 25

Figure 7: Modelled energy fluxes for Palcaraju glacier. 25

Figure 8: Precipitation time series for the raw and bias corrected RCM data. 26

Figure 9: Bias corrected daily mean RCM temperature data. 27

Figure 10: Difference between actual and natural RCM temperature. 27

Figure 11: Full time series ablation rate for Palcaraju glacier. 30

Figure 12: Influence of the discount rate on costs attributable to climate change. 35

Page 4: Melt rate of Palcaraju glacier, Cordillera Blanca, Peru ...applicable to ‘slow-onset’ impacts, such as glacial melt, in which climate change can be shown to materially affect the

4

1 Abstract

Litigation concerning the influence of anthropogenic greenhouse gas (GHG) emissions on the melt rate

of Palcaraju glacier in the Peruvian Andes, and the resultant change in risk of a glacier lake outburst

flood (GLOF) from Lake Palcacocha, has entered an evidentiary phase. While previous attribution

studies on anthropogenic influence on glacier mass balance have been largely restricted to the

regional or global scale, evidence of the influence of emissions on the scale of the individual glacier is

needed to establish causation in the specific case before the courts. This paper assesses the

contribution of climate change to the melt rate of Palcaraju glacier. A vertically distributed surface

energy and mass balance model was developed and run at daily resolution for 29 years (1988-2016).

Boundary conditions were provided by a coupled global/regional climate model simulating the climate

both with and without anthropogenic influence, to quantify impacts attributable to past emissions.

Anthropogenic climate forcing is found to increase melt rate for all years by 20-80% and advances the

timing of Lake Palcacocha reaching its current depth by 15-37 years. Additionally, an original

methodology is proposed for calculating the costs of damages attributable to GHG emissions that is

applicable to ‘slow-onset’ impacts, such as glacial melt, in which climate change can be shown to

materially affect the timing of the occurrence of a natural disaster that would, eventually, have

occurred naturally.

2 Summary

The impact of anthropogenic climate forcing on the height-resolved surface energy balance and melt

rate of Palcaraju glacier (9°23’49”S 77°22’47”W, 4625-6175 m above sea level (a.s.l.)) was calculated

as follows. A vertically distributed surface energy and mass balance model was developed (henceforth

referred to as the ‘glacier model’), following the methodology of Gurgiser, Marzeion, et al. (2013) as

closely as the available data permitted, and where necessary, using parameterisations from further

sources, most notably Arnold et al. (1996). The glacier model was driven by surface temperature and

fluxes from the Hadley Centre regional climate model (RCM) version 3P (HadRM3P) coupled to the

global atmosphere model HadAM3P from the weather@home distributed climate modelling system.

RCM precipitation was bias-corrected using satellite-based measurements from the Tropical Rainfall

Measuring Mission (TRMM), in the absence of complete local surface observations, while temperature

was bias-corrected using elevation-adjusted observations from a meteorological station in the nearby

village of Anta. Two ensembles of simulations were compared, one in which the global atmosphere is

driven with observed sea-surface temperatures (SSTs), sea ice fractions (SIFs), and atmospheric

composition (denoted ‘actual’), and the other with SSTs, SIFs and atmospheric composition adjusted

Page 5: Melt rate of Palcaraju glacier, Cordillera Blanca, Peru ...applicable to ‘slow-onset’ impacts, such as glacial melt, in which climate change can be shown to materially affect the

5

to simulate a counterfactual ‘world that might have been’ in the absence of human influence on

climate (denoted ‘natural’).

An increase in melt rate as a result of anthropogenic GHG emissions is detected for all 29 years studied.

The exact magnitude of this increase is subject to considerable uncertainty. Nevertheless a 20-50%

increase in melt rate can be attributed to anthropogenic influence on the climate, across the modelled

period. In the natural scenario (without human influence), meltwater production does occur in a

comparatively small ablation zone near the glacier snout. The primary glacier response to climate

change is a rise in the equilibrium line altitude, expanding the area of the ablation zone, supplemented

by an acceleration in the melt rate of the ablation zone present in the natural run. The increased rate

of glacier meltwater production suggests that the rate of expansion of Lake Palcacocha has increased

as a result of human GHG emissions, and that the lake has reached a critical level, 15-37 years sooner

than it would have done under a natural climate, necessitating the implementation of costly risk

reduction measures.

Existing literature on climate change-induced loss and damage focuses primarily on the impacts of

meteorological events. The results presented here indicate that the application of frameworks based

solely on meteorological events to slow-onset events such as increasing GLOF risk is inappropriate. For

many extreme weather events, GHG emissions may be thought of as increasing the likelihood of

occurrence of a meteorological event of a given intensity whose likelihood may otherwise be assumed

to be constant over time. For GLOFs, which occur naturally, the consequence of emissions is to bring

the next flood forward in time. In this paper, I propose a methodology for calculating the costs

attributable to the GHG emissions of individual emitters for the impacts of glacial melt, and other

instances in which the impact of climate change is to advance the timing of an event’s occurrence.

3 Introduction

The risk of a GLOF from Lake Palcacocha, approximately 20km upstream of the city of Huaraz (see

figure 1), has been widely documented (e.g. Somos-Valenzuela et al., 2016). Reduced to 514,800 m3

by a GLOF in 1941 (Vilímek et al., 2005), the lake volume has since increased 34-fold to 17,325,207 m3,

primarily due to meltwater from the Palcaraju glacier, and, as of 2014, had a maximum depth of 73 m

(Autoridad Nacional del Agua, 2014; Portocarrero Rodríguez and Engility Corporation, 2014). A small

landslide-induced GLOF occurred in 2003 but caused little damage to Huaraz. The 1941 GLOF was the

result of a large breach of the lake’s moraine dam, resulting in a debris flow which destroyed one third

of the city of Huaraz, and caused between 1800 and 6000 deaths (Evans and Clague, 1994; Vilímek et

al., 2005; Somos-Valenzuela et al., 2016). Recent research on Lake Palcacocha has indicated that the

Page 6: Melt rate of Palcaraju glacier, Cordillera Blanca, Peru ...applicable to ‘slow-onset’ impacts, such as glacial melt, in which climate change can be shown to materially affect the

6

lake has very high outburst potential, and the occurrence of a GLOF would most likely be due to

avalanche-induced waves overtopping the moraine dam (Emmer and Vilímek, 2013).

The main controls on a GLOF of this type are the lake depth and avalanche size; a 30m reduction in

lake depth would reduce overtopping volume by 90% following an avalanche (Somos-Valenzuela et

al., 2016). In light of this, siphons were installed in 2011 to lower the water surface by 3-5m and leave

a freeboard of 12m, slightly reducing the GLOF risk to Huaraz. Any anthropogenic influence on the rate

of meltwater production above the lake, and therefore the rate of rise of lake level can therefore be

linked to altering the GLOF hazard (Somos-Valenzuela et al., 2016).

Ongoing litigation in the German courts (Lliuya v RWE) seeks to recover, from RWE (a German power

utility company), costs incurred by Huaraz in protecting the city from this threat, proportional to RWE’s

contribution to historical anthropogenic GHG emissions (Frank, Bals and Grimm, 2019). Now in an

evidentiary phase, question 2c of the court’s call for evidence concerns the role of GHG emissions in

altering meltwater production from Palcaraju Glacier: the question that this paper seeks to answer.

While there is no doubt that GHG emissions cause harm, tracing specific damages to specific emitters

is more complex. For most meteorological events, climate science can provide only probabilistic

attribution (linking emissions to an increased likelihood of an event occurring). It has been suggested

that more than doubling of risk may be sufficient to attribute damages to emissions (Marjanac, Patton

and Thornton, 2017; Simlinger and Mayer, 2019). For GLOFs, however, it may be possible to

demonstrate a robust link between emissions and glacial melt rates; the question for the courts

concerns the liability of emitters for advancing the timing of events: a matter hitherto not discussed

in literature concerning climate change attribution, loss and damage, and litigation.

Tropical glaciers are particularly sensitive to climatic changes, due in part to the thermal homogeneity

of the tropical troposphere (Kaser, 1999; Rabatel et al., 2013). Palcaraju (see Figure 3) is a high-

altitude, outer tropical glacier for which subtropical conditions prevail with a dry season in the austral

winter of May to August, a transitional season from September to December, and a wet season from

January to April. Notable accumulation occurs only in the wet season, predominantly due to an

easterly flow of moisture from the Amazon Basin, when increased snowfall raises albedo and reduces

the melt rate (Rabatel et al., 2013).

Few studies have sought to attribute glacier mass loss to anthropogenic forcing, particularly due to

challenges with downscaling modelled atmospheric circulation to the scale of glaciers, complexities in

time lag of glacier response to climate change (10-40 years, depending on size of ice mass, its dynamic

and thermodynamic regime, slope hypsography and the climatic environment) and poor long term

Page 7: Melt rate of Palcaraju glacier, Cordillera Blanca, Peru ...applicable to ‘slow-onset’ impacts, such as glacial melt, in which climate change can be shown to materially affect the

7

precipitation and humidity records (Marshall, 2007; Rabatel et al., 2013; Schauwecker et al., 2014). All

Cordillera Blanca glaciers are documented to be retreating (Thompson et al., 2017) and the

Intergovernmental Panel on Climate Change states that there is ‘high confidence that a substantial

part of the mass loss of glaciers is likely due to human influence’ (Bindoff, et al., 2013). The global

synchronicity of recent mountain glacier fluctuations also indicate that their causes are global in scale

(Kaser, 1999). 25% ± 35% of global glacier mass loss over the period 1851-2010 has been attributed to

anthropogenic causes, rising to 69% ± 24% for the period 1991-2010. Anthropogenic influence on

glacier mass is, however, spatially heterogeneous, underlining the importance of local attribution of

climate change impacts on melt rates (Marzeion et al., 2014).

The glacier coverage in the Cordillera Blanca has shrunk from its Little Ice Age peak of 900 km2 to 482

km2 in 2010 (Somos-Valenzuela et al., 2016). This rate of retreat is above the global average and has

been highest in the past 50 years, with a 40% loss in glacier-coverage in Peru observed between 1970

and 2003-2010 (Rabatel et al., 2013; Thompson et al., 2017). Significantly, the freezing line altitude

rose by 160 m on precipitation days between 1964/72 and 1983/92 (Schauwecker et al., 2014), and

some Cordillera Blanca glaciers show a reasonable correlation of Equilibrium Line Altitude (ELA) with

temperature (Kaser, Ames and Zamora, 1990). However, regional variations in anthropogenic

influence on the climate system preclude an assumption that enhanced glacier meltwater production

can be attributed to human influence alone. Individual glacier topography and aspect mediate its

response to changing local climate conditions and in the Cordillera Blanca, glaciers with eastern aspect

have seen a 71% reduction in volume, compared with 39% for southwestern aspect glaciers (such as

Palcaraju), hypothesised as being the result of the presence of late afternoon clouds protecting

western-facing glaciers from direct solar irradiation (Vilímek et al., 2005). The largest source of

variation glacier melt rates in the Cordillera Blanca is precipitation, whereas temperature remains

fairly constant (Gurgiser, Marzeion, et al., 2013). Whilst widely documented in history, GLOFs are

becoming increasingly common in the Andes, with outbursts occurring from other proglacial lakes in

2008 and 2009 (McGuire, 2012).

Glacier melt is controlled predominantly by energy balance at the interface between the atmosphere

and the glacier (Hock, 2005). This energy balance is determined by the meteorological conditions

above the glacier and the physical properties of the glacier itself. The interactions between the glacier

and the atmosphere are complex and atmospheric supply of energy for melt is mediated by the

presence of snow and ice.

Page 8: Melt rate of Palcaraju glacier, Cordillera Blanca, Peru ...applicable to ‘slow-onset’ impacts, such as glacial melt, in which climate change can be shown to materially affect the

8

Figure 1: Satellite image of the southern Parque Nacional Huascarán, showing the city of Huaraz (blue

outline), Palcaraju glacier (red outline) and the location of the meteorological station at Anta (yellow

outline).

4 Data

4.1 Regional Climate Model

Meteorological inputs to the glacier model were taken from regional climate model outputs from the

weather@home simulations. The weather@home project is run on climateprediction.net, a

distributed volunteer computing climate modelling framework. Weather@home (Massey et al., 2015;

Guillod et al., 2017) consists of the Met Office Hadley Centre Atmosphere-only model, with HadAM3P

running globally at 1.25x1.875 degrees (latitude by longitude) to drive the Met Office Hadley Centre

regional model (HadRM3P). The HadRM3P simulations used in this study were run at 0.22° resolution

over South America (figure 2). This is a one-way nesting where HadAM3P provides boundary

conditions for HadRM3P, but HadRM3P does not feed back into HadAM3P. The HadAM3P model

generates ensembles of climate conditions for a counterfactual world with the climate conditions

without anthropogenic influence on the climate (‘natural’) and present day conditions (‘actual’),

driven by prescribed sea surface temperatures and radiative forcing (Massey et al., 2015; Guillod et

al., 2017). The glacier model was driven by an ensemble of 22 simulations of daily precipitation,

temperature, downwelling shortwave radiation, and wind speed under both actual and natural

conditions, over 29 years from 1988-2016. Each ensemble members differs in the initial conditions of

the driving HadAM3P.

Page 9: Melt rate of Palcaraju glacier, Cordillera Blanca, Peru ...applicable to ‘slow-onset’ impacts, such as glacial melt, in which climate change can be shown to materially affect the

9

Figure 2: Map of the domain modelled by HadRM3P that was used in this study. Width of the yellow

band denotes the ‘sponge’ layer over which RCM fields are relaxed to the driving AGCM. Figure used

courtesy of Dr Sihan Li.

4.2 Observations

Daily temperature observations were taken from the nearby Comandante Fap German Arias

meteorological station at Anta (figure 1; -9.347°,-77.598°, 2749.9m), sourced from the Global

Historical Climatology Network Daily dataset (Menne et al., 2012). The temperature dataset was

incomplete, with only 1792 days of data available over a 23-year period (1988-2010). For the bias

correction detailed below, months with greater than five days of observations available were used to

form the reference period (for which both observations and RCM data was available).

Precipitation observations in the Pacific-Andean region, including the area surrounding Palcaraju

glacier, were scarce (Ochoa et al., 2014). It was not possible to use surface observations since only few

days of data were available for most months and the data that was available was recorded at an

elevation approximately 2000m lower than the glacier snout. The topographic effect on precipitation

in the Cordillera Blanca is such that a strong elevational precipitation gradient is expected. It was also

necessary to have complete observational data for the correction of dry day frequency. Therefore,

following Malone et al. (2015), daily precipitation data was derived from Version 7 of the TRMM Real-

Time Multi-Satellite Precipitation Analysis (TMPA-RT). TRMM combines microwave, microwave-

Page 10: Melt rate of Palcaraju glacier, Cordillera Blanca, Peru ...applicable to ‘slow-onset’ impacts, such as glacial melt, in which climate change can be shown to materially affect the

10

calibrated infrared, and combined microwave-IR estimates of precipitation (Goddard Earth Sciences

Data and Information Services Center, 2016) and is one of the most widely used satellite-based

methods for producing global precipitation estimates. TRMM data is only available from March 2000,

limiting the bias correction reference period for precipitation to the latter half of the RCM time series.

Research in the Peruvian Andes suggests that TRMM tends to underestimate precipitation intensity

(Malone et al., 2015), but represents precipitation seasonality well (Scheel et al., 2011; Ochoa et al.,

2014).

4.3 Digital Elevation Model

Palcaraju glacier (shown in figure 3) was divided into 32 grid boxes with 50m vertical resolution,

covering the elevations from 4625 to 6175m above sea level, with the glacier model solving the daily

melt rate for each vertical level on the glacier surface, following the method of Caidong and Sorteberg

(2010). Glacier hypsometry was taken from the Randolph Glacier Inventory and was valid as of

12/08/2008, the most recent data available (glacier ID: RIGI60-16.02263;

http://www.glims.org/maps/info.html?anlys_id=312068). The average slope of Palcaraju glacier is

32.4° (Pfeffer et al., 2014; RGI Consortium, 2017).

Figure 3: Palcaraju glacier outline. Data from the Randolph Glacier Inventory (Pfeffer et al., 2014; RGI

Consortium, 2017).

Page 11: Melt rate of Palcaraju glacier, Cordillera Blanca, Peru ...applicable to ‘slow-onset’ impacts, such as glacial melt, in which climate change can be shown to materially affect the

11

5 Methods

5.1 Temperature bias correction

Since RCM bias may be of sufficient magnitude to introduce significant errors if fed directly into the

glacier model, temperature data was first bias corrected against observations from the closest

meteorological station. Daily temperature data from the climate model was corrected using a variance

scaling methodology described by Chen et al. (2011) and detailed below. The model temperatures

were lapse rate adjusted from the elevation of the nearest grid box to the weather station elevation

using a lapse rate of 8 °C/km, which is the lapse rate calculated over a similar elevational range for the

Cordillera Blanca by Carey et al. (2012) and Schauwecker et al. (2014).

Model temperatures (𝑇𝑅𝐶𝑀) were first adjusted by adding the difference between monthly mean

observational and model temperatures, sampled identically over the same reference period (the

period for which there is observational and modelled data), �̅�𝑜𝑏𝑠 and �̅�𝑅𝐶𝑀,𝑟𝑒𝑓, respectively (equation

1). Variations about this mean temperature were then adjusted to the correct standard deviation by

subtracting the monthly mean corrected temperature, �̅�𝑐𝑜𝑟, from the daily corrected temperature,

𝑇𝑐𝑜𝑟, and multiplying this by the ratio of the observed (𝑆𝑜𝑏𝑠) to model (𝑆𝑅𝐶𝑀,𝑟𝑒𝑓) standard deviation

over the same period (equation 2). This approximately doubled the RCM standard deviation.

Temperatures were then returned to the correct magnitude by multiplying by �̅�𝑐𝑜𝑟 (equation 3).

𝑇𝑐𝑜𝑟 = 𝑇𝑅𝐶𝑀 + (�̅�𝑜𝑏𝑠 − �̅�𝑅𝐶𝑀,𝑟𝑒𝑓) (1)

𝑆𝑐𝑜𝑟 = 𝑆𝑜𝑏𝑠 / 𝑆𝑅𝐶𝑀,𝑟𝑒𝑓 (2)

𝑇𝑐𝑜𝑟,𝑎𝑙𝑙 = (𝑇𝑐𝑜𝑟 − �̅�𝑐𝑜𝑟) × 𝑆𝑐𝑜𝑟 + �̅�𝑐𝑜𝑟 (3)

5.2 Precipitation bias correction

For precipitation data, a similar bias correction approach as is used for temperature is inappropriate

as precipitation data only has non-negative values and variance scaling could give rise to impossible

negative precipitation values (Hempel et al., 2013). A number of methods of statistical bias correction

were considered for precipitation, including local intensity scaling (LOCI) and power transformation

(Schmidli, Frei and Vidale, 2006; Teutschbein and Seibert, 2012). Linear scaling methods account for

bias in the mean but cannot correct for biases in the wet-day frequency and intensity. Evaluation of

TRMM in the La Plata Basin of South America found good agreement between TRMM and gauge

precipitation data on monthly time scales and on the occurrence of daily precipitation events, but

comparatively poor agreement on precipitation intensity estimates at daily scales (Su, Hong and

Page 12: Melt rate of Palcaraju glacier, Cordillera Blanca, Peru ...applicable to ‘slow-onset’ impacts, such as glacial melt, in which climate change can be shown to materially affect the

12

Lettenmaier, 2008). Since daily precipitation data was required by the glacier model and biases in both

wet-day frequency and intensity appeared to be present in the RCM, the Inter-Sectoral Impact Model

Intercomparison Project (ISI-MIP) approach described by Hempel et al. (2013) was selected and

adapted to correct only the monthly mean precipitation intensity and wet-day frequency. Monthly

mean precipitation was first scaled, using a month-specific correction factor, and wet-day frequency

was then corrected.

Bias-corrected data is inevitably limited by the availability of the observational dataset and the

representation of physical processes in the climate model. Stationarity in the bias in the actual data

(March 2000-2016 was available for TRMM, and 2001-2016 was used for the correction, which is a

shorter period than the RCM simulation years) with respect to natural data is assumed when applying

the bias correction for the natural run and years outside of the reference period for which there are

observations. Given that non-stationarities in the bias may be present, this introduces additional

uncertainty (Maraun, 2012; Hempel et al., 2013).

A correction factor, 𝑐, is derived from monthly mean precipitation, with 𝑐 limited to 10 to avoid

unrealistically high precipitation values, shown by equation 4. Month specific values of 𝑐 ranged from

0.95 to 4.67.

𝑐 = ∑ 𝑃𝑖𝑜𝑏𝑠

𝑚=16

𝑖=1

∑ 𝑃𝑖𝑅𝐶𝑀

𝑚=16

𝑖=1

⁄ (4)

where 𝑃𝑖𝑜𝑏𝑠 and 𝑃𝑖

𝑅𝐶𝑀 are the observational and modelled monthly mean precipitation for each month

in the reference time series, and 𝑚 is the number of occurrences of each month in the reference

period. Daily precipitation (𝑃𝑖𝑗𝑅𝐶𝑀) is then multiplied by the month-specific correction factor, 𝑐, to give

the mean-corrected daily precipitation, �̃�𝑖𝑗𝑅𝐶𝑀, accounting for the fact that the magnitude of bias in

RCM precipitation may vary seasonally, as shown by equation 5 (Maraun et al., 2010).

�̃�𝑖𝑗𝑅𝐶𝑀 = 𝑐 ⋅ 𝑃𝑖𝑗

𝑅𝐶𝑀 (5)

A wet day threshold was calculated to match the dry day frequency of the RCM to the dry day

frequency in the observational time series (𝑃1𝑅𝐶𝑀[𝑁𝑑𝑟𝑦]). The number of dry days in a month in the

observational dataset was multiplied by the number of ensemble members (22) to determine the dry

day threshold of the ensemble aggregate such that the average number of dry days in an RCM month

across all ensemble members was equal to the number of dry days in the same months of the

observations (equation 6). The frequency of dry days was then adjusted by setting all values below the

dry day threshold 𝜖𝑑 to zero (equation 7).

Page 13: Melt rate of Palcaraju glacier, Cordillera Blanca, Peru ...applicable to ‘slow-onset’ impacts, such as glacial melt, in which climate change can be shown to materially affect the

13

𝜖𝑑 = 0.5 ⋅ 𝑃𝑖𝑗𝑅𝐶𝑀|( 𝑃𝑖𝑗

𝑅𝐶𝑀 ≤ 𝑃1𝑅𝐶𝑀[𝑁𝑑𝑟𝑦]) + 0.5 ⋅ 𝑃𝑖𝑗

𝑅𝐶𝑀|(𝑃𝑖𝑗𝑅𝐶𝑀 > 𝑃1

𝑅𝐶𝑀[𝑁𝑑𝑟𝑦]) (6)

�̂�𝑖𝑗𝑅𝐶𝑀 = 0 if (𝑃𝑖𝑗

𝑅𝐶𝑀 ≤ 𝜖𝑑) (7)

This accounts for the fact that there are artificially large numbers of ‘drizzle days’ in RCMs (Maraun et

al., 2010; Rauscher et al., 2010; Hempel et al., 2013). Therefore, the number of observed dry days,

𝑁𝑑𝑟𝑦, are determined during the reference period by counting the number of days with precipitation

of less than 1 mm (to measurement noise in the TRMM data). Since precipitation values smaller than

𝜖𝑑 are set to zero, frequency of dry days can be increased in the model data. If there are more days

with zero precipitation in the RCM than observational dataset, 𝑁𝑑𝑟𝑦 is selected to equal the number

of days in order to calculate the threshold. Days with rainfall less than or equal to 𝜖𝑑 are adjusted to

0mm. Additional wet days cannot be introduced as this would introduce physical inconsistencies to

the glacier model, such as days with rainfall but no clouds.

This exclusion of drizzle days would alter monthly means and therefore could affect the long-term

trend, so each month is normalised and the sum of precipitation in days adjusted to dry-days is

redistributed uniformly among wet days by the additive constant 𝑚𝑖𝑑𝑎𝑡𝑎 as per equation 8:

�̂�𝑖𝑗𝑑𝑎𝑡𝑎 = {

𝑃𝑖𝑗𝑑𝑎𝑡𝑎 + 𝑚𝑖

𝑑𝑎𝑡𝑎 if wet

0 if dry(8)

In the application period (that not covered by observations), monthly mean values were scaled using

the month-specific correction factor, 𝑐, and then dry day frequency was corrected with the monthly

mean of the ensemble aggregate dry days set to the monthly mean number of dry days in the TRMM

observations. The eliminated drizzle precipitation amount was then redistributed among the wet days

of the month, as noted above.

Due to lack of observations, insolation and wind outputs from the RCM were not bias corrected and

were used directly.

5.3 Distributed energy and mass balance (DEMB) glacier model

The distributed surface energy and mass balance model followed the methodology of Gurgiser,

Marzeion, et al. (2013), which was used for the nearby Shallap glacier, as closely as possible. Due to

the absence of surface measurements on Palcaraju glacier, alternative parameterisations for some

energy fluxes were required to be taken from other sources (notably Arnold et al. (1996) and Hock

(2005)) and it was therefore necessary to code up the model from scratch (the full script is presented

in the appendix). Glacier modelling studies without in situ measurements of all or most of snow depth,

Page 14: Melt rate of Palcaraju glacier, Cordillera Blanca, Peru ...applicable to ‘slow-onset’ impacts, such as glacial melt, in which climate change can be shown to materially affect the

14

relative humidity, global radiation, cloud cover, surface albedo and glacier temperature, are highly

unusual. As far as possible, all parameterisations used in the model have been previously

demonstrated to be effective on other high-altitude tropical glaciers. Similar process-based mass

balance models have been used in various climatic conditions for alpine glaciers (e.g. Mölg and Hardy,

2004; Mölg et al., 2012; Gurgiser, Marzeion, et al., 2013; Gurgiser, Molg, et al., 2013; Prinz et al., 2016).

Energy balance melt models are a physically based approach for melt computation which assesses

energy fluxes to and from the glacier surface. A central assumption in such models is that at a surface

temperature of 0 °C, any surplus energy at the surface-air interface is assumed to be used immediately

for melting (Hock, 2005). Similar studies have run model simulations at hourly time-steps (e.g.

Gurgiser, Molg, et al., 2013) but the resolution of available input climatological data rendered this

impossible. For this study, the DEMB was run at daily resolution and forced with daily bias-corrected

model precipitation and bias corrected daily mean temperature data. Observations of the diurnal cycle

of temperature and precipitation in this region were not available: if they had been, then adding a

realistic diurnal cycle to daily bias-corrected modelled data would have been a preferable approach

given the non-linearities. That said, other approximations, such as the omission of shadowing effects

from surrounding steep topography, and the relatively coarse resolution of the RCM inputs, might

render more detailed modelling less than ideal.

The DEMB model calculates available melt energy, 𝑄𝑀, for each grid cell, using the following

relationship:

𝑄𝑀 = 𝑄𝑁 + 𝑄𝑆 + 𝑄𝐿 (9)

where 𝑄𝑁 is net radiation to the glacier surface, 𝑄𝑆 is the sensible heat flux, and 𝑄𝐿 is the latent heat

release due to sublimation (Gurgiser, Marzeion, et al., 2013). Variations in 𝑄𝑁 dominate 𝑄𝑀, although

increases in 𝑄𝐿 in the austral winter can also reduce the energy available for melt. An additional term,

𝑄𝑅, the heat flux due to rain on the ice surface, is sometimes included in melt models, but due to its

small magnitude, is often excluded (e.g. Arnold et al., 1996; Malone et al., 2015), as is done here,

although refreezing of meltwater was accounted for in calculating ablation. The effects of subsurface

melting and sublimation are not considered (as is the case in Hock and Holmgren, 2005), with all

ablation presumed to be in liquid form. Rupper and Roe (2008) note that in regions with high

precipitation (as is the case for the tropical Andes), ablation is dominated by melt and surface runoff,

and sublimation is comparatively small in magnitude. Subsurface melt is more sensitive to refreezing

than surface melt, and is of much smaller magnitude than surface melt, justifying its exclusion (Mölg

et al., 2012). The melt rate, 𝑀, is therefore calculated as follows:

Page 15: Melt rate of Palcaraju glacier, Cordillera Blanca, Peru ...applicable to ‘slow-onset’ impacts, such as glacial melt, in which climate change can be shown to materially affect the

15

𝑀 =𝑄𝑀

𝜌𝑤𝐿𝑓− 𝑅𝑀 (10)

where 𝜌𝑤 denotes the density of water (1000 kg m-3), 𝐿𝑓 the latent heat of fusion (3.34 × 105 J kg-1),

and 𝑅𝑀 is refrozen meltwater (Hock, 2005).

The vertical air temperature gradient was assumed to be -0.55 °C / 100m, as used by Gurgiser, Molg,

et al. (2013), for the nearby Shallap Glacier, based on measurements at Zongo Glacier, Bolivia (Sicart

et al., 2011). Following Gurgiser, Marzeion, et al. (2013), a vertical gradient in precipitation was not

included, since the vertical precipitation gradient on high tropical mountains is very low.

5.3.1 Net radiation

Evidence from Antisana Glacier, Ecuador (0°28’ S) demonstrates that there is a close relationship

between net shortwave radiation and surface ablation throughout the year (Favier et al., 2004). Net

radiation balance was calculated as follows:

𝑄𝑁 = (𝐼 + 𝐷𝑡)(1 − 𝛼) + 𝑅↓ + 𝑅↑ (11)

where 𝐼 is direct and diffuse solar radiation, 𝐷𝑡 is reflected radiation from surrounding terrain. 𝛼 is

albedo, 𝑅↓ is longwave sky radiation, and 𝑅↑ is emitted longwave radiation (Hock, 2005). As in Gurgiser,

Marzeion, et al. (2013), longwave radiation from the surrounding terrain was not included.

5.3.2 Shortwave radiation

Distributed modelling generally requires global radiation to be split into diffuse and direct components

as shaded grid boxes receive only diffuse shortwave radiation (Hock, 2005). However, since this model

does not attempt to account for obstruction of direct shortwave radiation by surrounding topography

(and Palcaraju glacier’s orientation and low-latitude location is such that this is likely of minor

influence), there is no need to model direct and diffuse shortwave radiation separately. RCM

shortwave radiation outputs could therefore be used directly.

The final component of global radiation, radiation reflected from surrounding terrain, 𝐷𝑡 is

parameterised as:

𝐷𝑡 = 𝛼𝑚 ⋅ 𝐼 ⋅ sin2 (𝛽

2) (12)

where 𝛽 is the gradient of the slope, 𝛼𝑚 is the mean albedo of surroundings, and 𝐼 is downwelling

shortwave radiation from the RCM (Arnold et al., 1996). Arnold et al. (1996) use a value of 0.4 for 𝛼𝑚.

The value of 𝐷𝑡 was found to be very small and could have been excluded from the model.

Page 16: Melt rate of Palcaraju glacier, Cordillera Blanca, Peru ...applicable to ‘slow-onset’ impacts, such as glacial melt, in which climate change can be shown to materially affect the

16

5.3.3 Longwave radiation

Emitted predominantly by atmospheric water vapour, CO2 and ozone, variations in longwave radiation

incident on the glacier surface are primarily controlled by cloud cover, and the amount and

temperature of water vapour. Melt models usually estimate longwave radiation from its correlation

with air temperature and vapour pressure at the surface, parameterised as:

𝑅↓ = 𝜖𝑐𝜎𝑇𝑎4𝐹(𝑛) (13)

where 𝜖𝑐 is the full-spectrum clear-sky emissivity, 𝜎 is the Stefan Boltzmann constant,

5.67 × 10−8Wm-2K-4, 𝑇𝑎 is air temperature (K) and 𝐹(𝑛) is a cloud factor which describes the

increase in radiation due to clouds, as a function of cloud amount, 𝑛. Effective emissivity, 𝜖𝑒 is defined

as the product of 𝜖𝑐 and 𝐹(𝑛) and ranges from 0.7 under clear-sky conditions to close to 1 under

overcast conditions (Hock, 2005). A parameterisation proposed by Arnold et al. (1996) is employed

here for 𝜖𝑒:

𝜖𝑐 ⋅ 𝐹(𝑛) = 𝜖𝑒 = 8.733 × 10−3 ⋅ 𝑇𝑎0.788(1 + 𝑘𝑛) (14)

where 𝑘 is a constant depending on cloud type, and 𝑛, from 0 to 1, is the fraction of the sky covered

by cloud. A constant value of 𝑘 = 0.26, the mean values for altostratus, altocumulus, stratocumulus,

stratus and cumulus cloud types, is used (Arnold et al., 1996). In the absence of model output data on

cloud coverage, 𝑛 was set at 1 for precipitation days and 0.3 for dry days, although these

approximations must be used cautiously. The choice of 𝑛 = 0.3 for dry days was based on the

following relationship, and an approximation of 10K for the daily temperature range (equation 15),

where 𝑥 is the daily temperature range (Arnold et al., 1996).

𝑛 = −0.098𝑥 + 1.285 (15)

Longwave outgoing radiation is controlled by the glacier temperature, and approximated by the

relationship in equation 16:

𝑅↑ = 𝜖𝜎𝑇𝑠4 (16)

where 𝜖 is the surface infra-red emissivity, assumed to be 1 (Gurgiser, Marzeion, et al., 2013), 𝜎 is the

Stefan Boltzmann constant and 𝑇𝑠 is the absolute temperature of the emitting surface (K). If the energy

available for melt is positive, the surface temperature is assumed to be 0 °C (Hock, 2005; Hock and

Holmgren, 2005). Outgoing longwave radiation is often prescribed as 316 Wm-2 (Oerlemans, 1993), as

infra-red emissivity of snow is fairly insensitive to snowpack parameters (Warren, 1982). Snow

Page 17: Melt rate of Palcaraju glacier, Cordillera Blanca, Peru ...applicable to ‘slow-onset’ impacts, such as glacial melt, in which climate change can be shown to materially affect the

17

emissivity has been measured as 99% (Griggs, 1968), but was here assumed to be 100% for both snow

and ice (Gurgiser, Molg, et al., 2013).

5.3.4 Albedo

Glacier albedo varies considerably, ranging from 0.1 for dirty ice to greater than 0.9 for fresh snow

(Hock, 2005). This is a significant control on the spatial and temporal distribution of meltwater

production. Summer snowfall, and resultant increased albedo reduces melt and runoff. Snow aging is

also an important consideration: fresh snow albedo may fall by 0.3 within a few days (Hock, 2005).

The performance of an energy-balance model is highly sensitive to the treatment of albedo

(Oerlemans, 1992).

Glacier surface albedo increases with cloudiness, as clouds preferentially absorb near-infrared

radiation, increasing the fraction of visible light incident on the glacier surface, for which albedo is

higher, and atmospheric water content. Snow albedo has been found to increase 3-15% on transition

from clear-sky to overcast conditions. Albedo modelling is complicated by difficulties in quantitatively

relating variations in albedo to their causes. Some studies have kept albedo constant, such as Malone

et al. (2015) which uses a value of 0.6. This was deemed insufficient given the high temporal resolution

of this model, and albedo was calculated here at daily, and 50m vertical, resolution.

Internal generation of albedo by the mass balance model is considered necessary due to considerable

variations of albedo in space and time. The dependence of albedo on crystal structure, ice and snow

morphology, dust and soot concentrations, morainic material, liquid water in veins, surface water

runoff, solar elevation and cloudiness complicates albedo parameterisation considerably, but these

are rarely accounted for in glacier modelling, and this paper also does not account for these factors.

Gurgiser, Marzeion, et al. (2013) use a parameterisation proposed by Oerlemans and Knap (1998). This

parameterises the albedo of a snow-covered glacier site at day 𝑖 as dependent on the age of the snow

at the surface:

𝛼𝑠𝑛𝑜𝑤𝑖 = 𝛼𝑓𝑖𝑟𝑛 + (𝛼𝑓𝑟𝑠𝑛𝑜𝑤 − 𝛼𝑓𝑖𝑟𝑛) ⋅ exp (

𝑠 − 𝑖

𝑡∗) (17)

where 𝛼𝑓𝑖𝑟𝑛 is 0.53, 𝛼𝑖𝑐𝑒 is 0.34, 𝛼𝑓𝑟𝑠𝑛𝑜𝑤 is 0.75. 𝛼𝑓𝑖𝑟𝑛 is the characteristic albedo for firn (old,

compacted snow), 𝛼𝑓𝑟𝑠𝑛𝑜𝑤 is that of fresh snow, and 𝑡∗ controls the rate at which the snow albedo

approaches the firn albedo following a snowfall event. 𝑠 is the number of the day on which the last

snowfall occurred. When the snow depth, 𝑑, decreases, albedo gradually decreases from that of snow

to that of ice (𝛼𝑖𝑐𝑒). Therefore, a final equation for albedo is as follows:

Page 18: Melt rate of Palcaraju glacier, Cordillera Blanca, Peru ...applicable to ‘slow-onset’ impacts, such as glacial melt, in which climate change can be shown to materially affect the

18

𝛼𝑖 = 𝛼𝑠𝑛𝑜𝑤𝑖 + 𝛼𝑖𝑐𝑒 − 𝛼𝑠𝑛𝑜𝑤

𝑖 ⋅ exp (−𝑑

𝑑∗) (18)

where 𝑑∗ is a characteristic scale for snow depth, with a value of 32 (Oerlemans and Knap, 1998).

When snow depth equals 𝑑∗, snow cover contributes 1/𝑒 to the albedo, and the underlying surface,

(1 −1

𝑒). If snow depth equals 3𝑑∗, the underlying surface contributes about 5% to the albedo.

Snowfall events are defined as days with ≥ 20mm of snow accumulation (Oerlemans and Knap, 1998),

and snowfall was calculated from precipitation according to the method used by Caidong and

Sorteberg (2010) and Tarboton et al. (1995):

𝑃𝑟 = 𝑃 | 𝑇𝑎 ≥ 3℃ (19)

𝑃𝑟 = 𝑃 (𝑇𝑎 − 𝑇𝑏

𝑇𝑟 − 𝑇𝑏) | − 1℃ < 𝑇𝑎 < 3℃ (20)

𝑃𝑟 = 0 | 𝑇𝑎 < −1℃ (21)

𝑃𝑠 = (𝑃 − 𝑃𝑟) (22)

where 𝑃 is the daily precipitation, which is partitioned into rain, 𝑃𝑟, and snow, 𝑃𝑠 (mm water equivalent

(w.e.)) based on air temperature, 𝑇𝑎. As daily mean temperature data is used, the threshold air

temperatures for rain and snow (3°C and -1°C, respectively) deviate significantly from zero to take into

account diurnal temperature variations, as suggested by Caidong and Sorteberg (2010). Due to

unavailable data, no wind-blown drifting snow or accumulation due to avalanches is accounted for,

and following the methodology of Caidong and Sorteberg (2010), snow thickness, 𝐻 (mm w.e.) is

calculated as follows:

𝐻 = 𝐻1 + 𝑃𝑠 −𝑄𝑚

𝜌𝑤𝐿𝑓

(23)

where 𝜌𝑤 is water density, 𝐿𝑓 is the specific latent heat of fusion, 𝑃𝑠 is daily snow accumulation and

𝐻1 is the snow thickness of the day before. The model was spun up for one year to equilibrate snow

depth over a full seasonal cycle (Caidong and Sorteberg, 2010).

5.3.5 Turbulent heat fluxes

Turbulent heat fluxes are driven by the temperature and moisture gradients between the air and the

surface, and by lower atmosphere turbulence. Although these are generally small when averaged over

extended periods of time, turbulent heat fluxes may exceed radiation flux on certain days, and the

Page 19: Melt rate of Palcaraju glacier, Cordillera Blanca, Peru ...applicable to ‘slow-onset’ impacts, such as glacial melt, in which climate change can be shown to materially affect the

19

highest melt rates frequently coincide with high turbulent flux values, notably on temperate glaciers

(Hock, 2005).

Arnold et al. (1996) provide a parameterisation for distributed energy and mass balance modelling, in

which turbulent heat fluxes are calculated for each cell of a melting glacier surface with a temperature

of 0 °C and vapour pressure equal to the saturated vapour pressure at 0 °C. This method requires an

assumption of adiabatic stratification in a Prandtl-type boundary layer, for which vertical fluxes of

energy and momentum constant with height (Kraus, 1973). Relationships for 𝑄𝑆 and 𝑄𝐿, the sensible

and latent heat fluxes, respectively, are therefore:

𝑄𝑆 = 𝐾𝑠𝑃𝑇𝑉 (24)

𝑄𝐿 = 𝐾1𝛿𝑒𝑉 (25)

where 𝐾𝑠 (4.42 × 10−6m kg-1 K-1 s2) and 𝐾1 (7.77 × 10−3m kg-1 s2) are coefficients, 𝑃 is atmospheric

pressure (Pa), 𝑇 is air temperature (°C), 𝑉 is wind speed (ms-1, taken from the RCM) and 𝛿𝑒 is the

difference between the vapour pressure of air and the saturation vapour pressure at the glacier

surface (Pa, Arnold et al., 1996). The vapour pressure of a melting surface was taken as 6.11 hPa, and

saturation vapour pressure calculated using an exponential approximation to the Clausius-Clapeyron

relation for moist air over ice and water (Hock, 2005). Numerical values of 𝐾𝑠 and 𝐾1 include an

adjustment for latent heat of fusion and water density, and vary with snow and ice surfaces and, for

𝐾1, whether the surface is experiencing condensation or evaporation.

Atmospheric pressure was calculated at the grid box elevation according to the barometric formula:

𝑃 = 𝑝0 ⋅ (1 −𝐿𝑅 ⋅ ℎ

𝑇0)

𝑔⋅𝑀𝑅0⋅𝐿

(26)

where 𝑝0 is the standard sea level atmospheric pressure (101325 Pa), 𝐿R is the lapse rate for dry air

(0.0065 K/m), ℎ is the grid box altitude, 𝑇0 is the sea level standard temperature (288.15K), 𝑔 is the

Earth-surface gravitational acceleration (9.80665 m s-2), 𝑀 is the molar mass of dry air (0.0289644 kg

mol-1) and 𝑅0 the universal gas constant (8.31447 J mol-1 K-1).

5.3.6 Other heat fluxes

The sensible heat flux of rain is generally unimportant in the overall surface energy balance of a glacier

and often neglected in models (e.g. Hock and Holmgren, 2005). As its magnitude, compared to other

fluxes, is small, heat flux from precipitation is not considered in the model, nor is mass redistribution

due to wind and gravity (Gurgiser, Marzeion, et al., 2013).

Page 20: Melt rate of Palcaraju glacier, Cordillera Blanca, Peru ...applicable to ‘slow-onset’ impacts, such as glacial melt, in which climate change can be shown to materially affect the

20

5.3.7 Refreezing of meltwater

Accurate parameterisation of meltwater refreezing is challenging even in complex snowpack models

(Etchevers et al., 2004), and some modelling efforts do not take this into account (e.g. Jiang et al.,

2010; Malone et al., 2015; Sagredo et al., 2014). However, this is still a variable widely included in

glacier modelling (e.g. Banwell et al., 2016, 2012) and its exclusion risks overestimating ablation losses

from the glacier and underestimating glacier mass balance. Banwell et al. (2012) find that 6% of surface

meltwater and rainwater refreezes in the snowpack (rather than becoming runoff) in West Greenland;

for a Tibetan glacier, Mölg et al. (2012) found an average of 13% of surface melt refreezes; and for

another Tibetan glacier, Fujita and Ageta (2000) as much as 20% of infiltrated water is refrozen and

does not run off from the glacier. Refrozen meltwater, 𝑀𝑅 for each time-step was calculated as follows,

such that the fraction of meltwater refreezing decreases as the glacier temperature approaches the

melting point (Oerlemans, 1991; Wright et al., 2007).

𝑀𝑅 = 𝑄𝜃 (27)

where 𝑄 is the surface energy budget, 𝜃 is the temperature of the thermally active layer, taken as

equivalent to an upper 2m thickness of solid ice, set to the mean annual air temperature, at the start

of the year (Oerlemans, 1991). Due to the negativity constraint on glacier temperature, temperature

was not permitted to rise above 273.15 K.

𝑥𝑖 =𝑄𝑖{1 − exp(𝜃𝑖)}

𝐿𝑟

(28)

Where 𝑥𝑖 is the mass of refrozen meltwater, 𝐿𝑟 is the latent heat generated by refreezing. Through

the year, glacier temperature is adjusted only by the addition of latent heat due to refreezing of

meltwater, calculated as:

𝐻𝑖𝑐𝑒 = 𝑄𝑖 − 𝑀𝑅 (29)

5.3.8 Initial conditions

Separate model simulations were run for each precipitation and shortwave radiation ensemble

member (to avoid physically improbable sunny days with large amounts of precipitation), and the

ensemble mean of wind velocity and bias corrected ensemble mean of temperature, for each year.

Using the methodology of Rye et al. (2010) each cell of the digital elevation model is initialised as bare

ice and spun-up for one year to generate initial conditions of snow depth on the glacier. Rye et al.

(2010) found that 20 annual cycles were required for mass balance and firn to reach a steady state (in

the high Arctic), Banwell et al. (2012) used 5 years of spin up for the West Greenland Ice Sheet,

Page 21: Melt rate of Palcaraju glacier, Cordillera Blanca, Peru ...applicable to ‘slow-onset’ impacts, such as glacial melt, in which climate change can be shown to materially affect the

21

Bougamont et al. (2005) used a spin up period of 6 months, and Gurgiser, Molg, et al. (2013) use a

three-month spin-up period for the nearby Shallap Glacier, also in the Cordillera Blanca of Peru. These

variations are attributed to varying initial conditions, which influence the time taken for the iterative

process to reach a stable solution, rather than the steady state itself (Banwell et al., 2012). For

Palcaraju glacier, a 1-year spin-up period was found to produce a stable melt rate, and therefore used

for this paper.

6 Results

6.1 Model evaluation

In the absence of measurements of ablation rate at Palcaraju glacier, the glacier model was evaluated

through a comparison with the results of Gurgiser, Marzeion, et al. (2013) from the nearby Shallap

glacier. The comparison was undertaken for 2007, the only complete year modelled for the Shallap

glacier. This was based on the expectation of a similar ablation rate over the same time period on

Shallap and Palcaraju glaciers and the success of Gurgiser, Marzeion, et al. (2013) in reproducing

measured ablation. Data from Shallap glacier indicates a strong semi-annual cycle, whereas the model

of Palcaraju glacier appears to produce only an annual cycle in melt rates. This is most likely the result

of reductions in albedo and therefore increasing net shortwave radiation to the glacier in April and

October, for Shallap glacier, which does not appear in the model use here, due to differences in the

meteorological inputs to the model, the comparatively slow response of albedo to dry periods in the

model, and differences in the glacier site. Nonetheless, across the whole year, ablation rate is found

to be of similar magnitude (20-40 mm/day in the ablation zone), although Palcaraju glacier (as

modelled here) tends to have a slightly higher melt rate for much of April – December 2007, in the

actual scenario. The altitudinal extent of the ablation zone is similar for both glaciers, peaking at

around 5000m (shown in figure 4). Melt rates are lowest in the wet season (austral summer) and the

elevation of the ablation zone increases through the austral winter, peaking at around 5100m in

September. The actual scenario shows a higher melt rate in the ablation zone for much of 2007, and

the equilibrium line altitude is raised from around 4800m (in the natural scenario) to over 5000m in

the actual scenario. Attribution of changed melt rate to anthropogenic climate forcing will be

discussed further in the in section 6.4.

Page 22: Melt rate of Palcaraju glacier, Cordillera Blanca, Peru ...applicable to ‘slow-onset’ impacts, such as glacial melt, in which climate change can be shown to materially affect the

22

Figure 4: Ablation time series for Shallap glacier calculated as daily means for 10m steps in altitude

(top; Gurgiser, Marzeion, et al., 2013) and ensemble mean modelled ablation rate of Palcaraju glacier

under actual (middle) and natural (bottom) climate conditions (50m vertical resolution), all for 2007.

The use of the ensemble mean in the middle and bottom panels smooths out short term variations in

melt rate.

Page 23: Melt rate of Palcaraju glacier, Cordillera Blanca, Peru ...applicable to ‘slow-onset’ impacts, such as glacial melt, in which climate change can be shown to materially affect the

23

A comparison of modelled mass balance values was also conducted with Shallap glacier (Gurgiser,

Marzeion, et al., 2013; figure 5). In line with the ablation data, above, an annual cycle in mass balance

is found for Palcaraju glacier, as compared to the semi-annual cycle present on Shallap glacier, where

mass balance peaks at the glacier snout in April and October. However, the equilibrium line altitude

and mass balance magnitude are similar for the two glaciers in the austral winter of June to September

(with the exception of a brief increase in Shallap mass balance in late July, following a snowfall event).

Glacier flow modelling would be required to evaluate whether the negative mass balance found below

5000m would lead to glacier retreat.

Figure 5: Mass balance for Shallap glacier (top, Gurgiser, Marzeion, et al., 2013) and Palcaraju

glacier, averaged across all ensemble members of the actual run, for 2007. The use of the ensemble

mean in the middle and bottom panels smooths out short term variations in melt rate.

Page 24: Melt rate of Palcaraju glacier, Cordillera Blanca, Peru ...applicable to ‘slow-onset’ impacts, such as glacial melt, in which climate change can be shown to materially affect the

24

6.2 Energy fluxes

Melt rate is most sensitive to variations in net shortwave radiation. This is controlled by seasonal

variations in radiation intensity (which are comparatively moderate at such low latitudes), cloud cover,

and albedo. Cloud cover alters the downwelling shortwave radiation produced by the RCM on a daily

scale. Albedo drives more gradual, high magnitude changes in net shortwave radiation, and is

modelled as being dependent on snow depth and the time elapsed since the last snowfall event. Net

shortwave radiation falls to its lowest levels in the wet season when cloud cover is high, and albedo at

the level of fresh snow due to frequent snowfall and thick snow depth. In the dry season, albedo is

reduced to that of ice (0.34) and net shortwave energy flux increases.

As shown in figure 7, similar to the results of Gurgiser, Marzeion, et al. (2013; figure 6), net longwave

radiation peaks in the months of the austral winter, although the modelled values are, in the dry

months, significantly lower than those found by Gurgiser, Marzeion, et al. (2013). The most significant

differences between the model results on Palcaraju and those from Shallap glacier is in the magnitude

of latent heat flux, and the near-absence of sensible heat flux, partly explained by the low or sub-

freezing surface temperatures for much of the glacier. In the model parameterisation, latent heat flux

was calculated as the product of wind velocity, the difference between the vapour pressure of air and

the saturation vapour pressure at the glacier surface, and a constant. In the absence of surface

measurements of saturation vapour pressure at the glacier surface, this was calculated using the

Arden Buck equation and therefore estimated based on daily ensemble mean air temperature at the

grid box elevation. The use of ensemble mean wind speeds as model inputs might also contributed to

the comparatively small variability in latent heat flux, particularly in the austral winter months.

Sensible heat flux was generally low on Shallap glacier, but could be high in June and July, due to high

surface air temperatures. The use of ensemble mean air temperatures to drive this glacier model

appears to obscure any short-term increases in air temperature and surface air temperature rarely

rises much above 0 °C, even at the lowest elevations of the glacier.

This attempt at modelling the energy fluxes on Palcaraju glacier is hindered by the lack of surface

observations, but the modelled energy available for melt is still similar for Palcaraju and Shallap

glaciers. In light of this, the model was considered to be sufficient for an assessment of the influence

of climate change on melt rate of Palacaraju glacier, although the exact magnitude of this change

predicted by this model (given the constraints in data availability) should be treated with caution.

Page 25: Melt rate of Palcaraju glacier, Cordillera Blanca, Peru ...applicable to ‘slow-onset’ impacts, such as glacial melt, in which climate change can be shown to materially affect the

25

Figure 6: Stacked chart of three-day running means of energy fluxes (orange = net shortwave, blue =

net longwave, red = sensible flux, grey = latent flux, QG = ground heat flux, black = energy available for

melt) spatially averaged over the glacier area, for 2007. Melt energy in the subsurface is indicated as

a dotted line. Grey and blue bars at top denote the dry and wet season, respectively (adapted from

Gurgiser, Marzeion, et al., 2013).

Figure 7: Stacked chart of modelled energy fluxes (net shortwave, net longwave, sensible and latent

fluxes) for the ablation zone (< 5000m) of Palcaraju glacier (actual run), averaged across all ensemble

members, 2007. Energy flux values are in W m-2. Note change in vertical axis scale relative to figure 6.

Page 26: Melt rate of Palcaraju glacier, Cordillera Blanca, Peru ...applicable to ‘slow-onset’ impacts, such as glacial melt, in which climate change can be shown to materially affect the

26

6.3 Meteorological data

6.3.1 Precipitation

One randomly selected ensemble member of the precipitation time series used for the glacier model

runs is presented below, in figure 8, for both the actual and natural scenarios alongside the TRMM

data used for the bias correction. For 20/29 years, the annual mean precipitation values are higher in

the actual runs, and in some years, by as much as 1 mm/day. Corrected model biases are noted in

section 5.2.

Figure 8: Precipitation time series for the raw and bias corrected RCM data (1 randomly chosen

ensemble member, 1987-2016) with the bias corrected data shown in blue, and raw RCM daily

precipitation values in green. The TRMM data used for bias correction is plotted for 2001-2016, in red.

6.3.2 Temperature

As explained above, RCM temperature data was bias corrected against lapse rate-adjusted

observations from the nearest meteorological station to the glacier. RCM data was corrected to

increase its monthly mean and standard deviation to that of the observational dataset, the results of

which are presented in figure 9 for actual and natural runs. Figure 10 shows the strong warming trend

in the temperature anomaly between 1987 and 2016. Over this 30-year period the temperature

anomaly rises by around 1 °C, with the strongest warming trend between 1992 and 2016. Periods of

Page 27: Melt rate of Palcaraju glacier, Cordillera Blanca, Peru ...applicable to ‘slow-onset’ impacts, such as glacial melt, in which climate change can be shown to materially affect the

27

cooling are observed in both the natural and actual runs in 1988-1989, 2007-2008, and 2010-2011, all

during La Niña events.

Figure 9: Bias corrected daily mean RCM temperature data from the factual (green) and natural (blue)

runs, at the bottom of Palcaraju glacier, Peru.

Figure 10: Bias corrected difference between actual and natural RCM temperature, Comandante

German Arias Graziani meteorological station, 2749.9m elevation, 1987-2016.

Page 28: Melt rate of Palcaraju glacier, Cordillera Blanca, Peru ...applicable to ‘slow-onset’ impacts, such as glacial melt, in which climate change can be shown to materially affect the

28

6.4 Attribution of anthropogenic influence on glacier melt rate

The model results for the area of the glacier below 5000m averaged across all ensemble members and

each year show an increase in melt rate in the actual scenario, as compared to the natural scenario,

for all years (figure 11, bottom panel). Melt rate anomalies are generally greatest in the late austral

winter (the dry season, with highest melt rates) and for most years, the greatest anomalies occur at

the top of ablation zone, due to the temperature driven increase in the equilibrium line altitude.

Changes in melt rate as a result of anthropogenic forcing are attributed here on century timescales,

and melt rate is elevated throughout the timescale studied. However, on the decadal time scales, the

contribution of rising temperatures between 1988 and 2016 to melt rate is dampened by other factors,

most likely increases in precipitation and cloud cover, and therefore a reduction in net shortwave

radiation flux at the glacier surface, due to increased albedo and decreased downwelling shortwave

radiation. The attributable increase in melt rate of the Palcaraju glacier ablation zone ranges from 5

mm/day to 35 mm/day, and natural melt rates in the ablation zone range from 25 to 45 mm/day.

While the exact magnitude of this increase is subject to considerable uncertainty, an increase in

ablation zone melt rate of 20-80% is seen for most of the time series. Since the increase in ablation

zone melt rate is between 20 and 50% for most of the time series, this would be equivalent to

advancing the timing of Lake Palcacocha reaching its current levels by approximately 15-37 years (since

lake levels were reduced significantly by the 1943 GLOF). Further research would be required to

explain why the dampening effect on increased melt rate described above appears to be

comparatively small prior to 1988 but prevents further increases in melt rate over the period studied,

despite the significant rise in temperature (ap proximately 1°C) over the same period, as shown in

figure 10.

Page 29: Melt rate of Palcaraju glacier, Cordillera Blanca, Peru ...applicable to ‘slow-onset’ impacts, such as glacial melt, in which climate change can be shown to materially affect the

29

Page 30: Melt rate of Palcaraju glacier, Cordillera Blanca, Peru ...applicable to ‘slow-onset’ impacts, such as glacial melt, in which climate change can be shown to materially affect the

30

Figure 11: Ablation rate, Palcaraju glacier, Cordillera Blanca, Peru, 1988-2016 under actual (top) and

natural (middle) climate conditions, and the attributable change in melt rate across the glacier, over

the same period (bottom). Note the episodic nature of the attributable change in melt rate, with strong

inter annual and decadal variability.

7 Further modelling work

7.1 RCM outputs

Due to the steep orography of the Cordillera Blanca, individual RCM grid boxes in this region cover a

large elevational range and may not accurately represent the values of meteorological variables at the

glacier site. Additionally, errors in the RCM itself, some of which may be inherited from the driving

GCM contributes to the uncertainty in the data and statistically adjusted results of these bias

corrections (Maraun, 2016). The unavailability of usable surface observations of wind speed,

precipitation and shortwave radiation is an obstacle to effective bias correction of the RCM outputs.

Wind speed and shortwave radiation were therefore taken directly from the RCM without bias

correction, precipitation was bias corrected for its dry day frequency and precipitation intensity using

data from the TRMM. These data have been found to accurately capture seasonality in the

precipitation rate of the Peruvian Andes (Malone et al., 2015) while underestimating total annual

precipitation rates observed by Thompson et al. (2013). Evaluation of TRMM in the La Plata Basin of

South America found good agreement between TRMM and gauge precipitation data on monthly time

scales and on the occurrence of daily precipitation events, but comparatively poor agreement on

precipitation intensity estimates at daily scales (Su, Hong and Lettenmaier, 2008). As a result, an

adapted version of the ISI-MIP bias correction approach (Hempel et al., 2013) was applied for

precipitation, in which precipitation was corrected for monthly mean intensity and dry day frequency,

whereas the fitting of a transfer function to the precipitation intensity of wet days was left out. This

could lead to biases remaining in the precipitation data driving the glacier model. Bias correction of

TRMM data with surface precipitation observations would allow the full ISI-MIP bias correction

methodology to be applied to RCM precipitation.

Surface temperature values from the RCM were bias corrected using a variance-scaling methodology

for lapse rate-adjusted observational data. This dataset was incomplete for parts of the time series.

Justified by the assumption of normally distributed daily mean temperature values within individual

months, months with greater than five days of temperature observations were included within the

reference period for bias correction.

Page 31: Melt rate of Palcaraju glacier, Cordillera Blanca, Peru ...applicable to ‘slow-onset’ impacts, such as glacial melt, in which climate change can be shown to materially affect the

31

The bias correction methods assume that the temperature precipitation bias within and outside of the

reference period (for which there are observations) of the actual run is constant, and that the

correction function remains valid for the natural RCM run. Accordingly it is unable to correct for any

errors in the natural RCM run which are not present in the actual data (Maraun et al., 2010).

Furthermore, the effectiveness of statistical bias correction in aligning the RCM values with

observations is dependent on the quality of the observational data set, deficiencies of which have

been noted previously.

7.2 Glacier model

The distributed surface energy and mass balance model devised for this paper, as described above,

provides parameterisations for all significant processes contributing to glacier melt rate. Following

protocols which have been demonstrated to be valid for high alpine glaciers, certain factors which

might contribute to surface energy balance and melt rate were excluded from the model. A high sky

view factor was assumed for the model, and so topographic shading was excluded (e.g. Oerlemans,

1992). The glacier’s orientation (facing a valley) and low-latitude location (i.e. the sun is high in the sky

throughout the year) suggests that little topographic shading of the ablation zone is likely. As a result

of this, diffuse and direct shortwave radiation incident on the glacier surface were taken together

directly from the RCM. Separating diffuse and direct radiation, as would have been required if shaded

grid cells received only diffuse and shortwave radiation (and light reflected from surrounding terrain)

would have introduced additional uncertainties, likely of greater magnitude than the exclusion of this

factor from the melt model.

Additionally, as is the case for the model used by Sicart et al. (2011) for modelling the Zongo glacier in

Bolivia, the melt model is not able to account for subsurface meltwater production. The exclusion of

sub-surface energy fluxes increase the melt rate sensitivity to albedo (Sicart et al., 2011). Although

subsurface meltwater has a very high likelihood of refreezing within the glacier, studies on a Tibetan

glacier have suggested that its exclusion can increase mass balance by as much as 21% (although as

mass balance is comprised of both accumulation and ablation, this represents a smaller proportion of

total melt from the glacier) (Mölg et al., 2012). The lower mean temperatures in the natural scenario

suggests that a greater proportion of subsurface melt would be refrozen. Including subsurface

meltwater production would, therefore, most likely increase the difference in melt rate between the

natural and actual runs. Furthermore, a widely acknowledged source of uncertainty in glacier

modelling is in the determination of turbulent energy fluxes (thought to be much larger than often

modelled) and surface albedo (Hock, 2005). Parameterisations for latent and sensible heat fluxes used

Page 32: Melt rate of Palcaraju glacier, Cordillera Blanca, Peru ...applicable to ‘slow-onset’ impacts, such as glacial melt, in which climate change can be shown to materially affect the

32

ensemble mean RCM output wind speed and air temperature data. More accurate parameterisations

require observations of humidity for calculating latent heat flux. This would reduce latent heat flux in

the wet season and significantly increase its magnitude in the dry season, a variation not captured by

this model (which uses an exponential approximation of the Clausius-Clapeyron relation to calculate

saturation vapour pressure with air temperature).

Confidence in results is also reduced by the lack of surface measurements of humidity, albedo, glacier

temperature, cloud cover, solar radiation, and melt rates, which are a significant obstacle to model

development and validation. For this reason, and as noted previously, remote glacier modelling

(without any surface observations) is highly unusual, and it was therefore necessary to source

parameterisations from several sources for the glacier model. The lack of measurements at the glacier

surface reduced the model sensitivity to short term variations in latent and sensible heat fluxes (in

particular), which can make large contributions to melt rate over short time scales, particularly in the

dry season of May to September.

However, given the motivation to undertake this work, in the context of ongoing legal proceedings

concerning the melt rate of this glacier, this assessment of the sign and approximate magnitude of the

change in melt rate in response to anthropogenic climate forcing was considered appropriate. Current

evidence provided to the court by plaintiffs notes the IPCC’s observation of glacial retreat on the global

and regional scale. This work aligns the impact of climate change on Palcaraju glacier with the global

trend of increasing glacial melt rates, and despite the uncertainties, a clear anthropogenic signal in

melt rate is detected. The assumptions made in constructing this model are consistent with those used

in previously validated glacier models (e.g. Arnold et al., 1996; Hock and Holmgren, 2005; Gurgiser,

Marzeion, et al., 2013), in almost all cases, for tropical glaciers. Nonetheless, there would be

considerable benefit to validating these results and improving the parameterisations using surface

observations. Further research to quantify the rate of increase of depth of Lake Palcacocha

attributable to anthropogenic climate forcing would need to include hydrological modelling of the lake

(including water outflow and evaporation from the lake surface). This would better constrain the

advance in the time at which point it becomes necessary to implement adaptation measures for

Huaraz. This is required if the cost attributable to climate change is to be calculated, as explained

below.

Page 33: Melt rate of Palcaraju glacier, Cordillera Blanca, Peru ...applicable to ‘slow-onset’ impacts, such as glacial melt, in which climate change can be shown to materially affect the

33

8 Discussion of legal context

Many natural disasters occur independently of human alteration of atmospheric GHG concentrations,

but growing numbers of events are demonstrably not ‘acts of god’ and can be attributed to

anthropogenic climate change (Stone and Allen, 2005; Thompson and Otto, 2015). Furthermore, there

is growing acknowledgement in legal and scientific circles that it may be possible to hold major

emitters liable for the proportion of the costs of events attributable to climate change for which they

are responsible (Allen, 2003; Allen and Lord, 2004; Simlinger and Mayer, 2019). As a result, growing

numbers of legal cases seeking damages for climate change impacts are being brought against fossil

fuel companies.

In circumstances in which the likelihood of an event happening increases with cumulative GHG

emissions and has at least doubled, it has been argued that increasing the risk of the event occurring

and causing the damage can be considered synonymous (Marjanac, Patton and Thornton, 2017). The

risk of such events occurring is determined by the atmospheric composition and global temperature

at the time of the event’s occurrence. Therefore, risk increases only as a result of anthropogenic

emissions and does not increase with time, even in the absence of emissions. In these cases, litigation

has sought to hold firms liable for the portion of the total cost of damages caused by the given event,

equal to the proportion of historical anthropogenic emissions for which the firms are responsible

(Thornton and Covington, 2016). However, this argument is valid only for events for which risk

increases only as a result of emissions, typically meteorological events such as floods and heatwaves.

Non-meteorological slow-onset events for which risk also increases with time even in the absence of

human influence, such as the accumulation of meltwater in a proglacial lake, need to be considered

differently.

Climate change may increase the likelihood of a meteorological event occurring in a given year,

whereas increased glacier melt accelerates the rate at which a proglacial lake is filled and brings

forward the next occurrence of a GLOF. If a GLOF would have occurred in the absence of human

influence on the climate, and if the return period of the event is long or the total number of

occurrences of a GLOF from a lake is limited, it is not accurate to attribute damages for the full cost of

the event to climate change. In the case of GLOFs, the return period is typically of the order of decades-

to-centuries, and erosion of terminal moraine may limit the total number of occurrences. Even if the

time to recharge the lake following a flood is reduced as a result of anthropogenic GHG emissions, it

is unlikely that the number of occurrences of the event will have increased over socially or

economically relevant time periods (for example, an additional occurrence of a 1/100-year GLOF

Page 34: Melt rate of Palcaraju glacier, Cordillera Blanca, Peru ...applicable to ‘slow-onset’ impacts, such as glacial melt, in which climate change can be shown to materially affect the

34

reduced to 1/80 by increased melt rates appears only once every 400 years). Rather, the period of

time before an upcoming event may have been reduced.

Further, unlike meteorological events, deterministic attribution to climate change and therefore the

collective action of emitters of long-lived GHGs of an increased rate of glacier meltwater production is

possible. This avoids a common obstacle to the success of litigation: whether an increase in the risk of

an event occurring can be considered sufficient to attribute damages to emitters. In the case of glacier

melt, attribution of changed melt rate may be demonstrable; but how to calculate the costs associated

with this is hitherto undiscussed in climate litigation literature and appears to have few parallels in

legal history (Simlinger and Mayer, 2019).

It can therefore be argued that the harm of a GLOF attributable to climate change is the additional

costs arising from the event occurring sooner than it otherwise would have done. One option,

consistent with mainstream economic practice in the climate damage literature (e.g. Markandya and

González-Eguino, 2019) is for the cost attributable to climate change to be calculated as the difference

between the present-day cost of, in the case of Huaraz, risk reduction (adaptation), and the discounted

present value of the natural event, that is, the discounted present value of the future costs that would

be incurred in the absence of human GHG emissions to obtain the same level of risk reduction. The

discounted present value is calculated as the cost divided by one plus the discount rate, raised to

power of the number of years forward in time the event has been moved as a result of anthropogenic

emissions, as shown in equation 30:

𝐶𝑎𝑛𝑡ℎ = 𝐶𝑎𝑙𝑙 − 𝐶𝑛𝑎𝑡 = 𝐶𝑎𝑙𝑙 −𝐶𝑎𝑙𝑙

(1 + 𝑑)𝑦= 𝐶𝑎𝑙𝑙 (1 −

1

(1 + 𝑑)𝑦) (30)

Equation 30: The cost of slow-onset damages attributable to climate change for events for which risk

is cumulative with both time and emissions. 𝐶𝑎𝑛𝑡ℎ is the cost attributable to anthropogenic GHG

emissions, 𝐶𝑎𝑙𝑙 is the total current cost of the event as a result of climate change, 𝐶𝑛𝑎𝑡 is the discounted

cost of the event occurring when it would have done in the absence of climate change, 𝑑 is the discount

rate, and 𝑦 is the number of years forward in time the event has been moved.

Page 35: Melt rate of Palcaraju glacier, Cordillera Blanca, Peru ...applicable to ‘slow-onset’ impacts, such as glacial melt, in which climate change can be shown to materially affect the

35

Figure 12: Schematic representation of the change in discounted present value of adaptation costs (left

axis) and costs attributable to anthropogenic GHG emissions (right axis) over a 30-year period for

Huaraz. Horizontal lines show the estimated cost attributable to anthropogenic climate forcing based

on the glacier model results under the three discount rates used in policy or proposed by prominent

authors: 1.4% (the average discount rate over the 200-year modelling period in the Stern Review on

the Economics of Climate Change (there is no single discount rate used in the Review since a separate

social discount rate is applied to possible future states of the world due to the uncertainties around

climate change impacts; Dietz, 2008; Stern et al., 2006)), 3.8% (UK Government, 2018) , and 6%

(Nordhaus, 2007, p. 690). This demonstrates how, for a slow-onset event whose timing is brought

forward by anthropogenic influence on climate, the use of a higher discount rate increases the present

value attributable to human influence, in contrast to the social cost of carbon.

Figure 12 demonstrates the importance of the discount rate (the rate of fall of ‘the value of an increase

in consumption at a time in the future relative to now’ (Stern et al., 2006)) for the calculation of the

costs of non-meteorological slow-onset events to climate change. In calculations of the social cost of

carbon (the marginal harm done by emitting one tonne of carbon dioxide) higher discount rates reduce

the costs attributable to emitters (Allen, 2016). By contrast, since higher discount rates ascribe a lower

discounted present value to future costs, it is under higher discount rates that the costs attributable

to climate change are larger according to this methodology, for a given damage. Therefore, as shown

by figure 12, if anthropogenic GHG emissions are responsible for advancing by 30 years the time at

which point adaptation costs of $4 million are required of the city of Huaraz, the cost attributable to

Page 36: Melt rate of Palcaraju glacier, Cordillera Blanca, Peru ...applicable to ‘slow-onset’ impacts, such as glacial melt, in which climate change can be shown to materially affect the

36

climate change is $3.304 million under a 6% discount rate, but only $1.364 million given a 1.4%

discount rate (compared to a present cost of $4 million).

For GLOFs, it is also inappropriate to divide the cost attributable to climate change among emitters in

proportion to their historic contribution to cumulative GHG emissions. Rather, the cost of damages

attributable to any one emitter is dependent on the proportion of the cost of moving the event

forward in time (𝐶𝑎𝑛𝑡ℎ) for which their emissions are responsible, and this depends on the length of

time over which an emitter’s historical emissions have contributed to GLOF risk. Carbon dioxide

emissions released in 1950, for example, will have contributed to the accumulation of meltwater in

Lake Palcacocha (and therefore the increase in GLOF risk) for 50 years longer than emissions produced

in 2000. Consequently, the age of a firm’s emissions is a significant factor in calculating their

contribution to the cost attributable to anthropogenic GHG emissions, as shown by equation 31.

Further research is necessary to calculate the accumulation of meltwater attributable to a pulse

emission of carbon dioxide over time, and therefore the time-dependent contribution of emissions of

different ages to GLOF risk. It is noted that the increase in melt rate in response to a pulse CO2 emission

is unlikely to be constant with time as melt rate increases non-linearly with emissions (as

demonstrated in this paper, and as is the case for numerous impacts of climate change (Harrington

and Otto, in press.)). Therefore, the contribution of the pulse emission to melt rate will change as

atmospheric CO2 levels increase.

𝐶𝑥 = 𝐶𝑎𝑙𝑙 (1 −1

(1 + 𝑑)𝑦) ×

𝐸𝑥

𝐸𝑎𝑙𝑙× 𝑤𝑥 (31)

Equation 31: The cost of damages attributable to an individual emitter for events for which risk of

occurrence is cumulative with time and emissions. 𝐶𝑥 is the cost attributable to emitter 𝑥, 𝐶𝑎𝑙𝑙 is the

total cost of damages, 𝑑 is the discount rate, 𝑦 is the number of years forward in time the event has

moved, 𝐸𝑥 is the historical emissions of firm 𝑥, 𝐸𝑎𝑙𝑙 is cumulative anthropogenic GHG emissions, and

𝑤𝑥 is a weighting factor which accounts for the time at which the emissions of the firm were produced

and therefore their cumulative contribution to the risk of the event occurring.

GLOF risk is dependent, at least in part, on lake volume and therefore cumulative meltwater

production. Elevated meltwater production in response to climate change may increase the rate of

proglacial lake expansion, bringing forward the timing of an outburst flood (or the time at which

adaptation measures are required to be implemented). If meltwater influx to a proglacial lake exceeds

lake drainage and evaporation in a natural climate, then damages sought from emitters should be

calculated according to this equation. If, however, it can be demonstrated that a GLOF would not have

Page 37: Melt rate of Palcaraju glacier, Cordillera Blanca, Peru ...applicable to ‘slow-onset’ impacts, such as glacial melt, in which climate change can be shown to materially affect the

37

happened in the absence of human enhancement of meltwater production through GHG emissions,

litigation could reasonably seek the full cost of the damages of the GLOF or of the cost of adaptation

to heightened GLOF risk (in effect, 𝑦 would be infinite in equations 30 and 31).

8.1 Implications for climate litigation

In Lliuya v. RWE, the claimant seeks to recover damages from RWE AG ‘proportionate to [RWE’s] level

of impairment (share of global GHG emissions) to cover the expenses for appropriate safety

precautions in favour of the claimant’s property from a glacial lake outburst flood from Lake

Palcacocha’ (Rechtsanwälte Günther, 2015). The validity of this claim is dependent on meltwater

production in the absence of anthropogenic emissions being insufficient to trigger a GLOF from Lake

Palcacocha. If this is not the case, the influence of climate change on the timing of the event provides

the justification for claiming of damages. The proposed formula (equation 31) using the difference

between realised costs and their discounted present value (in the absence of climate change) may

provide a useful means to help the courts accurately assess the quantum of damages. Furthermore,

the formula draws attention to the time dependency of the contribution of emissions to GLOF risk:

older emissions will have a greater cumulative contribution to meltwater, and the value of damages

attributed to individual emitters must take this into account.

It is important to stress that equations 30 and 31 are new expressions, developed for this study, with

a much broader application than GLOF risk alone: it would also apply, for example, to liability for the

costs of mitigating the risks of sea level rise in a location in which sea level is rising naturally due to

isostatic rebound. While it is consistent with conventional economic thinking regarding the calculation

of the present value of climate damages, it clearly merits further discussion with stakeholders and

alternative methods of attributing costs to impacts of this type might emerge.

The need to assess discounted future damages adds an additional layer of complexity to climate

litigation. In the case of GLOFs, scientific questions on attribution to climate change, economic analysis

of costs and ethical considerations (including time preferences and concerns about intergenerational

equity (e.g. Stern et al., 2006; Beckerman and Hepburn, 2007)) to determine the discount rate may all

need to be accepted by the courts, if litigation is to succeed. As this paper demonstrates, the impacts

of climate change are increasingly quantifiable by the scientific community. The question remains,

however, is whether in practice the courts will find emitters liable for damages in the face of this multi-

disciplinary complexity.

Page 38: Melt rate of Palcaraju glacier, Cordillera Blanca, Peru ...applicable to ‘slow-onset’ impacts, such as glacial melt, in which climate change can be shown to materially affect the

38

9 Conclusion

A strong anthropogenic signal is identified in the melt rate of Palcaraju glacier based on the output of

a regional climate model and an idealised glacier model. Below 5000m, the modelled melt rate for the

current climate exceeds the natural rate throughout. As discussed in section 6, there are a number of

uncertainties in modelling of a glacier in the absence of surface observations, but all assumptions used

in the glacier model developed for this paper have been used by other authors for alpine glaciers and

demonstrated to be valid.

Although the modelled melt rates under natural conditions are comparatively low, these are still likely

to be sufficient to fill Lake Palcacocha over time. It is clear, however, that the elevated melt rate in

response to anthropogenic GHG emissions will fill the lake much more rapidly. Uncertainties act to

both increase and decrease the meltwater production attributable to human influence on the climate

system: higher mean temperatures in the actual scenario would, for instance, reduce subsurface

refreezing of meltwater, but likely increase the evaporated fraction of ablation. Consequently, the lake

will reach a critical volume, at which point costly adaptation measures will be required to mitigate

GLOF risk, much sooner than would have been the case in the absence of human influence on the

climate system. While the findings of the present study are too uncertain (particularly in the absence

of direct in situ measurements) to determine precisely the proportion of adaptation costs attributable

to an individual emitter, this study does provide the basis for further research which, if undertaken,

would allow this question more accurately. Further research should be focussed in the following four

areas:

i. Further development of the glacier model to include a glacier flow model, and validation and

tuning of model parameterisations with surface observations, particularly of key surface fluxes

over this glacier.

i. Improvements to RCM bias correction with observations of surface downwelling shortwave

solar radiation, wind speed, temperature, and precipitation at locations close to the glacier.

Alternatively, if resources permit, a much higher-resolution non-hydrostatic model would

provide much more reliable information about attributable changes in precipitation, in

particular.

ii. Hydrological modelling of Lake Palcacocha to quantify outflow and evaporation and therefore

the rate at which the lake fills under the actual and natural climate scenarios.

iii. Age of the defendant company’s emissions to calculate cumulative contribution to meltwater

production.

Page 39: Melt rate of Palcaraju glacier, Cordillera Blanca, Peru ...applicable to ‘slow-onset’ impacts, such as glacial melt, in which climate change can be shown to materially affect the

39

With improvements to the glacier model and modelling of lake hydrology it would be possible to

calculate how much sooner the lake will reach a sufficient depth for a GLOF to occur than would have

been the case in a natural climate. For slow-onset climate change impacts, conventional approaches

to climate change litigation may make incorrect assumptions about the responsibility of emitters for

damages. Rather than attributing the entire cost of an event to human GHG emissions, an alternative

methodology for calculating emitters’ liability is proposed. In addition to how much sooner an event

occurs as a result of human influence on the climate system, the discount rate used is also a significant

control on the costs that can be recovered by plaintiffs. Higher discount rates decrease the discounted

present value of events under counterfactual conditions to a greater extent, and therefore attribute a

greater proportion of the current costs to GHG emissions.

Although devastating when they occur, GLOFs are rare and represent only a minor component of the

impacts of climate change on human societies. However, it is possible that this work could have

implications for climate litigation concerning other climate impacts, most notably sea level rise.

Further research is required to assess the time-dependency of the contribution of emitters to sea level

rise, but it is possible that a similar methodology may be required if sea level rise litigation is to align

itself with the best available science and if the contribution of emitters to these impacts is to be

correctly quantified.

10 References

Allen, M. R. (2003) ‘Liability for climate change’, Nature, 421, pp. 891–892.

Allen, M. R. (2016) ‘Drivers of peak warming in a consumption-maximizing world’, Nature Climate

Change. Nature Publishing Group, 6(7), pp. 684–686. doi: 10.1038/nclimate2977.

Allen, M. R. and Lord, R. (2004) ‘The blame game’, Nature, 432(7017), pp. 551–552. doi:

10.1038/432551a.

Arnold, N. S. et al. (1996) ‘A distributed surface energy-balance model for a small valley glacier. I.

Development and testing for Haut Glacier d’ Arolla, Valais, Switzerland’, Journal of Glaciology.

Cambridge University Press, 42(140), pp. 77–89. doi: 10.3189/S0022143000030549.

Autoridad Nacional del Agua (2014) Inventario de Glaciares y Lagunas. Lima. Available at:

www.ana.gob.pe (Accessed: 30 September 2018).

Banwell, A. F. et al. (2012) ‘Calibration and evaluation of a high-resolution surface mass-balance model

for Paakitsoq, West Greenland’, Journal of Glaciology. Cambridge University Press, 58(212), pp. 1047–

Page 40: Melt rate of Palcaraju glacier, Cordillera Blanca, Peru ...applicable to ‘slow-onset’ impacts, such as glacial melt, in which climate change can be shown to materially affect the

40

1062. doi: 10.3189/2012JoG12J034.

Banwell, A. F. et al. (2016) ‘Moulin density controls drainage development beneath the Greenland ice

sheet’, Journal of Geophysical Research: Earth Surface. Wiley-Blackwell, 121(12), pp. 2248–2269. doi:

10.1002/2015JF003801.

Beckerman, W. and Hepburn, C. J. (2007) ‘Ethics of the Discount Rate in the Stern Review on the

Economics of Climate Change’, World Economics, 8(1), pp. 187–210. Available at:

http://content.ebscohost.com/ContentServer.asp?T=P&P=AN&K=24957836&S=R&D=bth&EbscoCon

tent=dGJyMNLr40SeprI40dvuOLCmr1CeprRSr624TbKWxWXS&ContentCustomer=dGJyMPGus1CuqK

9RuePfgeyx44Dt6fIA (Accessed: 4 September 2018).

Bougamont, M., Bamber, J. L. and Greuell, W. (2005) ‘A surface mass balance model for the Greenland

Ice Sheet’, Journal of Geophysical Research: Earth Surface. Wiley-Blackwell, 110(F04018). doi:

10.1029/2005JF000348.

Caidong, C. and Sorteberg, A. (2010) ‘Modelled mass balance of Xibu glacier, Tibetan Plateau:

sensitivity to climate change’, Journal of Glaciology. Cambridge University Press, 56(196), pp. 235–

248. doi: 10.3189/002214310791968467.

Carey, M. et al. (2012) ‘An integrated socio-environmental framework for glacier hazard management

and climate change adaptation: Lessons from Lake 513, Cordillera Blanca, Peru’, Climatic Change,

112(3–4), pp. 733–767. doi: 10.1007/s10584-011-0249-8.

Chen, J., Brissette, F. P. and Leconte, R. (2011) ‘Uncertainty of downscaling method in quantifying the

impact of climate change on hydrology’, Journal of Hydrology. Elsevier, 401(3–4), pp. 190–202. doi:

10.1016/j.jhydrol.2011.02.020.

Dietz, S. (2008) A long-run target for climate policy: the Stern Review and its critics, Report for the

Committee on Climate Change. Available at: http://personal.lse.ac.uk/dietzs/A long-run target for

climate policy - the Stern Review and its critics.pdf (Accessed: 7 January 2019).

Emmer, A. and Vilímek, V. (2013) ‘Review Article: Lake and breach hazard assessment for moraine-

dammed lakes: an example from the Cordillera Blanca (Peru)’, Natural Hazards and Earth System

Science. Copernicus GmbH, 13(6), pp. 1551–1565. doi: 10.5194/nhess-13-1551-2013.

Etchevers, P. et al. (2004) ‘Validation of the energy budget of an alpine snowpack simulated by several

snow models (Snow MIP project)’, Annals of Glaciology. Cambridge University Press, 38, pp. 150–158.

doi: 10.3189/172756404781814825.

Page 41: Melt rate of Palcaraju glacier, Cordillera Blanca, Peru ...applicable to ‘slow-onset’ impacts, such as glacial melt, in which climate change can be shown to materially affect the

41

Evans, S. G. and Clague, J. J. (1994) ‘Recent climatic change and catastrophic geomorphic processes in

mountain environments’, Geomorphology. Elsevier BV, 10(1–4), pp. 107–128. doi: 10.1016/0169-

555x(94)90011-6.

Favier, V. et al. (2004) ‘One-year measurements of surface heat budget on the ablation zone of

Antizana Glacier 15, Ecuadorian Andes’, Journal of Geophysical Research Atmospheres. Wiley-

Blackwell, 109(18), p. D18105. doi: 10.1029/2003JD004359.

Frank, W., Bals, C. and Grimm, J. (2019) ‘The Case of Huaraz: First Climate Lawsuit on Loss and Damage

Against an Energy Company Before German Courts’, in Mechler, R. et al. (eds) Loss and Damage from

Climate Change. Cham: Springer, pp. 475–482. doi: 10.1007/978-3-319-72026-5_20.

Fujita, K. and Ageta, Y. (2000) ‘Effect of summer accumulation on glacier mass balance on the Tibetan

Plateau revealed by mass-balance model’, Journal of Glaciology. Cambridge University Press, 46(153),

pp. 244–252. doi: 10.3189/172756500781832945.

Goddard Earth Sciences Data and Information Services Center (2016) TRMM (TMPA-RT) Near Real-

Time Precipitation L3 1 day 0.25 degree x 0.25 degree V7, Edited by Andrey Savtchenko, Greenbelt,

MD, Goddard Earth Sciences Data and Information Services Center (GES DISC). doi:

10.5067/TRMM/TMPA/DAY-E/7.

Griggs, M. (1968) ‘Emissivities of natural surfaces in the 8-to14 Micron spectral region’, Journal of

Geophysical Research. Wiley-Blackwell, 73(24), pp. 7545–7551. doi: doi:10.1029/JB073i024p07545.

Guillod, B. P. et al. (2017) ‘weather@home 2: validation of an improved global–regional climate

modelling system’, Geoscientific Model Development, 10(5), pp. 1849–1872. doi: 10.5194/gmd-10-

1849-2017.

Gurgiser, W., Molg, T., et al. (2013) ‘Mass-balance model parameter transferability on a tropical

glacier’, Journal of Glaciology. Cambridge University Press, 59(217), pp. 845–858. doi:

10.3189/2013JoG12J226.

Gurgiser, W., Marzeion, B., et al. (2013) ‘Modeling energy and mass balance of Shallap Glacier, Peru’,

The Cryosphere, 7, pp. 1787–1802. doi: 10.5194/tc-7-1787-2013.

Harrington, L. J. and Otto, F. E. L. (no date) ‘Attributable damage liability in a non-linear climate’.

Hempel, S. et al. (2013) ‘A trend-preserving bias correction – the ISI-MIP approach’, Earth System

Dynamics, 4, pp. 219–236. doi: 10.5194/esd-4-219-2013.

Page 42: Melt rate of Palcaraju glacier, Cordillera Blanca, Peru ...applicable to ‘slow-onset’ impacts, such as glacial melt, in which climate change can be shown to materially affect the

42

Hock, R. (2005) ‘Glacier melt: a review of processes and their modelling’, Progress in Physical

Geography, 29(3), pp. 362–391. doi: 10.1191/0309133305pp453ra.

Hock, R. and Holmgren, B. (2005) ‘A distributed surface energy-balance model for complex topography

and its application to Storglaciären, Sweden’, Journal of Glaciology. Cambridge University Press,

51(172), pp. 25–36. doi: 10.3189/172756505781829566.

Jiang, X. et al. (2010) ‘A distributed surface energy and mass balance model and its application to a

mountain glacier in China’, Chinese Science Bulletin. SP Science China Press, 55(20), pp. 2079–2087.

doi: 10.1007/s11434-010-3068-9.

Kaser, G. (1999) ‘A review of the modern fluctuations of tropical glaciers’, Global and Planetary

Change. Elsevier BV, 22(1–4), pp. 93–103. doi: 10.1016/s0921-8181(99)00028-4.

Kaser, G., Ames, A. and Zamora, M. (1990) ‘Glacier Fluctuations and Climate in the Cordillera Blanca,

Peru’, Annals of Glaciology. Cambridge University Press (CUP), 14, pp. 136–140. doi:

10.3189/s0260305500008430.

Kraus, H. (1973) ‘Energy exchange at air-ice interface’, in The Role of Snow and Ice in Hydrology. (Proc.

Banff. Symp., September 1972), pp. 128–164. Available at: https://iahs.info/uploads/dms/107013.pdf

(Accessed: 8 December 2018).

Malone, A. G. O. et al. (2015) ‘Constraints on southern hemisphere tropical climate change during the

Little Ice Age and Younger Dryas based on glacier modeling of the Quelccaya Ice Cap, Peru’, Quaternary

Science Reviews, 125, pp. 106–116. doi: 10.1016/j.quascirev.2015.08.001.

Maraun, D. et al. (2010) ‘Precipitation downscaling under climate change: Recent developments to

bridge the gap between dynamical models and the end user’, Reviews of Geophysics. Wiley-Blackwell,

48(3), p. RG3003. doi: 10.1029/2009RG000314.

Maraun, D. (2012) ‘Nonstationarities of regional climate model biases in European seasonal mean

temperature and precipitation sums’, Geophysical Research Letters. Wiley-Blackwell, 39(6). doi:

10.1029/2012GL051210.

Maraun, D. (2016) ‘Bias Correcting Climate Change Simulations - a Critical Review’, Current Climate

Change Reports. Springer International Publishing, 2(4), pp. 211–220. doi: 10.1007/s40641-016-0050-

x.

Marjanac, S., Patton, L. and Thornton, J. (2017) ‘Acts of God, human influence and litigation’, Nature

Geoscience, 10(9), pp. 616–619. doi: 10.1038/ngeo3019.

Page 43: Melt rate of Palcaraju glacier, Cordillera Blanca, Peru ...applicable to ‘slow-onset’ impacts, such as glacial melt, in which climate change can be shown to materially affect the

43

Markandya, A. and González-Eguino, M. (2019) ‘Integrated Assessment for Identifying Climate Finance

Needs for Loss and Damage: A Critical Review’, in Mechler, R. et al. (eds) Loss and Damage from

Climate Change. Cham: Springer, pp. 343–362. doi: https://doi.org/10.1007/978-3-319-72026-5_14

343.

Marshall, S. J. (2007) ‘Modelling Glacier Response to Climate Change’, in Glacier Science and

Environmental Change. Malden, MA, USA: Blackwell Publishing, pp. 163–173. doi:

10.1002/9780470750636.ch32.

Marzeion, B. et al. (2014) ‘Attribution of global glacier mass loss to anthropogenic and natural causes’,

Science. American Association for the Advancement of Science (AAAS), 345(6199), pp. 919–921. doi:

10.1126/science.1254702.

Massey, N. R. et al. (2015) ‘weather@home-development and validation of a very large ensemble

modelling system for probabilistic event attribution’, Quarterly Journal of the Royal Meteorological

Society. Wiley-Blackwell, 141, pp. 1528–1545. doi: 10.1002/qj.2455.

McGuire, B. (2012) ‘Hazardous Responses of the Solid Earth to a Changing Climate’, Climate Forcing of

Geological Hazards. John Wiley & Sons, Ltd, pp. 1–33. doi: 10.1002/9781118482698.ch1.

Menne, M. J. et al. (2012) ‘An Overview of the Global Historical Climatology Network-Daily Database’,

Journal of Atmospheric and Oceanic Technology, 29(7), pp. 897–910. doi: 10.1175/JTECH-D-11-

00103.1.

Mölg, T. et al. (2012) ‘The footprint of Asian monsoon dynamics in the mass and energy balance of a

Tibetan glacier’, The Cryosphere, 6, pp. 1445–1461. doi: 10.5194/tc-6-1445-2012.

Mölg, T. and Hardy, D. R. (2004) ‘Ablation and associated energy balance of a horizontal glacier surface

on Kilimanjaro’, Journal of Geophysical Research D: Atmospheres. Wiley-Blackwell, 109(16), p. D16104.

doi: 10.1029/2003JD004338.

Nordhaus, W. D. (2007) ‘A Review of the Stern Review on the Economics of Climate Change’, Journal

of Economic Literature, XLV, pp. 686–702. Available at:

https://pubs.aeaweb.org/doi/pdfplus/10.1257/jel.45.3.686 (Accessed: 4 September 2018).

Ochoa, A. et al. (2014) ‘Evaluation of TRMM 3B42 precipitation estimates and WRF retrospective

precipitation simulation over the Pacific-Andean region of Ecuador and Peru’, Hydrol. Earth Syst. Sci,

18, pp. 3179–3193. doi: 10.5194/hess-18-3179-2014.

Oerlemans, J. (1991) ‘The mass balance of the Greenland ice sheet: sensitivity to climate change as

Page 44: Melt rate of Palcaraju glacier, Cordillera Blanca, Peru ...applicable to ‘slow-onset’ impacts, such as glacial melt, in which climate change can be shown to materially affect the

44

revealed by energy-balance modelling’, The Holocene, 1, pp. 40–48. doi:

10.1177/095968369100100106.

Oerlemans, J. (1992) ‘Climate sensitivity of glaciers in southern Norway: application of an energy-

balance model to Nigardsbreen, Hellstugubreen and Alfotbreen’, Journal of Glaciology. Cambridge

University Press, 38(129), pp. 223–232. doi: 10.3189/S0022143000003634.

Oerlemans, J. (1993) ‘A model for the surface balance of ice masses: Part I: alpine glaciers’, Zeitschrift

für Gletscherkunde und Glazialgeologie, 27, pp. 63–83.

Oerlemans, J. and Knap, W. H. (1998) ‘A 1 year record of global radiation and albedo in the ablation

zone of Morteratschgletscher, Switzerland’, Journal of Glaciology. Cambridge University Press,

44(147), pp. 231–238. doi: 10.3189/S0022143000002574.

Pfeffer, W. T. et al. (2014) ‘The Randolph Glacier Inventory: a globally complete inventory of glaciers’,

Journal of Glaciology. Cambridge University Press, 60(221), pp. 537–552. doi:

10.3189/2014JoG13J176.

Portocarrero Rodríguez, C. A. and Engility Corporation (2014) The Glacial Lake Handbook: reducing risk

from dangerous glacial lakes in the Cordillera Blanca, Peru. Washington D.C. Available at:

https://pdf.usaid.gov/pdf_docs/PBAAA087.pdf (Accessed: 30 September 2018).

Prinz, R. et al. (2016) ‘Climatic controls and climate proxy potential of Lewis Glacier, Mt. Kenya’, The

Cryosphere, 10, pp. 133–148. doi: 10.5194/tc-10-133-2016.

Rabatel, A. et al. (2013) ‘Current state of glaciers in the tropical Andes: a multi-century perspective on

glacier evolution and climate change’, The Cryosphere. Copernicus GmbH, 7(1), pp. 81–102. doi:

10.5194/tc-7-81-2013.

Rauscher, S. A. et al. (2010) ‘Resolution effects on regional climate model simulations of seasonal

precipitation over Europe’, Climate Dynamics. Springer-Verlag, 35, pp. 685–711. doi: 10.1007/s00382-

009-0607-7.

Rechtsanwälte Günther (2015) Lliuya v RWE AG. Available at:

https://germanwatch.org/en/download/20822.pdf.

RGI Consortium (2017) Randolph Glacier Inventory–A Dataset of Global Glacier Outlines: Version 6.0:

Technical Report, Global Land Ice Measurements from Space, Digital Media. https://doi. org/10.7265.

Colorado, USA. doi: https://doi.org/10.7265.

Page 45: Melt rate of Palcaraju glacier, Cordillera Blanca, Peru ...applicable to ‘slow-onset’ impacts, such as glacial melt, in which climate change can be shown to materially affect the

45

Rupper, S. and Roe, G. H. (2008) ‘Glacier changes and regional climate: A mass and energy balance

approach’, Journal of Climate, 21(20), pp. 5384–5401. doi: 10.1175/2008JCLI2219.1.

Rye, C. J. et al. (2010) ‘Modeling the surface mass balance of a high Arctic glacier using the ERA-40

reanalysis’, Journal of Geophysical Research: Earth Surface. Wiley-Blackwell, 115(F2). doi:

10.1029/2009JF001364.

Sagredo, E. A., Rupper, S. and Lowell, T. V. (2014) ‘Sensitivities of the equilibrium line altitude to

temperature and precipitation changes along the Andes’, Quaternary Research. Cambridge University

Press, 81(02), pp. 355–366. doi: 10.1016/j.yqres.2014.01.008.

Schauwecker, S. et al. (2014) ‘Climate trends and glacier retreat in the Cordillera Blanca, Peru,

revisited’, Global and Planetary Change, 119, pp. 85–97. doi: 10.1016/j.gloplacha.2014.05.005.

Scheel, M. L. M. et al. (2011) ‘Evaluation of TRMM Multi-satellite Precipitation Analysis (TMPA)

performance in the Central Andes region and its dependency on spatial and temporal resolution’,

Hydrology and Earth System Sciences, 15(8), pp. 2649–2663. doi: 10.5194/hess-15-2649-2011.

Schmidli, J., Frei, C. and Vidale, P. L. (2006) ‘Downscaling from GCM precipitation: A benchmark for

dynamical and statistical downscaling methods’, International Journal of Climatology. Wiley-Blackwell,

26(5), pp. 679–689. doi: 10.1002/joc.1287.

Sicart, J.-E. et al. (2011) ‘Analysis of seasonal variations in mass balance and meltwater discharge of

the tropical Zongo Glacier by application of a distributed energy balance model’, Journal of

Geophysical Research. Wiley-Blackwell, 116(D13), p. D13105. doi: 10.1029/2010JD015105.

Simlinger, F. and Mayer, B. (2019) ‘Legal Responses to Climate Change Induced Loss and Damage’, in

Mechler, R. et al. (eds) Loss and Damage from Climate Change. Cham: Springer, pp. 179–203. doi:

10.1007/978-3-319-72026-5_20.

Somos-Valenzuela, M. A. et al. (2016) ‘Modeling a glacial lake outburst flood process chain: the case

of Lake Palcacocha and Huaraz, Peru’, Hydrology and Earth System Sciences. Copernicus GmbH, 20(6),

pp. 2519–2543. doi: 10.5194/hess-20-2519-2016.

Stern, N. H. et al. (2006) Stern Review: The Economics of Climate Change. Cambridge, UK: Cambridge

University Press.

Stone, D. A. and Allen, M. R. (2005) ‘The End-to-End Attribution Problem: From Emissions to Impacts’,

Climatic Change. Kluwer Academic Publishers, 71(3), pp. 303–318. doi: 10.1007/s10584-005-6778-2.

Page 46: Melt rate of Palcaraju glacier, Cordillera Blanca, Peru ...applicable to ‘slow-onset’ impacts, such as glacial melt, in which climate change can be shown to materially affect the

46

Su, F., Hong, Y. and Lettenmaier, D. P. (2008) ‘Evaluation of TRMM Multisatellite Precipitation Analysis

(TMPA) and Its Utility in Hydrologic Prediction in the La Plata Basin’, Journal of Hydrometeorology,

9(4), pp. 622–640. doi: 10.1175/2007JHM944.1.

Tarboton, D. G., Chowdhury, T. G. and Jackson, Thomas, H. (1995) ‘A spatially distributed energy

balance snowmelt model’, in Biogeochemistry of Seasonally Snow-Covered Catchments (Proceedings

of a Boulder Symposium, July 1995). IAHS Publ. no. 228, pp. 141–155.

Teutschbein, C. and Seibert, J. (2012) ‘Bias correction of regional climate model simulations for

hydrological climate-change impact studies: Review and evaluation of different methods’, Journal of

Hydrology. Elsevier, 456–457, pp. 12–29. doi: 10.1016/J.JHYDROL.2012.05.052.

Thompson, A. and Otto, F. E. L. (2015) ‘Ethical and normative implications of weather event attribution

for policy discussions concerning loss and damage’, Climatic Change. Springer Netherlands, 133(3), pp.

439–451. doi: 10.1007/s10584-015-1433-z.

Thompson, L. G. et al. (2013) ‘Annually Resolved Ice Core Records of Tropical Climate Variability over

the Past ~1800 Years’, Science. American Association for the Advancement of Science, 340(6135), pp.

945–950. doi: 10.1126/science.269.5220.46.

Thompson, L. G. et al. (2017) ‘Ice Core Records of Climate and Environmental Variability in the Tropical

Andes of Peru: Past Present and Future / Registros en Núcleos de Hielo de la Variabilidad Climática y

Ambiental en los Andes Tropicales del Perú: Pasado, Presente y Futuro’, Revista de Glaciares y

Ecosistemas de Montaña, 3, pp. 25–40.

Thornton, J. and Covington, H. (2016) ‘Climate change before the court’, Nature Geoscience, 9(1), pp.

3–5. doi: 10.1038/ngeo2612.

UK Government (2018) Updated short-term traded carbon values used for modelling purposes.

Available at:

https://assets.publishing.service.gov.uk/government/uploads/system/uploads/attachment_data/file

/671191/Updated_short-term_traded_carbon_values_for_modelling_purposes.pdf (Accessed: 7

January 2019).

Vilímek, V. et al. (2005) ‘Influence of glacial retreat on natural hazards of the Palcacocha Lake area,

Peru’, Landslides. Springer-Verlag, 2(2), pp. 107–115. doi: 10.1007/s10346-005-0052-6.

Warren, S. G. (1982) ‘Optical properties of snow’, Reviews of Geophysics and Space Physics. Wiley-

Blackwell, 20(1), p. 67. doi: 10.1029/RG020i001p00067.

Page 47: Melt rate of Palcaraju glacier, Cordillera Blanca, Peru ...applicable to ‘slow-onset’ impacts, such as glacial melt, in which climate change can be shown to materially affect the

47

Wright, A. P. et al. (2007) ‘Modeling the refreezing of meltwater as superimposed ice on a high Arctic

glacier: A comparison of approaches’, Journal of Geophysical Research. Wiley-Blackwell, 112(F4), p.

F04016. doi: 10.1029/2007JF000818.

Appendix

#--- Vertically Distributed Surface Energy and Mass Balance Model ---#

#--- November 2018 ---#

#--- Import required packages ---#

import numpy as np

import math

import statistics

import matplotlib.pyplot as plt

import matplotlib.cbook as cbook

import matplotlib.cm as cm

from matplotlib.path import Path

from matplotlib.patches import PathPatch

from netCDF4 import Dataset

import scipy.special as special

import glob

from datetime import datetime

from netCDF4 import date2num,num2date

import matplotlib.patches as mpatches

from math import *

from tempfile import TemporaryFile

import datetime as dt

import matplotlib.dates as mdates

#--- Subroutines that input data from excel spreadsheet as matlab arrays ---#

import Topo_Import

import THREE_VALUE_GRAB

#--- Conversion Factors and Constants ---#

secs_per_day = 3600. * 24.

pi = 3.14159

deg_rad = pi/180.

days_per_year = 360.

#--- Parameters ---#

# Lapse Rate

LR = 0.0055 # degrees per km Gurgiser et al 2013

Lap = 0.0065

Tfreeze = 273.15 # freezing temp of H20 (K)

rhow = 1000. # Liquid at 0C

Lf = 334000.0 # Latent heat of fusion of water - Arnold et al 1996

psia = 0.84 # Atmospheric clear-sky transmissivity (Cuffey & Paterson, 2010)

atm0 = 101325.0 # Atmospheric pressure at sea level (??)

g = 9.80665 # earth-surface gravitational acceleration

M = 0.0300644 # Molar mass of dry air

gc = 8.31447 # Ideal gas constant

Kelvin = 273.15

L = 1367.0 # Solar constant (W/m^2) - Varley et al 1996.

e = 0.017 # eccentricity

precession = 0. # precessional angle Rel. to solstaice (degrees)

obliquity = 23.44 # obliquity angle (degrees)

Page 48: Melt rate of Palcaraju glacier, Cordillera Blanca, Peru ...applicable to ‘slow-onset’ impacts, such as glacial melt, in which climate change can be shown to materially affect the

48

latitude = -9.375 # latitude of site (degrees)

precess = precession * deg_rad # precessional angle (radians)

obl = obliquity * deg_rad # obliquity angle (radians)

lat = latitude * deg_rad # latitude (radians)

psi = lat

sigma = 0.0000000567 # Stefan Boltzman constant (Hock, 2005)

alpham = 0.4 # Arnold et al (1996) - albedo of surroundings

# Define constants for longwave terrain radiation:

Lta = 0.77 # Wm**-2sr**-1

Ltb = 0.54 # Wm**-2sr**-1 from Pluss & Ohmura (1997)

Lb = 100.2 # Wm**2sr**-1 Emitted radiance of a 0C black body (Hock & Holmgren, 2005)

etas = 1.0 # Emissivity of snow cover = 1, Hock & Holmgren (2005)

cp = 1004.0 # Gas phase specific heat (J/(kg K)), at 298K and 1 bar

Ks = 0.00000442 # Sensible heat scalar (snow)

K1 = 0.00777 # for evaporating snow.

#K1 = 0.01114 # for evaporating ice

#--- Model definitions ---#

def xTa(nat_Tgl,LR,b, T_step):

yTa = nat_Tgl[T_step] - (LR * (b - 4625.0))

return yTa

def xpd(nat_ppt, T_step):

ypd = nat_ppt[T_step]

return ypd

def xELA(nat_Tgl, LR, T_step):

yELA = 4625.0 + ((nat_Tgl[T_step] - 273.15) / LR)

return yELA

def xs(Ta, pd, T_step, lev, j, k):

if Ta[lev,T_step] < 276.15 and pd[lev,T_step,k,j] > 10.0:

ys = T_step

elif Ta[lev,T_step-1] < 276.15 and pd[lev,T_step-1,k,j] > 20.0 and T_step > 0:

ys = T_step - 1

elif Ta[lev,T_step-2] < 276.15 and pd[lev,T_step-2,k,j] > 20.0 and T_step > 1:

ys = T_step - 2

elif Ta[lev,T_step-3] < 276.15 and pd[lev,T_step-3,k,j] > 20.0 and T_step > 2:

ys = T_step - 3

elif Ta[lev, T_step-4] < 276.15 and pd[lev,T_step-4,k,j] > 20.0 and T_step > 3:

ys = T_step - 4

elif Ta[lev, T_step-5] < 276.15 and pd[lev,T_step-5,k,j] > 20.0 and T_step > 4:

ys = T_step - 5

elif Ta[lev, T_step-6] < 276.15 and pd[lev,T_step-6,k,j] > 20.0 and T_step > 5:

ys = T_step - 6

elif Ta[lev, T_step-7] < 276.15 and pd[lev,T_step-7,k,j] > 20.0 and T_step > 6:

ys = T_step - 7

elif Ta[lev, T_step-8] < 276.15 and pd[lev,T_step-8,k,j] > 20.0 and T_step > 7:

ys = T_step - 8

elif Ta[lev, T_step-9] < 276.15 and pd[lev,T_step-9,k,j] > 20.0 and T_step > 8:

ys = T_step - 9

elif Ta[lev, T_step-10] < 276.15 and pd[lev,T_step-10,k,j] > 20.0 and T_step > 9:

ys = T_step - 10

elif Ta[lev, T_step-11] < 276.15 and pd[lev,T_step-11,k,j] > 20.0 and T_step > 10:

ys = T_step - 11

elif Ta[lev, T_step-12] < 276.15 and pd[lev,T_step-12,k,j] > 20.0 and T_step > 11:

ys = T_step - 12

elif Ta[lev, T_step-13] < 276.15 and pd[lev,T_step-13,k,j] > 20.0 and T_step > 12:

ys = T_step - 13

elif Ta[lev, T_step-14] < 276.15 and pd[lev,T_step-14,k,j] > 20.0 and T_step > 13:

Page 49: Melt rate of Palcaraju glacier, Cordillera Blanca, Peru ...applicable to ‘slow-onset’ impacts, such as glacial melt, in which climate change can be shown to materially affect the

49

ys = T_step - 14

elif Ta[lev, T_step-15] < 276.15 and pd[lev,T_step-15,k,j] > 20.0 and T_step > 14:

ys = T_step - 15

elif Ta[lev, T_step-16] < 276.15 and pd[lev,T_step-16,k,j] > 20.0 and T_step > 15:

ys = T_step - 16

elif Ta[lev, T_step-17] < 276.15 and pd[lev,T_step-17,k,j] > 20.0 and T_step > 16:

ys = T_step - 17

elif Ta[lev, T_step-18] < 276.15 and pd[lev,T_step-18,k,j] > 20.0 and T_step > 17:

ys = T_step - 18

elif Ta[lev, T_step-19] < 276.15 and pd[lev,T_step-19,k,j] > 20.0 and T_step > 18:

ys = T_step - 19

elif Ta[lev, T_step-20] < 276.15 and pd[lev,T_step-20,k,j] > 20.0 and T_step > 19:

ys = T_step - 20

elif Ta[lev, T_step-21] < 276.15 and pd[lev,T_step-21,k,j] > 20.0 and T_step > 20:

ys = T_step - 21

elif Ta[lev, T_step-22] < 276.15 and pd[lev,T_step-22,k,j] > 20.0 and T_step > 21:

ys = T_step - 22

elif Ta[lev, T_step-23] < 276.15 and pd[lev,T_step-23,k,j] > 20.0 and T_step > 22:

ys = T_step - 23

elif Ta[lev, T_step-24] < 276.15 and pd[lev,T_step-24,k,j] > 20.0 and T_step > 23:

ys = T_step - 24

elif Ta[lev, T_step-25] < 276.15 and pd[lev,T_step-25,k,j] > 20.0 and T_step > 24:

ys = T_step - 25

elif Ta[lev, T_step-26] < 276.15 and pd[lev,T_step-26,k,j] > 20.0 and T_step > 25:

ys = T_step - 26

elif Ta[lev, T_step-27] < 276.15 and pd[lev,T_step-27,k,j] > 20.0 and T_step > 26:

ys = T_step - 27

elif Ta[lev, T_step-28] < 276.15 and pd[lev,T_step-28,k,j] > 20.0 and T_step > 27:

ys = T_step - 28

elif Ta[lev, T_step-29] < 276.15 and pd[lev,T_step-29,k,j] > 20.0 and T_step > 28:

ys = T_step - 29

elif Ta[lev, T_step-30] < 276.15 and pd[lev,T_step-30,k,j] > 20.0 and T_step > 29:

ys = T_step - 30

elif Ta[lev, T_step-31] < 276.15 and pd[lev,T_step-31,k,j] > 20.0 and T_step > 30:

ys = T_step - 31

elif Ta[lev, T_step-32] < 276.15 and pd[lev,T_step-32,k,j] > 20.0 and T_step > 31:

ys = T_step - 32

elif Ta[lev, T_step-33] < 276.15 and pd[lev,T_step-33,k,j] > 20.0 and T_step > 32:

ys = T_step - 33

elif Ta[lev, T_step-34] < 276.15 and pd[lev,T_step-34,k,j] > 20.0 and T_step > 33:

ys = T_step - 34

elif Ta[lev, T_step-35] < 276.15 and pd[lev,T_step-35,k,j] > 20.0 and T_step > 34:

ys = T_step - 35

elif Ta[lev, T_step-36] < 276.15 and pd[lev,T_step-36,k,j] > 20.0 and T_step > 35:

ys = T_step - 36

elif Ta[lev, T_step-37] < 276.15 and pd[lev,T_step-37,k,j] > 20.0 and T_step > 36:

ys = T_step - 37

elif Ta[lev, T_step-38] < 276.15 and pd[lev,T_step-38,k,j] > 20.0 and T_step > 37:

ys = T_step - 38

elif Ta[lev, T_step-39] < 276.15 and pd[lev,T_step-39,k,j] > 20.0 and T_step > 38:

ys = T_step - 39

elif Ta[lev, T_step-40] < 276.15 and pd[lev,T_step-40,k,j] > 20.0 and T_step > 39:

ys = T_step - 40

elif Ta[lev, T_step-41] < 276.15 and pd[lev,T_step-41,k,j] > 20.0 and T_step > 40:

ys = T_step - 41

elif Ta[lev, T_step-42] < 276.15 and pd[lev,T_step-42,k,j] > 20.0 and T_step > 41:

ys = T_step - 42

elif Ta[lev, T_step-43] < 276.15 and pd[lev,T_step-43,k,j] > 20.0 and T_step > 42:

ys = T_step - 43

elif Ta[lev, T_step-44] < 276.15 and pd[lev,T_step-44,k,j] > 20.0 and T_step > 43:

ys = T_step - 44

elif Ta[lev, T_step-45] < 276.15 and pd[lev,T_step-45,k,j] > 20.0 and T_step > 44:

Page 50: Melt rate of Palcaraju glacier, Cordillera Blanca, Peru ...applicable to ‘slow-onset’ impacts, such as glacial melt, in which climate change can be shown to materially affect the

50

ys = T_step - 45

elif Ta[lev, T_step-46] < 276.15 and pd[lev,T_step-46,k,j] > 20.0 and T_step > 45:

ys = T_step - 46

elif Ta[lev, T_step-47] < 276.15 and pd[lev,T_step-47,k,j] > 20.0 and T_step > 46:

ys = T_step - 47

elif Ta[lev, T_step-48] < 276.15 and pd[lev,T_step-48,k,j] > 20.0 and T_step > 47:

ys = T_step - 48

elif Ta[lev, T_step-49] < 276.15 and pd[lev,T_step-49,k,j] > 20.0 and T_step > 48:

ys = T_step - 49

elif Ta[lev, T_step-50] < 276.15 and pd[lev,T_step-50,k,j] > 20.0 and T_step > 49:

ys = T_step - 50

elif Ta[lev, T_step-51] < 276.15 and pd[lev,T_step-51,k,j] > 20.0 and T_step > 50:

ys = T_step - 51

elif Ta[lev, T_step-52] < 276.15 and pd[lev,T_step-52,k,j] > 20.0 and T_step > 51:

ys = T_step - 52

elif Ta[lev, T_step-53] < 276.15 and pd[lev,T_step-53,k,j] > 20.0 and T_step > 52:

ys = T_step - 53

elif Ta[lev, T_step-54] < 276.15 and pd[lev,T_step-54,k,j] > 20.0 and T_step > 53:

ys = T_step - 54

elif Ta[lev, T_step-55] < 276.15 and pd[lev,T_step-55,k,j] > 20.0 and T_step > 54:

ys = T_step - 55

elif Ta[lev, T_step-56] < 276.15 and pd[lev,T_step-56,k,j] > 20.0 and T_step > 55:

ys = T_step - 56

elif Ta[lev, T_step-57] < 276.15 and pd[lev,T_step-57,k,j] > 20.0 and T_step > 56:

ys = T_step - 57

elif Ta[lev, T_step-58] < 276.15 and pd[lev,T_step-58,k,j] > 20.0 and T_step > 57:

ys = T_step - 58

elif Ta[lev, T_step-59] < 276.15 and pd[lev,T_step-59,k,j] > 20.0 and T_step > 58:

ys = T_step - 59

elif Ta[lev, T_step-60] < 276.15 and pd[lev,T_step-60,k,j] > 20.0 and T_step > 59:

ys = T_step - 60

elif Ta[lev, T_step-61] < 276.15 and pd[lev,T_step-61,k,j] > 20.0 and T_step > 60:

ys = T_step - 61

elif Ta[lev, T_step-62] < 276.15 and pd[lev,T_step-62,k,j] > 20.0 and T_step > 61:

ys = T_step - 62

elif Ta[lev, T_step-63] < 276.15 and pd[lev,T_step-63,k,j] > 20.0 and T_step > 62:

ys = T_step - 63

elif Ta[lev, T_step-64] < 276.15 and pd[lev,T_step-64,k,j] > 20.0 and T_step > 63:

ys = T_step - 64

elif Ta[lev, T_step-65] < 276.15 and pd[lev,T_step-65,k,j] > 20.0 and T_step > 64:

ys = T_step - 65

elif Ta[lev, T_step-66] < 276.15 and pd[lev,T_step-66,k,j] > 20.0 and T_step > 65:

ys = T_step - 66

elif Ta[lev, T_step-67] < 276.15 and pd[lev,T_step-67,k,j] > 20.0 and T_step > 66:

ys = T_step - 67

elif Ta[lev, T_step-68] < 276.15 and pd[lev,T_step-68,k,j] > 20.0 and T_step > 67:

ys = T_step - 68

elif Ta[lev, T_step-69] < 276.15 and pd[lev,T_step-69,k,j] > 20.0 and T_step > 68:

ys = T_step - 69

elif Ta[lev, T_step-70] < 276.15 and pd[lev,T_step-70,k,j] > 20.0 and T_step > 69:

ys = T_step - 70

elif Ta[lev, T_step-71] < 276.15 and pd[lev,T_step-71,k,j] > 20.0 and T_step > 70:

ys = T_step - 71

elif Ta[lev, T_step-72] < 276.15 and pd[lev,T_step-72,k,j] > 20.0 and T_step > 71:

ys = T_step - 72

elif Ta[lev, T_step-73] < 276.15 and pd[lev,T_step-73,k,j] > 20.0 and T_step > 72:

ys = T_step - 73

elif Ta[lev, T_step-74] < 276.15 and pd[lev,T_step-74,k,j] > 20.0 and T_step > 73:

ys = T_step - 74

elif Ta[lev, T_step-75] < 276.15 and pd[lev,T_step-75,k,j] > 20.0 and T_step > 74:

ys = T_step - 75

elif Ta[lev, T_step-76] < 276.15 and pd[lev,T_step-76,k,j] > 20.0 and T_step > 75:

Page 51: Melt rate of Palcaraju glacier, Cordillera Blanca, Peru ...applicable to ‘slow-onset’ impacts, such as glacial melt, in which climate change can be shown to materially affect the

51

ys = T_step - 76

elif Ta[lev, T_step-77] < 276.15 and pd[lev,T_step-77,k,j] > 20.0 and T_step > 76:

ys = T_step - 77

elif Ta[lev, T_step-78] < 276.15 and pd[lev,T_step-78,k,j] > 20.0 and T_step > 77:

ys = T_step - 78

elif Ta[lev, T_step-79] < 276.15 and pd[lev,T_step-79,k,j] > 20.0 and T_step > 78:

ys = T_step - 79

elif Ta[lev, T_step-80] < 276.15 and pd[lev,T_step-80,k,j] > 20.0 and T_step > 79:

ys = T_step - 80

elif Ta[lev, T_step-81] < 276.15 and pd[lev,T_step-81,k,j] > 20.0 and T_step > 80:

ys = T_step - 81

elif Ta[lev, T_step-82] < 276.15 and pd[lev,T_step-82,k,j] > 20.0 and T_step > 81:

ys = T_step - 82

elif Ta[lev, T_step-83] < 276.15 and pd[lev,T_step-83,k,j] > 20.0 and T_step > 82:

ys = T_step - 83

elif Ta[lev, T_step-84] < 276.15 and pd[lev,T_step-84,k,j] > 20.0 and T_step > 83:

ys = T_step - 84

elif Ta[lev, T_step-85] < 276.15 and pd[lev,T_step-85,k,j] > 20.0 and T_step > 84:

ys = T_step - 85

else:

ys = 0.

return ys

def xPr(Ta, pd, T_step, lev, j, k): # parameterisation from Caidong and Sorteberg 2010

if Ta[lev,T_step] >= 276.15:

yPr = pd[lev,T_step,k,j]

elif 276.15 > Ta[lev,T_step] >= 272.15:

yPr = pd[lev,T_step,k,j] * ((Ta[lev,T_step] - 272.15)/4.0)

else:

yPr = 0.0

return yPr

def xPs(Pr, pd, T_step, lev, k,j):

yPs = pd[lev,T_step,k,j] - Pr

return yPs

def xSNDEPTH(snow_depth, T_step, lev, Ps, model_elev, rhow, Lf,k,j):

if k == 0:

if T_step == 0:

ySNDEPTH = 0.

else:

ySNDEPTH = snow_depth[lev,T_step-1,k,j] + Ps - model_elev[lev,T_step-1,k,j]

else:

if T_step == 0:

ySNDEPTH = snow_depth[lev,359,0,j] + Ps - model_elev[lev,359,0,j]

else:

ySNDEPTH = ySNDEPTH = snow_depth[lev,T_step-1,k,j] + Ps - model_elev[lev,T_step-1,k,j]

return ySNDEPTH

def zSNDEPTH(aSNDEPTH):

if aSNDEPTH >= 0.:

bSNDEPTH = aSNDEPTH

else:

bSNDEPTH = 0.

return bSNDEPTH

def xSNALBEDO(T_step, s):

ySNALBEDO = 0.53 + (0.22 * math.exp((s - T_step) / 21.9))

return ySNALBEDO # Oerlemans & Knap 1998

def xALBEDO(SNALBEDO, SNDEPTH):

Page 52: Melt rate of Palcaraju glacier, Cordillera Blanca, Peru ...applicable to ‘slow-onset’ impacts, such as glacial melt, in which climate change can be shown to materially affect the

52

yALBEDO = SNALBEDO + 0.34 - (SNALBEDO * math.exp((-1. * SNDEPTH)/32.))

return yALBEDO # Oerlemans & Knap 1998

def aALBEDO(zALBEDO):

if zALBEDO > 0.75:

bALBEDO = 0.75

elif zALBEDO < 0.34:

bALBEDO = 0.34

else:

bALBEDO = zALBEDO

return bALBEDO

def xn(pd, T_step, lev, k,j):

if pd[lev,T_step,k,j] < 0.1:

yn = 0.3

else:

yn = 1.0

return yn

# 2. Incoming shortwave radiation calculation

def xKt(alpham, nat_glob_rad, T_step): # Shortwave reflected from surrounding terrain

yKt = alpham * nat_glob_rad[T_step] * 0.07783604

return yKt

def xSHORTIN(nat_glob_rad, T_step, Kt):

ySHORTIN = nat_glob_rad[T_step] + Kt

return ySHORTIN

def xSHORTNET(SHORTIN, ALBEDO):

ySHORTNET = SHORTIN * (1. - ALBEDO)

return ySHORTNET

#--- 3. Longwave radiation ---#

# General equation for incoming L: Hock, 2005; using Arnold et al (1996) for ec*F(n)

def xLONGIN(sigma, Ta, T_step, n, lev, k,j):

yLONGIN = sigma * ((Ta[lev,T_step])**4.) * 0.008733 * ((Ta[lev,T_step])**0.788) * (1. + (0.26*n)) # Incoming solar radiation (l/w)

return yLONGIN

#--- Glacier surface temperature (Oerlemans, 1991) ---#

def xmean_temp(Ta):

ymean_temp = np.array(np.mean(Ta, axis=1)) # deg C

return ymean_temp

#def xlatent_minus(melt_en, glac_temp, lev, T_step, k, j):

# if T_step == 0:

# zlatent_release = 0.

# else:

# zlatent_release = melt_en[lev,T_step-1,k,j] - (melt_en[lev,T_step-1,k,j]**glac_temp[lev,T_step-1,k,j])

# return zlatent_release

#def xglacier_temp(mean_temp, lev, T_step, glac_temp, latent_min, k,j):

def xglacier_temp(mean_temp, lev):

yglacier_temp = mean_temp[lev] - 273.15

return yglacier_temp

def zglacier_temp(glacier_tem):

if glacier_tem > 0.:

aglacier_temp = 0.

else:

Page 53: Melt rate of Palcaraju glacier, Cordillera Blanca, Peru ...applicable to ‘slow-onset’ impacts, such as glacial melt, in which climate change can be shown to materially affect the

53

aglacier_temp = glacier_tem

return aglacier_temp

# Longwave radiation from surrounding terrain

#def xLONGTERR(glacier_temp, Lb, Lta, Ltb, T_step, Ta, lev, k,j): # Pluss and Ohmura, 1997. Does not include obstruction by

surrounding terrain.

# yLONGTERR = Lb + (Lta * (Ta[lev,T_step,k,j] - 273.15)) + (Ltb * glacier_temp)

# return yLONGTERR

# Longwave outgoing Radiation

def xLONGOUT(sigma, glacier_temp):

yLONGOUT = sigma * ((glacier_temp+273.15)**4.)

return yLONGOUT

def xLONGNET(LONGIN, LONGOUT):

yLONGNET = LONGIN - LONGOUT

return yLONGNET

#--- 4. Turbulent heat fluxes ---#

#--- a) Sensible heat ---#

def xAIR_PRESS(atm0, Lap, g, M, gc, b):

#yAIR_PRESS = atm0 * ((1.0 - (Lap * b)/288.15)**((g*M)/(gc*Lap))) # Atmospheric pressure at grid-box level

yAIR_PRESS = atm0 *((288.15/(288.15+(Lap * b)))**((g*M)/(gc*Lap)))

return yAIR_PRESS

# Based on Arnold et al (1996)

def xSENSIBLE(Ks, AIR_PRESS, Ta, T_step, nat_vw, lev, k,j):

ySENSIBLE = Ks * AIR_PRESS * (Ta[lev,T_step] - 273.15) * nat_vw[T_step]

return ySENSIBLE

#--- b) Latent heat ---#

# Based on Arnold et al (1996)

#--- Arden Buck equation for saturation vapour pressure over ice ---#

def xsatvpi(Ta, lev, T_step):

if Ta[lev,T_step] >= 273.15:

esice = 100.0 * 0.61121 * math.exp((18.678 - ((Ta[lev,T_step]-273.15)/234.5)) * ((Ta[lev,T_step]-273.15)/(257.14 + Ta[lev,T_step]

- 273.15)))

else:

esice = 100.0 * 0.61115 * math.exp((23.036-((Ta[lev,T_step]-273.15)/333.7))*((Ta[lev,T_step]-273.15)/(279.82+Ta[lev,T_step]-

273.15)))

return esice

#def vapour_press():

# yvappress = ?????

# return yvappress

def xdeltae(satvpi): # Difference between vapour pressure of air and saturation vapour pressure at the glacier surface

ydeltae = satvpi - 611.

return ydeltae

def xLATENT(K1, deltae, nat_vw, T_step):

yLATENT = K1 * deltae * nat_vw[T_step]

return yLATENT

#--- Calculate energy available for melt ---#

def xMELT_ENERGY(SHORTIN, LONGNET, ALBEDO, SENSIBLE, LATENT):

yMELT_ENERGY = (SHORTIN * (1.0 - ALBEDO) + LONGNET + SENSIBLE + LATENT)

if yMELT_ENERGY > 0.:

zMELT_ENERGY = yMELT_ENERGY

Page 54: Melt rate of Palcaraju glacier, Cordillera Blanca, Peru ...applicable to ‘slow-onset’ impacts, such as glacial melt, in which climate change can be shown to materially affect the

54

else:

zMELT_ENERGY = 0.

return zMELT_ENERGY

#--- Calculate refreezing (Oerlemans, 1991) ---#

def xmeng(MELT_ENERGY, glacier_temp):

ymeng = MELT_ENERGY**glacier_temp

return ymeng

def xlatent_release(MELT_ENERGY, meng):

ylatent_release = MELT_ENERGY - meng

return ylatent_release

def xrefrozen(latent_released, Lf): # Oerlemans, 1991

yrefrozen = latent_released / Lf

return yrefrozen

def zrefrozen(arefrozen, rhow, Lf, MELT_ENERGY, secs_per_day):

if arefrozen < (MELT_ENERGY * secs_per_day * 1000. / (rhow * Lf)):

brefrozen = arefrozen

else:

brefrozen = MELT_ENERGY * secs_per_day * 1000. / (rhow * Lf)

return brefrozen

#--- Calculate net melt ---#

def xTOTAL_MELT_ELEV(rhow, Lf, MELT_ENERGY, secs_per_day, refrozen):

yTOTAL_MELT_ELEV = MELT_ENERGY * secs_per_day * 1000. / (rhow * Lf) - refrozen # Melt

if yTOTAL_MELT_ELEV < 100.:

zTOTAL_MELT_ELEV = yTOTAL_MELT_ELEV

else:

zTOTAL_MELT_ELEV = 1.

return zTOTAL_MELT_ELEV

def xDAY_TOTAL_MELT(TOTAL_MELT_ELEV):

yDAY_TOTAL_MELT = np.array(np.mean(TOTAL_MELT_ELEV, axis=0)) # Is this correct?

return yDAY_TOTAL_MELT

def xensemble_mean(model_elev):

yensemble_mean = np.array(np.mean(model_elev, axis=3))

return yensemble_mean

def xensemble_mean_sw(short_net):

yensemble_mean_sw = np.array(np.mean(short_net, axis=3))

return yensemble_mean_sw

def xensemble_mean_lw(longnet_arr):

yensemble_mean_lw = np.array(np.mean(longnet_arr, axis=3))

return yensemble_mean_lw

def xensemble_mean_sens(sensflux):

yensemble_mean_sens = np.array(np.mean(sensflux, axis=3))

return yensemble_mean_sens

def xensemble_mean_latent(latentflux):

yensemble_mean_latent = np.array(np.mean(latentflux, axis=3))

return yensemble_mean_latent

def xlevmean_sw(sw_mean):

ylevmean_sw = np.array(np.mean(sw_mean,axis=0))

return ylevmean_sw

def xlevmean_lw(lw_mean):

Page 55: Melt rate of Palcaraju glacier, Cordillera Blanca, Peru ...applicable to ‘slow-onset’ impacts, such as glacial melt, in which climate change can be shown to materially affect the

55

ylevmean_lw = np.array(np.mean(lw_mean,axis=0))

return ylevmean_lw

def xlevmean_sens(sens_mean):

ylevmean_sens = np.array(np.mean(sens_mean,axis=0))

return ylevmean_sens

def xlevmean_latent(latent_mean):

ylevmean_latent = np.array(np.mean(latent_mean,axis=0))

return ylevmean_latent

def main():

list_of_ensmem = glob.glob('/soge-home/users/shil4723/dissertation/rcm/precip/factual/raw/model/2007/x_sr_precip_*.nc')

n_ens = len(list_of_ensmem)

nat_list_of_ensmem = glob.glob('/soge-

home/users/shil4723/dissertation/rcm/precip/counterfactual/raw/model/2007/x_sr_precip_*.nc')

nat_ens = len(nat_list_of_ensmem)

melt_en = np.zeros((8,360,2,22))

snow_depth = np.zeros((8,360,2,22))

glac_temp = np.zeros((8,360,2,22))

latent_release = np.zeros((8,360,2,22))

pd = np.zeros((8,360,2,22))

mean_temp = np.zeros((8))

ensemble_mean = np.zeros((8,360,2))

albedo_array = np.zeros((8,360,2,22))

model_elev = np.zeros((8,360,2,22))

Ta=np.empty((8,360))

Ta[:] = np.nan

short = np.zeros((8,360,2,22))

short_terr = np.zeros((8,360,2,22))

short_net = np.zeros((8,360,2,22))

long = np.zeros((8,360,2,22))

longterrain = np.zeros((8,360,2,22))

longoutgoing = np.zeros((8,360,2,22))

longnet_arr = np.zeros((8,360,2,22))

sensflux = np.zeros((8,360,2,22))

latentflux = np.zeros((8,360,2,22))

refreeze = np.zeros((8,360,2,22))

snow = np.zeros((8,360,2,22))

sw_in = np.zeros((8,360,2,22))

sw_mean = np.zeros((8,360,2))

lw_mean = np.zeros((8,360,2))

sens_mean = np.zeros((8,360,2))

latent_mean = np.zeros((8,360,2))

sw = np.zeros((360,2))

lw = np.zeros((360,2))

sens = np.zeros((360,2))

latent = np.zeros((360,2))

#--- Set up mountain topography ---#

# topo = 'Palcaraju_topo.xlsx'

# x, y, w = THREE_VALUE_GRAB.GetData(topo)

y = [4625., 4675., 4725., 4775., 4825., 4875., 4925., 4975.]

az = 215. # slope azimuth

grad = 32.4 # slope gradient

# Temperature

file_temp = '/soge-home/users/shil4723/dissertation/rcm/temp/counterfactual/biascorr/model/glacier_2007.nc'

temp = Dataset(file_temp, mode='r')

Tgl = temp.variables['temp'][:] # Daily temp at bottom of glacier

Tgl_units = temp.variables['temp'].units

Page 56: Melt rate of Palcaraju glacier, Cordillera Blanca, Peru ...applicable to ‘slow-onset’ impacts, such as glacial melt, in which climate change can be shown to materially affect the

56

temp.close()

nat_file_temp = '/soge-home/users/shil4723/dissertation/rcm/temp/factual/biascorr/model/glacier_2007.nc'

nat_temp = Dataset(nat_file_temp, mode='r')

nat_Tgl = nat_temp.variables['temp'][:]

nat_Tgl_units = nat_temp.variables['temp'].units

nat_temp.close()

for T_step in np.arange(360):

for lev in np.arange(8):

b = y[lev]

tmp = xTa(Tgl, LR, b, T_step)

print tmp

Ta[lev,T_step] = tmp

lev += 1

T_step += 1

for j in np.arange(6):

precipitation_file = '/soge-home/users/shil4723/dissertation/rcm/precip/factual/raw/model/21/x_sr_precip_' + str(j+1) + '.nc'

precip = Dataset(precipitation_file, mode='r')

time = precip.variables['time0'][:]

time_units = precip.variables['time0'].units

t = num2date(time[:],time_units)

ppt = precip.variables['precip'][:] # Daily precipitation

ppt_units = precip.variables['precip'].units

glob_rad = precip.variables['sr_flux'][:]

glob_rad_units = precip.variables['sr_flux'].units

precip.close()

nat_precipitation_file = '/soge-home/users/shil4723/dissertation/rcm/precip/counterfactual/raw/model/21/x_sr_precip_' +

str(j+1) + '.nc'

nat_precip = Dataset(nat_precipitation_file, mode='r')

nat_ppt = nat_precip.variables['precip'][:]

nat_glob_rad = nat_precip.variables['sr_flux'][:]

nat_precip.close()

# Temperature

# file_temp = '/soge-home/users/shil4723/dissertation/rcm/temp/factual/biascorr/model/glacier_' + str(q+1) + '.nc'

file_temp = '/soge-home/users/shil4723/dissertation/rcm/temp/factual/biascorr/model/glacier_21.nc'

temp = Dataset(file_temp, mode='r')

Tgl = temp.variables['temp'][:] # Daily temp at bottom of glacier

Tgl_units = temp.variables['temp'].units

temp.close()

nat_file_temp = '/soge-home/users/shil4723/dissertation/rcm/temp/counterfactual/biascorr/model/glacier_21.nc'

nat_temp = Dataset(nat_file_temp, mode='r')

nat_Tgl = nat_temp.variables['temp'][:]

nat_Tgl_units = nat_temp.variables['temp'].units

nat_temp.close()

# Wind Speed

wind = '/soge-home/users/shil4723/dissertation/rcm/wind/factual/wind_21.nc'

vel = Dataset(wind, mode='r')

vw = vel.variables['v_wind'][:]

vel.close()

nat_wind = '/soge-home/users/shil4723/dissertation/rcm/wind/counterfactual/wind_21.nc'

nat_vel = Dataset(wind, mode='r')

nat_vw = nat_vel.variables['v_wind'][:]

nat_vel.close()

n_days = len(t)

Page 57: Melt rate of Palcaraju glacier, Cordillera Blanca, Peru ...applicable to ‘slow-onset’ impacts, such as glacial melt, in which climate change can be shown to materially affect the

57

for k in np.arange(2):

# Calculate the glacier dynamics and surface energy balance for the number of steps per year

for T_step in np.arange(360):

#--- Defining grid-cell elevations ---#

for lev in np.arange(8):

b = y[lev]

ypd = xpd(nat_ppt, T_step)

pd[lev,T_step,k,j] = ypd

ELA = xELA(nat_Tgl, LR, T_step)

s = xs(Ta, pd, T_step, lev, j, k)

Pr = xPr(Ta, pd, T_step, lev, j, k)

Ps = xPs(Pr, pd, T_step, lev, k,j)

snow[lev,T_step,k,j] = Ps

aSNDEPTH = xSNDEPTH(snow_depth, T_step, lev, Ps, model_elev, rhow, Lf,k,j)

SNDEPTH = zSNDEPTH(aSNDEPTH)

snow_depth[lev,T_step,k,j] = SNDEPTH

SNALBEDO = xSNALBEDO(T_step, s)

zALBEDO = xALBEDO(SNALBEDO, SNDEPTH)

ALBEDO = aALBEDO(zALBEDO)

albedo_array[lev,T_step,k,j] = ALBEDO

n = xn(pd, T_step, lev,k,j)

Kt = xKt(alpham, nat_glob_rad, T_step)

short_terr[lev,T_step,k,j] = Kt

SHORTIN = xSHORTIN(nat_glob_rad, T_step, Kt)

sw_in[lev,T_step,k,j] = SHORTIN

SHORTNET = xSHORTNET(SHORTIN, ALBEDO)

short_net[lev,T_step,k,j] = SHORTNET

LONGIN = xLONGIN(sigma, Ta, T_step, n, lev, k,j)

long[lev,T_step,k,j] = LONGIN

mean_temp[:] = np.squeeze(xmean_temp(Ta))[:]

glacier_tem = xglacier_temp(mean_temp, lev)

glacier_temp = zglacier_temp(glacier_tem)

glac_temp[lev,T_step,k,j] = glacier_temp

LONGOUT = xLONGOUT(sigma, glacier_temp)

longoutgoing[lev,T_step,k,j] = LONGOUT

LONGNET = xLONGNET(LONGIN, LONGOUT)

longnet_arr[lev,T_step,k,j] = LONGNET

AIR_PRESS = xAIR_PRESS(atm0, LR, g, M, gc, b)

SENSIBLE = xSENSIBLE(Ks, AIR_PRESS, Ta, T_step, nat_vw, lev,k,j)

sensflux[lev,T_step,k,j] = SENSIBLE

satvpi = xsatvpi(Ta, lev, T_step)

deltae = xdeltae(satvpi)

LATENT = xLATENT(K1, deltae, nat_vw, T_step)

latentflux[lev,T_step,k,j] = LATENT

MELT_ENERGY = xMELT_ENERGY(SHORTIN, LONGNET, ALBEDO, SENSIBLE, LATENT)

melt_en[lev,T_step,k,j] = MELT_ENERGY

meng = xmeng(MELT_ENERGY, glacier_temp)

latent_released = xlatent_release(MELT_ENERGY, meng)

latent_release[lev,T_step,k,j] = latent_released

arefrozen = xrefrozen(latent_released, Lf)

refrozen = zrefrozen(arefrozen, rhow, Lf, MELT_ENERGY, secs_per_day)

refreeze[lev,T_step,k,j] = refrozen

TOTAL_MELT_ELEV = xTOTAL_MELT_ELEV(rhow, Lf, MELT_ENERGY, secs_per_day, refrozen)

model_elev[lev,T_step,k,j] = TOTAL_MELT_ELEV

lev += 1

T_step += 1

k += 1

j += 1

ensemble_mean[:] = np.squeeze(xensemble_mean(model_elev))[:]

#--- Plotting script inserted here ---#

main()