into the smoke: a research on household air ...oa.upm.es/49583/1/candela_de_la_sota_sandez.pdfcada...

125
UNIVERSIDAD POLITÉCNICA DE MADRID ESCUELA TÉCNICA SUPERIOR DE INGENIEROS INDUSTRIALES INTO THE SMOKE: A RESEARCH ON HOUSEHOLD AIR POLLUTION AND CLIMATE IMPACTS OF BIOMASS COOKSTOVES IN SENEGAL Tesis doctoral Candela de la Sota Sández Ingeniera química 2017

Upload: others

Post on 19-Aug-2021

3 views

Category:

Documents


0 download

TRANSCRIPT

Page 1: INTO THE SMOKE: A RESEARCH ON HOUSEHOLD AIR ...oa.upm.es/49583/1/CANDELA_DE_LA_SOTA_SANDEZ.pdfCada año, la contaminación del aire interior provoca la muerte prematura de 4,3 millones

UNIVERSIDAD POLITÉCNICA DE MADRID

ESCUELA TÉCNICA SUPERIOR DE INGENIEROS INDUSTRIALES

INTO THE SMOKE:

A RESEARCH ON HOUSEHOLD AIR POLLUTION AND CLIMATE

IMPACTS OF BIOMASS COOKSTOVES IN SENEGAL

Tesis doctoral

Candela de la Sota Sández

Ingeniera química

2017

Page 2: INTO THE SMOKE: A RESEARCH ON HOUSEHOLD AIR ...oa.upm.es/49583/1/CANDELA_DE_LA_SOTA_SANDEZ.pdfCada año, la contaminación del aire interior provoca la muerte prematura de 4,3 millones

DEPARTAMENTO DE INGENIERÍA QUÍMICA INDUSTRIAL Y MEDIO

AMBIENTE

ESCUELA TÉCNICA SUPERIOR DE INGENIEROS INDUSTRIALES

INTO THE SMOKE:

A RESEARCH ON HOUSEHOLD AIR POLLUTION AND CLIMATE

IMPACTS OF BIOMASS COOKSTOVES IN SENEGAL

Autora:

Candela de la Sota Sández

Ingeniera Química

Director:

Julio Lumbreras Martín

Doctor Ingeniero Industrial

2017

Page 3: INTO THE SMOKE: A RESEARCH ON HOUSEHOLD AIR ...oa.upm.es/49583/1/CANDELA_DE_LA_SOTA_SANDEZ.pdfCada año, la contaminación del aire interior provoca la muerte prematura de 4,3 millones

Agradecimientos

Gracias a todas aquellas personas que me han acompañado durante estos años en la

realización de este proyecto de investigación, que para mí ha sido tan importante.

Gracias a las familias de las comunidades de Senegal y Nicaragua por abrirme las

puertas de sus casas, especialmente a las mujeres, que han compartido sus

experiencias y su tiempo conmigo. Sin ellas, esta tesis no hubiera sido posible.

Gracias a la UPM, ONGAWA, IDAEA-CSIC, CERER, GACC, UNAM y a la Fundación

Iberdrola España, por su fundamental apoyo económico y técnico.

Por último, quiero mostrar mi más profundo rechazo al gobierno de este país, que, con

sus recortes y políticas, se empeña en destruir la investigación científica, entre muchas

otras cosas. Espero que no lo consigáis.

Page 4: INTO THE SMOKE: A RESEARCH ON HOUSEHOLD AIR ...oa.upm.es/49583/1/CANDELA_DE_LA_SOTA_SANDEZ.pdfCada año, la contaminación del aire interior provoca la muerte prematura de 4,3 millones

Table of contents

Resumen ......................................................................................................................................... i

Abstract .........................................................................................................................................iv

List of figures ................................................................................................................................ vii

List of tables .................................................................................................................................. ix

Abbreviations ................................................................................................................................ x

1. Introduction .......................................................................................................................... 1

1.1. Household energy for cooking. Global situation and impacts ...................................... 2

1.2. Overview of clean and improved cooking solutions and related challenges ................ 4

1.3. Research opportunities ................................................................................................. 8

1.4. Research aim and overview .......................................................................................... 9

1.5. Personal motivation .................................................................................................... 10

2. Research context ................................................................................................................. 11

2.1. Cookstoves and household air pollution ..................................................................... 12

2.2. Carbonaceous aerosol emissions from cookstoves .................................................... 14

2.3. Household energy for cooking in sub-Saharan and Western Africa ........................... 17

3. Inter-comparison of methods to measure black carbon from cookstoves ......................... 21

3.1. Background.................................................................................................................. 22

3.2. Methods ...................................................................................................................... 23

3.2.1. Sample collection .................................................................................................... 23

3.2.2. Determination of BC and EC on filters .................................................................... 25

3.2.3. Method limitations .................................................................................................. 27

3.2.4. Technical requirements and features of the methods ........................................... 28

3.2.5. Data analysis ............................................................................................................ 29

3.3. Results and discussion ................................................................................................. 29

3.3.1. BC analysis by the cell-phone system compared with EC by thermo-optical analysis

29

3.3.2. Reflectance analysis by smoke stain reflectometer compared with EC by thermo-

optical analysis ........................................................................................................................ 32

3.3.3. Analysis of filter substrate on BC and EC quantification ....................................... 33

4. Quantification of carbonaceous aerosol emissions from cookstoves in Senegal ............... 36

4.1. Background.................................................................................................................. 37

4.2. Methods ...................................................................................................................... 37

Page 5: INTO THE SMOKE: A RESEARCH ON HOUSEHOLD AIR ...oa.upm.es/49583/1/CANDELA_DE_LA_SOTA_SANDEZ.pdfCada año, la contaminación del aire interior provoca la muerte prematura de 4,3 millones

4.2.1. Laboratory sampling ................................................................................................ 37

4.2.2. Field sampling .......................................................................................................... 38

4.2.3. Stoves and fuels....................................................................................................... 39

4.2.4. EC and OC analysis .................................................................................................. 41

4.2.5. Data Analysis ........................................................................................................... 41

4.3. Results and discussion ................................................................................................. 41

4.3.1. Laboratory estimation of EC and OC emission using the WBT ............................... 41

4.3.2. Field estimation of EC and OC emissions ................................................................ 48

4.3.3. Laboratory vs. field results ...................................................................................... 54

4.3.4. Estimation of total emissions and net climate effect of cooking-related

carbonaceous aerosol in Senegal ............................................................................................ 55

5. Indoor air pollution from biomass cookstoves in Senegal .................................................. 58

5.1. Background....................................................................................................................... 59

5.2. Methods ........................................................................................................................... 60

5.2.1. Household selection and study design .................................................................... 60

5.2.2. Indoor Air Pollution monitoring .............................................................................. 64

5.2.3. Instrument quality control ...................................................................................... 67

5.2.4. Data Analysis ........................................................................................................... 70

5.3. Results and discussion ................................................................................................. 71

5.3.1. Indoor air pollutant concentrations during cooking activities ................................ 71

5.3.2. Indoor air pollutant concentrations during cooking periods as a function of the

type of stove ............................................................................................................................ 74

5.3.3. Effect of household-level determinants on indoor air pollution ............................ 80

5.3.4. Comparison between IAP meter and DustTrak units .............................................. 82

6. Conclusions ......................................................................................................................... 84

7. Limitations and further research......................................................................................... 90

8. References ........................................................................................................................... 93

Page 6: INTO THE SMOKE: A RESEARCH ON HOUSEHOLD AIR ...oa.upm.es/49583/1/CANDELA_DE_LA_SOTA_SANDEZ.pdfCada año, la contaminación del aire interior provoca la muerte prematura de 4,3 millones

i

Resumen

Más de la mitad de la población mundial depende de los combustibles sólidos (leña,

carbón vegetal, residuos agrícolas y animales) para cocinar. Estos combustibles son

quemados en el interior de los hogares en cocinas poco eficientes, convirtiendo los

hogares en entornos muy peligrosos para la salud.

Cada año, la contaminación del aire interior provoca la muerte prematura de 4,3

millones de personas, un número mayor que la suma de las muertes producidas por el

paludismo, el sida y la tuberculosis. Las mujeres, y los niños que a menudo las

acompañan, inhalan estas sustancias durante horas, ya que ellas son las responsables

de preparar la comida para sus familias. Además, ellas son las principales responsables

de recolectar el combustible para cocinar, poniendo en peligro su seguridad y

reduciendo de manera importante su tiempo disponible para la educación, el descanso

y las actividades generadoras de ingresos.

Pero el problema no se queda en casa. La quema de combustibles sólidos a nivel

doméstico supone el 12% de las emisiones mundiales de partículas finas (PM2.5) y el

25% de las emisiones mundiales de carbono negro (BC, por sus siglas en inglés) o

carbono elemental (EC) , que se considera el segundo mayor contribuyente al

calentamiento global, después del CO2.

Afortunadamente, la comunidad mundial está intensificando sus esfuerzos para

garantizar el acceso a una energía limpia y segura para cocinar en los hogares, y se

considera ya un aspecto esencial para poder avanzar en los objetivos planteados en la

Agenda 2030 para el Desarrollo Sostenible.

Existe una amplia gama de soluciones de cocinado limpias y eficientes, cada una de

ellas con una determinada eficiencia, coste, ventajas y desafíos en cuanto a la

satisfacción de las necesidades de las personas usuarias. Aunque estas soluciones de

cocinado puedan parecer simples, la naturaleza estocástica de la combustión de la

biomasa, combinada con las variaciones naturales durante su uso diario, la amplia

variedad de tecnologías y combustibles disponibles, dificultan la evaluación de las

mismas, especialmente de algunos parámetros o contaminantes, como el BC.

Hasta la fecha, se han realizado pocos estudios para cuantificar las emisiones de BC de

la quema de biomasa a nivel residencial, en parte debido a que los métodos analíticos

disponibles son relativamente costosos y complejos. Además, el efecto de

calentamiento del BC es bastante incierto, puesto que se emite siempre junto con el

carbono orgánico (OC), tradicionalmente asociado a efectos de enfriamiento.

Por otra parte, la necesidad de estudios para cuantificar los impactos relacionados con

la quema de combustibles para cocinar no es homogénea en todo el mundo.

Page 7: INTO THE SMOKE: A RESEARCH ON HOUSEHOLD AIR ...oa.upm.es/49583/1/CANDELA_DE_LA_SOTA_SANDEZ.pdfCada año, la contaminación del aire interior provoca la muerte prematura de 4,3 millones

ii

Tradicionalmente, la mayoría de los estudios se han centrado en Asia, y muy pocos

estudios se han llevado a cabo en otras regiones del mundo, como en África

subsahariana.

En este contexto, esta tesis pretende contribuir al conocimiento existente sobre los

impactos en el clima y la calidad del aire interior producidos por el cocinado con

biomasa a nivel residencial, con especial énfasis en las emisiones de BC y OC, y

centrándose en la región de África Occidental, donde reside un tercio de la población

total de África subsahariana y donde el 75% de la gente todavía utiliza madera y

carbón para cocinar.

El Capítulo 3 presenta una intercomparación de tres métodos analíticos para estimar

las emisiones de BC procedentes de los sistemas de cocinado. Dos métodos

relativamente baratos y fáciles de usar, un reflectómetro (Difussion Systems Ltd.) y un

sistema de análisis a través de teléfono móvil (Nexleaf Analytics), fueron validados con

un dispositivo más preciso, el Analizador OC-EC (Sunset Laboratory).

Los resultados muestran una buena concordancia entre las dos técnicas de bajo coste,

lo que supone una importante ayuda para superar las actuales barreras metodológicas

para la determinación de emisiones de BC procedentes de cocinas en lugares con

recursos limitados. Asimismo, la facilidad de uso del sistema basado en el teléfono

móvil puede facilitar la participación de las usuarias de las cocinas en los procesos de

recopilación de datos, aumentando así la cantidad y la frecuencia de la recolección de

información.

En el capítulo 4 se presenta una cuantificación de emisiones de aerosoles carbonosos

(EC y OC) de cuatro tipos de cocina ampliamente utilizados en Senegal y países

vecinos: i) Tres piedras; ii) cocina tipo rocket; iii) cocina mejorada básica de cerámica);

iv) gasificador. Para todas las cocinas analizadas, los Factores de Emisión (FE) (g/MJ) de

EC y OC se vieron significativamente afectados por la especie forestal, desatanco la

necesidad de tener en cuenta el tipo de combustible al presentar los FE de los sistemas

de cocinado. En cuanto al efecto del tipo de cocina, la cocina rocket dio lugar a los FE

de EC más altos, mientras que las otras tres cocinas mostraron valores muy similares.

En lo que se refiere a emisiones totales por test, el gasificador mostró los valores más

bajos, y la rocket, los más elevados.

Asimismo, la cocina tradicional y la rocket fueron analizadas en una aldea rural de

Senegal, con el fin de obtener datos de emisiones de EC y OC, y entender cuán

representativos son los estudios estandarizados de laboratorio de las prácticas de

cocinado reales de África Occidental. Al igual que en los resultados de laboratorio, la

estufa tipo rocket mostró un FE y emisiones totales de EC más alto que la cocina

tradicional. Además, el FE y emisiones totales de OC fueron más bajos con la cocina

rocket, por lo que las emisiones de aerosoles carbonosos de este tipo de cocinas

Page 8: INTO THE SMOKE: A RESEARCH ON HOUSEHOLD AIR ...oa.upm.es/49583/1/CANDELA_DE_LA_SOTA_SANDEZ.pdfCada año, la contaminación del aire interior provoca la muerte prematura de 4,3 millones

iii

producirían un efecto neto de calentamiento positivo, en comparación con la cocina

tradicional.

El capítulo 5 presenta un análisis de concentración de PM2.5, partículas ultrafinas, BC y

monóxido de carbono (CO) en el interior de las viviendas en la misma aladea rural de

Senegal. Los resultados confirman que la contaminación del aire en los hogares en esta

zona se deben principalmente a las actividades de cocinado. Además, la instalación de

la cocina tipo rocket contribuyó a una reducción significativa PM2.5, partículas

ultrafinas y CO, pero incrementó la concentración de BC, lo cual es coherente con los

resultados de FE presentados en el capítulo 4.

Los resultados también evidencian que, además de un cambio en la fuente de emisión

(es decir, la cocina y/o el combustible), las estrategias enfocadas en la mejora de la

calidad del aire doméstico deben enfocarse en las prácticas de ventilación y el uso de

los materiales de construcción más adecuados.

Según los resultados presentados en los capítulos 4 y 5, la implementación de estufas

tipo rocket tendría un efecto positivo con respecto a la reducción de la contaminación

interior, pero efectos más inciertos sobre el clima. Esto demuestra que los efectos en

el clima y en salud de las soluciones de cocinado no siempre están alineados y destaca

que ambas dimensiones deben ser consideradas en las decisiones tecnológicas. No

obstante, en el estudio no se incluyen otros contaminantes emitidos y factores

importantes, como la renovabilidad de la biomasa empleada para cocinar. Por lo tanto,

los resultados presentados no significan que el uso de cocinas tipo rocket no tenga

beneficios medio ambientales, ya que el panorama completo es mucho más

complicado e incierto.

La información presentada en esta tesis ayuda a tener una imagen más precisa de los

impactos producidos por la quema de biomasa para cocinar a nivel residencial en la

región de África Occidental. Además, puede ser útil para los implementadores y

tomadores de decisiones de la región a la hora de seleccionar las soluciones

tecnológicas más apropiadas en función de los impactos esperados, y para que los

diseñadores optimicen los co-beneficios para el medio ambiente y la salud de las

soluciones de cocinado.

Se necesitan más estudios de laboratorio y de terreno que analicen diferentes sistemas

de cocinado, cubriendo otras áreas de África Occidental y que también estudien el

efecto de la estacionalidad sobre las emisiones, con el fin de obtener datos más

representativos para su uso en los inventarios de emisiones y modelos climáticos.

Page 9: INTO THE SMOKE: A RESEARCH ON HOUSEHOLD AIR ...oa.upm.es/49583/1/CANDELA_DE_LA_SOTA_SANDEZ.pdfCada año, la contaminación del aire interior provoca la muerte prematura de 4,3 millones

iv

Abstract

Over half of the world’s population still rely on solid fuels, like fuelwood, charcoal,

coal, agricultural residues and animal wastes for cooking. These fuels are burnt inside

homes in inefficient stoves, turning the home into one of the most health damaging

environments.

Each year, 4.3 million people prematurely die because of the air pollution inside their

homes, a number above the deaths produced by malaria, aids and tuberculosis

together. Women, and children who are often with them, inhale the smoke for hours,

as they are responsible for preparing meals for their families. Moreover, they are the

primary responsible for gathering fuel for cooking, facing safety risks and suffering

significant constraints on their available time for education, rest and activities for

income generation.

And the problem does not stay at home. Residential solid fuel burning accounts for up

to 12% of global ambient fine particulate matter (PM2.5) emissions and 25% of global

emissions of black carbon (BC) or elemental carbon (EC), which is considered the

second biggest climate forcer after CO2.

Fortunately, the global community is intensifying its efforts to expand and accelerate

access to clean household energy for cooking, since addressing this issue is essential to

make progress toward the Sustainable Development Agenda.

A broad range of stove and fuel solutions have been developed, each one with

differentiated efficiencies, costs, distribution models and challenges in term of

meeting user needs. While household’ stoves often look simple in appearance, the

stochastic nature of biomass combustion, combined with natural variations in daily

use, the wide variety of fuel-stoves combinations available and their scattered

location, make the stove evaluation challenging, especially for some parameters or

pollutants, as the BC.

To date, quantitative data of BC emissions estimates from residential biomass burning

for cooking is scarce, in part due to the relatively costly and complex analytical

methods available. Moreover, the warming effect of BC is highly uncertain, as it is

always co-emitted with organic carbon (OC), which has been traditionally linked to

cooling effects in the atmosphere.

On the other hand, the need of studies to quantify coking-related impacts is not

homogeneous across the globe. Traditionally, most studies have been focused in Asia

with far few studies in other regions of the world.

Page 10: INTO THE SMOKE: A RESEARCH ON HOUSEHOLD AIR ...oa.upm.es/49583/1/CANDELA_DE_LA_SOTA_SANDEZ.pdfCada año, la contaminación del aire interior provoca la muerte prematura de 4,3 millones

v

In this context, this thesis aims to contribute to the existing knowledge about climate

and indoor air quality impacts of biomass cookstoves, with a particular emphasis on BC

and OC emissions, and focusing in the West African region.

Chapter 3 presents an intercomparison of analytical methods to estimate BC emissions

from cookstoves. The performance of two relatively low-cost and easy-to-use

methods, a cell-phone based system and smoke stain reflectometer, was validated

against a more accurate device, the Sunset OCEC Analyzer (TOT).

A good agreement was observed between the two low cost techniques and the

reference system for the aerosol types and concentrations assessed. This validation

helps to overcome current methodological barriers to determine BC emissions from

cookstoves in resource-constrained locations. Moreover, the easy use of the cell-

phone based system may allow engaging cookstove users in the data collection

process, increasing the amount and frequency of data collection and helping to raise

public awareness about environmental and health issues related to cookstoves.

Chapter 4 presents a carbonaceous aerosols (EC and OC) emissions characterization of

a set of stove-fuel combinations widely used in Senegal and neighbouring countries: i)

Three-stone; ii) rocket stove; iii) Basic improved ceramic stove); and iv) natural draft

gasifier.

For all the stoves tested, EC and OC Emission Factors (EFs) (g/MJ) where significantly

affected by the wood specie, highlighting the need of taking into account the fuel type

when reporting cookstove carbonaceous aerosol EFs. The highest average EC EF per

WBT was found with the rocket stove, while EC EF’s were fairly uniform across the

other two improved cookstoves and the three stones. Considering EC and OC total

emissions per test (Water Boiling Test), the gasifier induced to the smallest total

emissions and the rocket stoves to the highest.

To understand how real cooking practices in this region influenced EC and OC

emissions and how representative are the standardized studies of actual cooking

practices in the region, the three stones and the rocket stove were tested in a rural

village of Senegal. Similarly to laboratory results, rocket stove showed the highest EC

EF and EC total emissions per cooking task. Moreover, OC emissions were reduced

with the rocket stove, so carbonaceous emissions of this stove type would produce a

net positive warming effect when compared the traditional stove.

Chapter 5 presents a real-world assessment of household concentrations of fine

particulate matter, ultrafine particles, BC and carbon monoxide conducted in the same

rural village. Results confirmed that HAP in this area is mainly due to cooking activities

and showed that the installation of the rocket stove contributed to a significant

reduction of fine and ultrafine particulate matter and CO concentrations, but increased

indoor BC concentrations. Findings also evidence that, in addition to a switch in the

Page 11: INTO THE SMOKE: A RESEARCH ON HOUSEHOLD AIR ...oa.upm.es/49583/1/CANDELA_DE_LA_SOTA_SANDEZ.pdfCada año, la contaminación del aire interior provoca la muerte prematura de 4,3 millones

vi

emission source (i.e. cookstove and/or fuel), successful strategies focused on the

improvement of household air quality should improve ventilation practices and

appropriate construction materials.

According to results presented in chapters 4 and 5, rocket stoves would have a positive

effect with regard to HAP reduction, but more uncertain effects on climate. This proves

that the climate and health-relevant properties of stoves do not always scale together

and highlights that both dimensions should be considered in technological decisions.

Notwithstanding, this research work is focused carbonaceous aerosol emissions, but it

does not include climate impacts from other pollutants emitted or factors such as

biomass renewability, so results do not mean that rocket stove types do not have

climate benefits, as the full picture is much more complicated and uncertain.

The information presented in this thesis helps to have a more accurate picture of the

climate and health impacts of residential burning of biofuels for cooking in the West

African region. Moreover, it gives useful insights for the selection of the more

appropriate technologies and it may be useful for stove designers to optimize

environmental and health co-benefits of cooking solutions.

Further laboratory and field studies of different cooking technologies and fuels are

needed, covering different stove-fuel combinations, in other areas of West Africa, and

also studying the seasonality effect on emissions to have more representative EF data

to be used in emission inventories and climate prediction models.

Page 12: INTO THE SMOKE: A RESEARCH ON HOUSEHOLD AIR ...oa.upm.es/49583/1/CANDELA_DE_LA_SOTA_SANDEZ.pdfCada año, la contaminación del aire interior provoca la muerte prematura de 4,3 millones

vii

List of figures

Figure 1. Global polluting fuel use in 2014. Source: (WHO, 2016) ................................... 2

Figure 2. Household energy for cooking and SDG's linkages. Source: WHO, 2016.......... 3

Figure 3. Potential benefits related to the sustained use of improved and clean cooking

solutions. Source: Adapted from (Mazorra, 2017) ........................................................... 6

Figure 4. Emission rates by stove type and ISO Tiers classification. Source: Adapted

from Putti et al., 2015 ..................................................................................................... 13

Figure 5. Main BC sources by region and sector (2005). Source: adapted from (CCAC,

2014) ............................................................................................................................... 15

Figure 6. Household energy consumption for cooking by fuel in Sub-saharan Africa.

Source: IEA, 2014 ............................................................................................................ 18

Figure 7.Share of cooking fuel use in West African countries. Source: own elaboration

with data from (Saho and Reiss, 2013) .......................................................................... 19

Figure 8. Laboratory Emission Monitoring System (LEMS). Source: Aprovecho Research

Centre ............................................................................................................................. 24

Figure 9. Schematic of the gravimetric system of the LEMS. Source: Aprovecho

Research Center ............................................................................................................. 24

Figure 10. Nexleaf BC reference card. Source: Nexleaf Ltd............................................ 26

Figure 11. BC determinations (µg/cm2) by cell-phone system plotted against EC loads

(µg/cm2) by TOT. Source: own elaboration .................................................................... 29

Figure 12. EC load (µg/cm2) determined by TOT plotted against reflectance readings

from the smoke stain reflectometer. Source: own elaboration .................................... 33

Figure 13. Comparison of BC concentrations determined by the reflectometer (right)

and Nexleaf system (left) on quartz filters and co-located glass fiber filters. Source:

own elaboration ............................................................................................................. 33

Figure 14. Glass fiber filters Hi-Q environmental products, FPAE-102, 0.3 μm

aerodynamic equivalent diameter. Source: own elaboration ....................................... 34

Figure 15. Quartz fibre filters Pallflex, tissuquartz, 2500 QAT-UP. Source: own

elaboration ..................................................................................................................... 34

Figure 16. Typical kitchen in Bibane, rural village in the region of Fatick, Senegal.

Source:own elaboration. ................................................................................................ 38

Figure 17. Portable Emission Monitoring System (PEMS), designed by the Aprovecho

Research Centre. Source: own elaboration. ................................................................... 39

Figure 18. Cookstoves tested in the study (from left to right: three stones, Noflaye

Jegg, Jambaar Bois and gasifier Prime Square. Source: own elaboration. ..................... 40

Figure 19. Laboratory estimation of EC emission using the WBT.. Source: own

elaboration ..................................................................................................................... 42

Figure 20. Laboratory estimation of OC emission using the WBT. Source: own

elaboration. .................................................................................................................... 45

Figure 21. Field estimation of EC and OC emissions. Source: own elaboration............. 49

Page 13: INTO THE SMOKE: A RESEARCH ON HOUSEHOLD AIR ...oa.upm.es/49583/1/CANDELA_DE_LA_SOTA_SANDEZ.pdfCada año, la contaminación del aire interior provoca la muerte prematura de 4,3 millones

viii

Figure 22. Boxplots of EC EF (g/MJ) for each type of stove and test. Source: own

elaboration. .................................................................................................................... 54

Figure 23. Figure 23a: 100 years Global Warming Potential (from Elemental carbon and

Organic Carbon emitted) per year of use of Noflaye Jeg stove and Three stones;

Figure23b: 100 years Global Warming Potential (from Elemental carbon and Organic

Carbon emitted) per year from household fuelwood cookstoves in Senegal. Results are

presented for a rate exfiltration of 26% and 80%. Source: own elaboration. ............... 56

Figure 24. Laboratory comparison. Relationship between DustTrak monitors (code:

DustTrak 2 ) and OPC (GRIMM) concentrations (µg/m3) for the period 12/02/2016-

26/02/2016 at the Barcelona (BCN) site. Source: own elaboration .............................. 67

Figure 25. Laboratory comparison. Relationship between DustTrak monitors (code:

DustTrak 3) and OPC (GRIMM) concentrations (µg/m3) for the period 12/02/2016-

26/02/2016 at the Barcelona (BCN) site. Source: own elaboration .............................. 67

Figure 26. Laboratory comparison. Relationship between Microaeth (AE51_787,

AE51_ACS and AE51_5) and MAAP concentrations (µg/m3) obtained for the period

12/02/2016-26/02/2016 at the BCN site. Source:own elaboration .............................. 68

Figure 27. Laboratory comparison. Relationship between DISCmini (MDS, MD impact)

and CPC concentrations (pt/m3) obtained for the period 12/02/2016-26/02/2016 at

the BCN site. Source: own elaboration .......................................................................... 68

Figure 28. Field intercomparison. Relationship of DustTrak monitors (DST2 and DST3)

concentrations obtained at the household BI-01-A. Source: own elaboration. ............ 68

Figure 29. Field intercomparison. Relationship of Microaeths monitors (AE51_5,

AE51_ACS and AE51_787) concentrations obtained at the household BI-01-A. Source:

own elaboration. ............................................................................................................ 69

Figure 30. Field intercomparison. Relationship of Discmini monitors (Impact and MDS)

concentrations obtained at the household BI-01-A. Source: own elaboration ............. 69

Figure 31. PM1, PM2.5, PM10 (µg/m3), BC (µg/m3), UFP (/m3), mean diameter (nm) and

CO (ppm) concentrations measured on March 3, 2016 in a household with rocket

cookstove Noflaye Jegg, during lunch and dinner preparation at household BI-02-A .

Source: own elaboration ................................................................................................ 71

Figure 32. PM1, PM2.5, PM10, BC, UFP, Size, LDSA and CO concentrations measured

during lunch preparation on March 3 2016 in a household with improved cookstove

Noflaye Jegg at household BI-02-A. Source: own elaboration ...................................... 73

Figure 33. Box-plots of pollutants concentrations for three-stone fire and the rocket

stove. The boxplot presents minimum, first quartile, median, (middle dark line), third

quartile and maximum values. Source: own elaboration. ............................................. 79

Figure 34. Correlation of DustTrak response with IAP meter for PM2.5 concentrations

(µg/m3) at 1 min resolution. Source: own elaboration .................................................. 82

Page 14: INTO THE SMOKE: A RESEARCH ON HOUSEHOLD AIR ...oa.upm.es/49583/1/CANDELA_DE_LA_SOTA_SANDEZ.pdfCada año, la contaminación del aire interior provoca la muerte prematura de 4,3 millones

ix

List of tables

Table 1. Technical requirements and features of the methods evaluated during the

intercomparison. Source: own elaboration.................................................................... 28

Table 2. Wilcoxon test results and absolute value relative difference between results

from the Nexleaf system and TOT, as a function of aerosol type and EC load.. Source:

own elaboration ............................................................................................................. 31

Table 3. Summary of correction equations obtained as a function of analytical method,

aerosol and filter type.. Source: own elaboration. ........................................................ 35

Table 4. CO/CO2 ratio (expressed in %) for the stove-fuel combinations analysed in the

laboratory. Source: own elaboration ............................................................................. 42

Table 5. Summary of EC and OC emission factors, EC/OC ratios and total emissions per

WBT from the stove-fuel combinations tested in the laboratory .................................. 47

Table 6. Comparison of Measured EC and OC EF for fuelwood combustion with

previous studies. Source: own elaboration. ................................................................... 51

Table 7. Summary of EC, OC and PM Emission Factors, ratios and total emissions from

in-field emission testing for each household included in the study. Source:own

elaboration ..................................................................................................................... 53

Table 8. EC EFs and % difference between laboratory and field data sets.). Source:own

elaboration ..................................................................................................................... 55

Table 9. BC and OC emissions in different regions of the world determined in this study

and in previous publications. Source: own elaboration.. ............................................... 57

Table 10. Summary of participant kitchen characteristics. Source: own elaboration ... 61

Table 11. Participant kitchen characteristics. Improved cookstoves. Source: own

elaboration ..................................................................................................................... 62

Table 12. Participant kitchen characteristics. Traditional cookstoves. Source:own

elaboration ..................................................................................................................... 63

Table 13. Devices installed per household. Traditional cookstove. Source: own

elaboration ..................................................................................................................... 66

Table 14. Devices installed per household. Improved cookstove. Source: own

elaboration ..................................................................................................................... 66

Table 15. Descriptive statistics of indoor air pollutants concentrations by stove type.

n=number of households sampled.. Source: own elaboration ...................................... 76

Table 16. Summary of pollutant concentration during cooking activities reported for

previous studies and within this study. Source: own elaboration ................................. 78

Table 17. Linear Mixed Effect Modelling of log-transformed PM2.5. Source: own

elaboration ..................................................................................................................... 81

Table 18. Field intercomparison. Relationship of IAP meter(5014 and 5025) and

DustTrak (2 and 3) concentrations. DustTrak vs IAP meter ........................................... 83

Page 15: INTO THE SMOKE: A RESEARCH ON HOUSEHOLD AIR ...oa.upm.es/49583/1/CANDELA_DE_LA_SOTA_SANDEZ.pdfCada año, la contaminación del aire interior provoca la muerte prematura de 4,3 millones

x

Abbreviations

ACCES African Clean Cooking Energy Solution Initiative

ANOVA Analysis of variance

BC Black carbon

CCAC Climate and Clean Air Coalition to Reduce Short-Lived Climate Pollutants

CDM Clean Development Mechanism

CH4 Methane

CO Carbon Monoxide

EC Elemental carbon

ECOWAS Economic Community of West African States

ECREEE ECOWAS Centre for Renewable Energy and Energy Efficiency

GACC Global Alliance for Clean cookstoves

GHG Greenhouse gases

GWP Global Warming Potential

HAP Household Air Pollution

IAP Indoor air pollution

IEA International Energy Agency

ISO International Organization for Standardization

LEMS Laboratory emission monitoring system

LMI Low and middle income

LPG Liquefied petroleum gas

MAC Mass absorption cross-section

Mt Million Tonnes

NAMA’s Nationally Appropriate Mitigation Actions

OC Organic carbon

PAH’s Polyaromatic hydrocarbons

PEMS Portable Emission monitoring system

PIC’s Products of incomplete combustion

PM Particulate matter

PM1 Particulate matter 1 micrometers or less in diameter

PM10 Particulate matter 10 micrometers or less in diameter

Page 16: INTO THE SMOKE: A RESEARCH ON HOUSEHOLD AIR ...oa.upm.es/49583/1/CANDELA_DE_LA_SOTA_SANDEZ.pdfCada año, la contaminación del aire interior provoca la muerte prematura de 4,3 millones

xi

PM2.5 Particulate matter 2.5 micrometers or less in diameter

SDG’s Sustainable Development Goals

SE4all Sustainable Energy for All

SLCP’s Short Lived Climate Pollutants

TOT Thermo-optical transmission

UFP Ultrafine particles

UNEP United Nations Environment Programme

UNFCCC United Nations Framework Convention on Climate Change

UNICEF United Nations Children’s Fund

VOC’s Volatile organic compounds

WACCA West Africa Clean Cooking Alliance

WB The World Bank

WBT Water Boiling Test

WHO World Health Organization

Page 17: INTO THE SMOKE: A RESEARCH ON HOUSEHOLD AIR ...oa.upm.es/49583/1/CANDELA_DE_LA_SOTA_SANDEZ.pdfCada año, la contaminación del aire interior provoca la muerte prematura de 4,3 millones

1

1. Introduction

Over half of the world’s population still rely on polluting and inefficient energy systems

to meet their daily cooking needs. This has important social, economic and

environmental consequences. Women are disproportionately affected.

Page 18: INTO THE SMOKE: A RESEARCH ON HOUSEHOLD AIR ...oa.upm.es/49583/1/CANDELA_DE_LA_SOTA_SANDEZ.pdfCada año, la contaminación del aire interior provoca la muerte prematura de 4,3 millones

2

1.1. Household energy for cooking. Global situation and impacts

Nowadays, around 3100 million people rely on solid fuels, like fuelwood, charcoal,

coal, agricultural residues and animal wastes, as the primary source of energy to satisfy

their basic energy needs (WHO, 2016). These fuels are most usually burnt inside homes

either in open fires or poorly efficient traditional stoves, which lead to important

consequences both on families’ living conditions and the environment.

The African Region, the South-East Asia Region, the Western Pacific Region and some

Central American countries have the highest proportions of households primarily using

solid fuels for cooking (Figure 1), with significant differences between urban and rural

areas. Over 20% of urban households rely primarily on polluting fuels and

technologies, while the ratio in rural areas ascends to an 80% (WHO, 2016).

Solid fuels are primarily burned in poorly ventilated spaces with inefficient stoves or

open fires, causing high levels of household air pollution (HAP), the single most

important environmental health risk factor worldwide (WHO, 2016).

Indeed, HAP causes the premature death of 4.3 million people each year, a number

above the deaths produced by malaria and tuberculosis together (WHO, 2014a) and

also leads to many negative health impacts, including low birth weight and stillbirths,

tuberculosis, asthma, respiratory infections, cancers, chronic headaches, cataract, and

burns (Bruce et al., 2014; Ezzati and World Health Organization, 2004; WHO, 2016).

Women are particularly at high risk of disease from exposure to HAP, owing to their

greater involvement in daily cooking and other domestic activities, together with their

children, who are under their care most of the day (Batliwala and Reddy, 2003; Desai

et al., 2004; Torres-Duque et al., 2008) (WHO, 2016).

Figure 1. Global polluting fuel use in 2014. Source: (WHO, 2016)

Page 19: INTO THE SMOKE: A RESEARCH ON HOUSEHOLD AIR ...oa.upm.es/49583/1/CANDELA_DE_LA_SOTA_SANDEZ.pdfCada año, la contaminación del aire interior provoca la muerte prematura de 4,3 millones

3

But the problem does not stay at home. Burning solid fuels in inefficient devices at

household level is a major source of ambient air pollution and releases high emissions

of some important contributors to global climate change, such as carbon dioxide (CO2),

methane (CH4), black carbon (BC), and other short-lived climate pollutants (SLCP’s).

Recent studies estimate that residential solid fuel burning accounts for up to 25% of

global BC emissions (Bond et al., 2013) and that 12% of global ambient fine particulate

matter (PM2.5) emissions come from cooking with solid fuels (Smith et al., 2014).

Furthermore, inefficient combustion for cooking purposes also leads to high levels of

fuel consumption. About one third of global wood fuel harvesting is unsustainable,

meaning that the rate of depletion of renewable biomass outstrips the rate of

regrowth, resulting in a net addition of heat-trapping carbon to the atmosphere (Bailis

et al., 2015).

Women and children, mostly girls, are the primary responsible of gathering fuel for

cooking in most low and middle income (LMI) countries (Shailaja, 2000;

Wickramasinghe, 2003). During this activity, they face safety risks, such as physical and

sexual attacks, in addition to suffer significant constraints on their available time for

education, rest and activities for income generation (WHO, 2016).

These facts lead to the conclusion that the global community must intensify in an

urgent and inescapable manner its efforts to expand and accelerate access to clean

household energy (WHO, 2016). Fortunately, today a growing consensus exits that

expanding access to clean household energy for cooking (together with heating and

lighting) is essential to make progress toward several Sustainable Development Goals

(SDG’s), and presents an enormous opportunity to leverage the synergies currently

offered by initiatives that encompass energy, gender, health and climate change

(Mazorra, 2017; WHO, 2016).

Integrating clean cooking into development agendas can catalyse impacts to end

energy poverty (SDG 7), to reduce global mortality and improving overall wellbeing

(SDG 3), to promote gender equality and women’s empowerment (SDG 5) and SDG 11

and 13, regarding action to fight against climate change and to make cities and human

settlements inclusive, safe, resilient and sustainable (Figure 2) .

.

Figure 2. Household energy for cooking and SDG's linkages. Source: WHO, 2016

Page 20: INTO THE SMOKE: A RESEARCH ON HOUSEHOLD AIR ...oa.upm.es/49583/1/CANDELA_DE_LA_SOTA_SANDEZ.pdfCada año, la contaminación del aire interior provoca la muerte prematura de 4,3 millones

4

1.2. Overview of clean and improved cooking solutions and related

challenges

There is no “one size fits all” when it comes to designing an approach to address the

issue of household energy for cooking. A broad range of stove and fuel solutions have

been developed in order to make the shift from polluting cooking options to clean and

improved ones, each one with differentiated efficiencies, costs, distribution models

and challenges in term of meeting user needs.

Some of these solutions imply a switch to other type of fuels different from the

traditionally used biomass, usually known as clean fuels, as they produce very low or

any indoor air pollution (IAP) in homes (Putti et al., 2015):

LPG stoves: In this stoves, liquefied petroleum gas (LPG) from a canister burns very

cleanly (from an indoor air pollution perspective) and efficiently. A typical LPG cooking

system is made up of a steel cylinder filled with LPG, a pressure controller, a tube

connecting the cylinder to the pressure controller and the burner, and finally the

burner itself. The burner can consist of one or more cooking tops. LPG stoves are very

convenient for users as they heat up quickly and temperature can be precisely

controlled. However, LPG is a fossil fuel, with a measurable carbon footprint.

Solar cookstoves: Generate heat by directly capturing sun’s solar thermal energy. The

fuel is available at no cost, and, as no smoke is produced, solar cookers do not produce

any health or impact associated with cooking. However, these stoves can take more

time to cook, and the efficiency depends of the strength of the sun at any given time.

Moreover, it is very difficult to adapt to some meal times (i.e. it is not possible to cook

at night or breakfast very early in the morning).

Alcohol cookstoves: Ethanol and methanol, pressurized or non-pressurized, typically

achieve high combustion rates, and are therefore very efficient. They emit very low

levels of carbon monoxide (CO), volatile organic compounds (VOC’s) and BC.They can

also produce a carbon footprint. Moreover, the fuel is not widely available and could

be costly.

Biogas stoves: Domestic biogas digesters convert organic wastes (animal manure,

human waste, and crop residues) into biogas that includes methane, which is then

effectively used in simple gas stoves. Biogas typically produces little household air

pollution with no additional greenhouse gases (GHG) emitted to the atmosphere.

However, cultural barriers and maintenance costs difficult the adoption of biogas

stoves in many regions of the world.

Electric cookstoves: This stoves convert electrical energy into heat. Its use is limited to

areas with electricity access, excluding most rural communities. Electric stoves do not

Page 21: INTO THE SMOKE: A RESEARCH ON HOUSEHOLD AIR ...oa.upm.es/49583/1/CANDELA_DE_LA_SOTA_SANDEZ.pdfCada año, la contaminación del aire interior provoca la muerte prematura de 4,3 millones

5

produce emissions at household levels, although environmental impact depends on

the energy source.

During recent years, an increasing access to low cost LPG, electric stoves and other

clean fuel and technologies across Asian, African, and Latin American population, has

been produced, primarily between the middle-class in urban areas. However, previous

experiences shown that most of the time there is no direct transition to these clean

modern fuels. The stacking model, which means the parallel use of multiple types of

fuel and technologies in a single household, offers a much more realistic picture than

the idea of a fixed “energy ladder”, whereby choice of fuel graduates from solid to

non-solid fuels as households get richer (IEA, 2014; Masera et al., 2000; Ruiz-Mercado

et al., 2011; Ruiz-Mercado and Masera, 2015). Indeed, even after gaining access to LPG

or electricity, traditional stoves may be kept in use for heating, to keep mosquitoes

away, to cook certain types of food, to prepare food for their animals, etc.

Moreover, modern cooking appliances, such as LPG or electrical cookers, would be out

of reach in the short-medium term for a large proportion of LMI countries population,

while biomass would continue to be the fuel that people can most easily access and

afford (Foell et al., 2011).

Therefore, gaining access to clean cooking facilities encompasses not only switching to

alternative fuels, but also access to improved biomass cookstoves (fired with

fuelwood, charcoal or pellets) that are more efficient, safer, and more environmentally

sustainable than the traditional facilities, and still using solid biomass (IEA, 2014). A

wide range of biomass stoves have been developed and the most common are as

follows:

Rocket stoves: They consist in a L-shaped combustion chamber where the flow of hot

gases is directed closer to the pot. Production of rocket stoves range from centrally

produced products to locally produced artisanal products, which leads to high

variations in the emission reductions and efficiency achieved. The rocket stove design

is the most commonly type of improved cookstove used in sub-Saharan Africa.

Plancha cookstoves: They are designed to enclose the fire and to exhaust the emissions

from combustion though a chimney. They are specialized for typical gastronomical

practices in Central in South America. As in the case of rocket stoves, their

performance varies greatly depending on design and conditions.

Natural and forced draft gasifiers: They convert biomass into a combustible gas using a

two-stage combustion process: firstly, the biomass fuel is heated and converted into a

combustible gas in the pre-combustion (or gasification) chamber; then the gas is

completely oxidized in the second-stage combustion chamber, resulting in very few

emissions (Roth, 2011). Forced draft gasifiers have a fan to blow air into the

combustion chamber, resulting in a more complete combustion of the biomass fuel.

Page 22: INTO THE SMOKE: A RESEARCH ON HOUSEHOLD AIR ...oa.upm.es/49583/1/CANDELA_DE_LA_SOTA_SANDEZ.pdfCada año, la contaminación del aire interior provoca la muerte prematura de 4,3 millones

6

Performance of these stoves can be improved by using processed biomass fuels, such

as pellets or briquettes.

Improved charcoal stoves: Typically consist of ceramic liner in a metal cladding,

providing a higher efficiency, a hotter flame and an improved combustion when

compared to traditional all metal charcoal stoves. Although performance varies a lot

through design and production characteristics, in general, charcoal stoves emit very

little particulate matter (PM), but considerable CO.

Improved and clean cooking solutions presented above, both using alternative and

biomass fuels, have the potential to bring important and interconnected benefits

(Figure 3), which are highly dependent on the fuel-technology choice, as well as the

type of use and maintenance during daily cooking activities.

Being conscious of the potential benefits achieved through the use of clean and

improved cooking solutions, a wide variety of actors (governments, corporations,

foundations, civil society, individuals…) have been involved in their promotion since

the 1980’s in many regions of the world. However, even though international and

national policies, programmes and interventions in the last 40 years have advanced

solutions, they have failed to substantially alter long-term trends (WHO, 2016).

This is strongly linked with the fact that interventions have traditionally focused in the

installation of stoves, while overlooking a lot of other key factors, such as having a

deep knowledge of user needs and preferences, the complexity linked to behaviour

Figure 3. Potential benefits related to the sustained use of improved and clean cooking solutions. Source: Adapted from (Mazorra, 2017)

Page 23: INTO THE SMOKE: A RESEARCH ON HOUSEHOLD AIR ...oa.upm.es/49583/1/CANDELA_DE_LA_SOTA_SANDEZ.pdfCada año, la contaminación del aire interior provoca la muerte prematura de 4,3 millones

7

changes, or the highly importance of gender roles, inequalities and relative power in

household decision-making.

As main responsible for household energy provision and cooking activities, women are

not only the mainly impacted by problems of traditional used of solid fuels, but

essential in influencing from switching to cleaner cooking options (Austin and Mejia,

2017; Ryan, 2014). However, in patriarchal households, women’s first-hand

experiences of indoor air pollution often carry little sway and their preferences for

healthier cooking options may be trumped by unilateral influence of some men over

house-hold energy decisions (Ryan, 2014; WHO, 2016).

Other barriers to the wider uptake of improved solutions for cooking are related with

the lack of awareness of health hazards and environmental degradation, the

unaffordability of the proposed cooking solutions, the lack of supply and repair

services, as well as religious and cultural believes (Puzzolo, 2013).

Moreover, recent studies showed that many improved biomass cookstoves on the

market today, though a great improvement regarding traditional cooking, still produce

enough PM2.5 and BC emissions to be considered a health and environmental hazard

(Grieshop et al., 2011; Jetter et al., 2012; Kar et al., 2012; Mazorra, 2017; Smith et al.,

2014; Sota et al., 2014).

During the last years, great efforts have been made to understand and overcome the

varied and complex barriers to increase the adoption and sustained use of improved

and clean cooking solutions, with the key support of international initiatives like the

Global Alliance for Clean Cookstoves (GACC), the Climate and Clean Air Coalition

(CCCA), the World Health Organization (WHO), the United Nations Children’s Fund

(UNICEF), the Sustainable Development Goals (SDG’s), or the Sustainable Energy for All

initiative (SE4all).

Together, these organizations are working in developing innovative technological and

finance solutions, increasing the cross-sector cooperation, promoting interdisciplinary

research, expanding the knowledge base of impacts of traditional cooking,

demonstrating the cost-effectiveness of clean cooking interventions, developing the

market sector of clean and improved stove and fuels, monitoring and evaluating the

adoption and effectiveness of the solutions promoted, searching for solutions to

address gendered barriers to the adoption of cooking solutions, enabling policy and

regulatory environment, and promoting the creation of more gender-responsive

policies.

Page 24: INTO THE SMOKE: A RESEARCH ON HOUSEHOLD AIR ...oa.upm.es/49583/1/CANDELA_DE_LA_SOTA_SANDEZ.pdfCada año, la contaminación del aire interior provoca la muerte prematura de 4,3 millones

8

1.3. Research opportunities

Strengthening the base of evidence of the impacts related to the use of pollutant

cooking systems and demonstrating the magnitude of the health, environmental, and

socio-economic benefits of clean cookstoves and fuels is a critical priority that will help

drive investment into solving this issue (Anenberg et al., 2013; Mazorra, 2017).

Hundreds, if not thousands, of stove related innovations have been implemented

around the world (Simon et al., 2014), but not all provide the same benefits. While

these stoves often look simple in appearance, the inherent stochastic nature of

biomass combustion, combined with natural variations in daily use and the wide

variety of fuel-stoves combinations available, make stoves evaluation challenging

(L’Orange et al., 2012).

In order to provide transparency regarding the performance (i.e. efficiency in fuel use,

emissions, safety, etc.) and subsequent potential benefits of different cooking

solutions, standard certification procedures and performance benchmarks are being

developed (Anenberg et al., 2013).

Putting these new guidelines into practice require increasing the evaluation of stoves

and fuels performance under both laboratory and various real-world conditions.

Pollutants different than the traditionally measured PM2.5 or CO, such as ultrafine

particles (UFP), BC and other co-pollutants, which leads to important health and

climate impacts (CCAC, 2014; Kinney, 2015; WHO, 2016)) should be further explored

(Anenberg et al., 2013).

Moreover, the need of studies to quantify coking-related impacts is not homogeneous

across the globe. Traditionally, most studies have been focused in Asia (mainly India

and China), with far few studies in other regions in the world. In the West African

region, for example, where more than 730 million people rely on biomass for cooking

(Saho and Reiss, 2013), very little research has been conducted to measure the impact

of improved and clean cooking solutions.

This issue is particularity significant, as the impacts of available household cooking

options can vary sharply from one region to another (Grieshop et al., 2011). In the case

of BC and other SLCP’s, the impacts on climate and ambient air quality are highly

spatially variable owing to several factors, and thus global-scale assessments may not

properly represent the consequences on national and regional scale (Lacey et al.,

2017).

Page 25: INTO THE SMOKE: A RESEARCH ON HOUSEHOLD AIR ...oa.upm.es/49583/1/CANDELA_DE_LA_SOTA_SANDEZ.pdfCada año, la contaminación del aire interior provoca la muerte prematura de 4,3 millones

9

1.4. Research aim and overview

This thesis aims to contribute to the existing knowledge about climate and indoor air

quality impacts of biomass cookstoves, with a particular emphasis on BC and co-

pollutant emissions, and focusing in the West African region. To that end, this

dissertation addresses several specific research goals, which are presented in the

following chapters.

Chapter 2 frames the context of the research. It develops the key concepts for

understanding the importance of IAP and BC and co-pollutant emissions from

cookstoves, and provides an overview of the situation of household clean energy

access in sub-Saharan and Western Africa.

Chapter 3 presents an inter-comparison of analytical methods to estimate BC

emissions from cookstoves. The performance of two relatively low-cost and easy-to-

use methods is validated against a more accurate device, with two different aerosols

types and different filtrates substrates. This validation helps to overcome the current

existing methodological barriers to determine BC emissions from cookstoves in

resource-constrained locations.

Chapter 4 presents a characterization, both in field and laboratory conditions, of BC

and organic carbon (OC) emissions from a set of stove-fuel combinations used in

Senegal and neighbouring countries. This novel information helps to have a more

accurate picture of the climate impacts of residential burning of biofuels for cooking in

the West African region.

Chapter 5 sets out the first study developed in Senegal to provide a field assessment of

household concentrations of PM2.5, UFP, BC and CO from the combustion of fuelwood

for cooking. The findings add to the small number of quantitative studies on cook-

smoke exposures in West-Africa and establish the bases for interventions to reduce

IAP and corresponding adverse health impacts in this region.

Chapter 6 shows the primary contributions of the research work, and Chapter 7

concludes this thesis by outlining the limitations of the study and future research

areas.

Page 26: INTO THE SMOKE: A RESEARCH ON HOUSEHOLD AIR ...oa.upm.es/49583/1/CANDELA_DE_LA_SOTA_SANDEZ.pdfCada año, la contaminación del aire interior provoca la muerte prematura de 4,3 millones

10

1.5. Personal motivation

My interest on cooking energy started in January 2013 in the region of Casamance

(Senegal), where I had the chance to participate in an initiative of improved stoves

implementation. When I first visited those rural communities, I felt really impressed by

the scale and severity of problems related to the daily and essential action of cooking:

enormous levels of smoke and heat inside cooking areas, where women spent around

5-6 hours per day, cost and time demanded for gathering fuel, etc.

When I came back to Spain, I decided to start a PhD on this issue to contribute with

solutions to that critical situation, which is still minimally discussed at the energy and

environment arenas in comparison with other advanced energy technologies,

commercial fuels, and large-scale power plants.

Another motivation to conduct this research was to make this problem visible,

because, despite of its severity, it still unknown for many people. The use of polluting

technology to cover household energy needs exacerbates gender inequality it’s a big

impediment to the personal development of many people and also and impediment

for humanity to aspire to an equitable and sustainable development. This situation can

and must change.

Lastly, I chose this research because it was an opportunity to look at the human

dimension of energy use and environmental change, which is, surprisingly, not very

commonly studied in engineering studies.

Page 27: INTO THE SMOKE: A RESEARCH ON HOUSEHOLD AIR ...oa.upm.es/49583/1/CANDELA_DE_LA_SOTA_SANDEZ.pdfCada año, la contaminación del aire interior provoca la muerte prematura de 4,3 millones

11

2. Research context

Page 28: INTO THE SMOKE: A RESEARCH ON HOUSEHOLD AIR ...oa.upm.es/49583/1/CANDELA_DE_LA_SOTA_SANDEZ.pdfCada año, la contaminación del aire interior provoca la muerte prematura de 4,3 millones

12

2.1. Cookstoves and household air pollution

Under ideal conditions, the combustion of any fuel produces only useful heat, CO2 and

water. But household combustion is far from ideal, rather always incomplete and often

quite inefficient (WHO, 2016). During burning of solid fuels for cooking activities,

products of incomplete combustion (PIC’s), such as PM, CO, poly-aromatic

hydrocarbons (PAH’s), VOC’s, BC and many other substances are emitted. All of them

are toxic to human beings in different ways and to varying degrees, turning the home

into one of the most health damaging environments (WHO, 2016).

In dwellings with poor ventilation, 24-hour levels of PM2.5 can reach 2,500 μg/m3, i.e.

100 times the levels recommended as safe by WHO guidelines (WHO, 2006), and peaks

during cooking may be as high as 20,000 μg/m3 (see chapter 5). Exposure to IAP causes

4.3 million premature deaths each year and many adverse health outcomes, such as

low birth weight and stillbirths, tuberculosis, asthma, respiratory infections, cancers,

chronic headaches, cataract, and burns (Bruce et al., 2014; Ezzati and World Health

Organization, 2004; WHO, 2016)

Gender roles in the household are major determinants of relative health risks to IAP

(WHO, 2015). Women and children are at a particularly high risk of diseases, as they

are the ones who spend most time in and around the home (WHO, 2016). IAP is the

second largest environmental risk of non-communicable diseases in women of LMI

countries , and it is the case of over half of childhood pneumonia deaths (the largest

cause of death in children under 5 years) (WHO and EPA, 2016).

Moreover, solid cook fuel is sufficiently polluting and widespread to appreciably affect

ambient (outdoor) air pollution (Smith et al., 2014). Indeed, recent studies estimated

that IAP from cooking accounts for up to 12% of global ambient PM2.5 pollution (Smith

et al., 2014). Moreover, in some countries as India and China, as much as 30%

of ambient air pollution is caused by household emissions(Zhang and Kotani, 2012) .

In view of this situation, in 2014 the WHO published the “Guidelines for indoor air

quality: household fuel combustion”, specifying targets for emissions rates for CO and

PM2.5 from household fuels and energy devices. These guidelines were developed with

a pragmatic approach providing both interim targets (for 60% of homes meeting the

targets) and aspirational targets (for 90% of homes meeting the targets). Moreover,

the document provides guidance on the policy for a transition to a sustained adoption

of clean fuels and technologies and emphasizes the importance of addressing all main

household energy end uses for health benefits (WHO, 2014b).

These guidelines are also serving as the basis for an ongoing process to develop

international cookstove standards by the GACC in collaboration with the International

Organization for Standardization (ISO). The ISO process is still ongoing, but it has

Page 29: INTO THE SMOKE: A RESEARCH ON HOUSEHOLD AIR ...oa.upm.es/49583/1/CANDELA_DE_LA_SOTA_SANDEZ.pdfCada año, la contaminación del aire interior provoca la muerte prematura de 4,3 millones

13

already developed a framework for rating stoves from a baseline (Tier 0) to an

aspirational situation (Tier 4), considering four indicators: fuel efficiency, emissions

(total and indoor) and safety (GACC, 2012). PM2.5 and CO emissions rates that define

Tier 4 were determined based on the WHO Guidelines, while other Tier boundaries for

Indoor Emissions were defined to be progressively lower relative to Tier 4.

Figure 4 shows that liquid and gas fuels such as LPG, propane, biogas, and alcohol,

offer very low PM2.5 and CO emission rates. Well performing gasifiers are the biomass

stoves providing the lowest emissions of both PM2.5 and CO. Improved stoves using

biomass fuels, such as the rocket stove, may perform better in terms of emissions over

the baseline of a traditional stove, but it may not produce low enough emissions at

point-of-use to result in meaningful health benefits (WHO, 2016). Charcoal stoves

typically have lower PM2.5 emissions than rocket or traditional stoves, but can create

dangerous levels of CO.

At this point, the challenge is to select cooking solutions that are affordable and

acceptable to households; and, at the same time, sufficiently clean and able to

effectively displace traditional fires to achieve dramatic emission reductions (Martin et

al., 2014). Given that the alternative of providing use of clean fuels such as LPG may

not be realistic for a big portion of population in LMI countries in the short-medium

term (Martin et al., 2014), improved and clean biomass-burning stoves can be an

important step towards cleaning up indoor air (WHO, 2016, p. 20).

In addition to emissions, a variety of factors affect pollutant concentrations levels in a

closed space: ventilation, spatial characteristics, emission from external sources,

emissions that can re-infiltrate, combination of different stoves, etc. (WHO and EPA,

Figure 4. Emission rates by stove type and ISO Tiers classification. Source: Adapted from Putti et al., 2015

Page 30: INTO THE SMOKE: A RESEARCH ON HOUSEHOLD AIR ...oa.upm.es/49583/1/CANDELA_DE_LA_SOTA_SANDEZ.pdfCada año, la contaminación del aire interior provoca la muerte prematura de 4,3 millones

14

2016). Therefore, strategies and interventions to reduce IAP should focus on reducing

the emission source as much as possible through a switch to clean and improved

cooking solutions, while promoting adequate ventilation practices (chimneys, kitchen

materials, windows, etc.) and ensuring a correct and sustained use of technologies

(WHO, 2014b).

The key to any of these strategies is to develop a monitoring and evaluation system

that documents stove use, emissions, indoor air concentration of pollutants and

exposure levels in and around the household (Martin et al., 2014). Moreover,

additional parameters different than PM2.5 and CO, such as particle composition,

number of particles, and surface area, should also be further explored (Anenberg et al.,

2013).

2.2. Carbonaceous aerosol emissions from cookstoves

Black carbon (BC), or Elemental Carbon (EC), is a carbonaceous aerosol formed from

the incomplete combustion of biomass and fossil fuels. It is a SLCP, with days to weeks

of lifetime in the atmosphere, and considered a dangerous air pollutant and the

second biggest climate forcer after CO2 (Jacobson, 2000).

BC refers to the light-absorbing carbon, measured by change in light transmittance or

reflection, whereas EC is measured by thermal evolution under high-temperature

oxidation (Venkataraman et al., 2005). Even if EC and BC are not exactly equivalent

(Petzold et al., 2013), both terms are often used interchangeably (Bond and Bergstrom,

2006; Chen et al., 2005; Li et al., 2009; Shen et al., 2010; Venkataraman et al., 2005).

BC contributes to climate change in a variety of ways: when airborne, it absorbs solar

radiation, stimulating earth warming. When deposited on ice and snow, BC decreases

the surface albedo, making them less reflective and more light absorbent. In this way,

it accelerates the melting effect, especially in highly vulnerable environments as the

Arctic and in glaciated regions like the Himalayas (Li et al., 2016).

Furthermore, BC causes impacts on environmental conditions at local and regional

scale, such as visibility reductions and changes in rainfall patterns, which can be

substantially larger than global impacts (Ramanathan and Carmichael, 2008).

Each region of the world has a unique mix of natural and pollution carbonaceous

aerosol sources. In many regions, such as North America, Latin America, and Europe,

diesel combustion for transportation is the dominant contributor to BC (Bond, 2004),

while 60–80% of total BC emissions within Asia and Africa comes from residential

burning of biofuels, mainly for cooking purposes (Bond et al., 2013) (see Figure 5).

Nowadays, global inventories indicate that residential sources are responsible for over

25% of BC emissions worldwide (Bond et al., 2013), although important regional

Page 31: INTO THE SMOKE: A RESEARCH ON HOUSEHOLD AIR ...oa.upm.es/49583/1/CANDELA_DE_LA_SOTA_SANDEZ.pdfCada año, la contaminación del aire interior provoca la muerte prematura de 4,3 millones

15

variations in emissions are expected in the coming decades, with decreases of up to

50% in North America and Europe due to mitigation measures in the transport sector,

and significant increases in Asia and Africa (CCAC, 2014). By 2030, BC from traditional

bioenergy use in Asia and Africa is expected to contribute to close to half of all global

BC emissions (UNEP and WMO, 2011).

BC is always co-emitted with organic carbon (OC), the other major component of

carbonaceous aerosols, which has been traditionally linked to light scattering

properties with consequent cooling effects in the atmosphere (Chung, 2002).

However, OC has recently been found to be a source of light absorption in the

atmosphere due to the presence of brown carbon, which strongly absorbs radiation in

the ultraviolet wavelengths (Bond et al., 2007; Chung et al., 2012; Feng et al., 2013;

Saleh et al., 2014). The ratio of BC to OC, the portion of brown carbon and other co-

pollutants with cooling effects (such as sulphate aerosols) depend on the source of

emissions and determines the net warming or net cooling effect (CCAC, 2014).

In addition to climate effect, BC is a powerful air pollutant with detrimental impacts on

public health. Recent studies have reported significant associations between long-term

exposure to black carbon and all-cause and cardiopulmonary mortality, cardiovascular

and respiratory diseases (Janssen et al., 2012, 2011).

Therefore, addressing BC emissions would improve the chances to remain global

temperature rise below the maximum target of two degrees Celsius set under the

international climate negotiations (UNEP and WMO, 2011), while providing significant

Figure 5. Main BC sources by region and sector (2005). Source: adapted from (CCAC, 2014)

Page 32: INTO THE SMOKE: A RESEARCH ON HOUSEHOLD AIR ...oa.upm.es/49583/1/CANDELA_DE_LA_SOTA_SANDEZ.pdfCada año, la contaminación del aire interior provoca la muerte prematura de 4,3 millones

16

health benefits. Moreover, short BC lifetimes mean that climate benefits would be

achieved quickly after mitigation (CCAC, 2014; WHO, 2016).

In recent years there has been an increase in studies and institutions addressing BC’s

impacts. An example of this is the creation of the Climate and Clean Air Coalition to

Reduce Short-Lived Climate Pollutants (CCAC), a voluntary partnership of

governments, intergovernmental organizations, businesses, scientific institutions and

civil society organizations committed to reduce SLCP’s, hosted by the United Nations

Environment Programme (UNEP) and the WHO. The CCAC has proposed sixteen cost-

effective control measures to reduce BC and other SLCP’s (CCAC, 2014; WHO, 2016),

one of them being the replacement of traditional cooking with modern fuels and

clean-burning biomass stoves.

Another remarkable initiative is the development of a new methodology for the

quantification of BC and co-emitted species from cookstove projects, promoted by the

Gold Standard Foundation and a group of cookstove organizations (The Gold Standard,

2015). These pollutants are not covered by the existing methodologies used for Clean

Development Mechanism (CDM) projects and other mechanisms established under the

United Nations Framework Convention on Climate Change (UNFCCC), which can result

in incorrect judgments about the relative benefits of clean and improved stoves

solutions (Johnson et al., 2008a).

The new methodology will award a certification, which should not be confused with

carbon offsets. Organizations can not compensate emissions of GHG, like CO2, by

reducing BC elsewhere, since BC impacts are highly localized depending on the region,

season, and local weather. Meanwhile, inclusion of BC co-benefits in cookstoves

projects may provide an opportunity to attract new additional funding for further

scaling up of these activities (Hamrick, 2015).

However, despite the ongoing efforts to increase capacity for testing BC and other

SLCPs, most biomass stoves technology worldwide have not been tested or proved to

reduce BC yet (Jetter, 2015). The complexity of studying aerosol emissions from

cooking is related to the wide distribution and remote location of the sources (Johnson

et al., 2008b), the complicated effect of OC (Feng et al., 2013), the relatively complex

BC and OC measurement methods (de la Sota et al., 2017; Ramanathan et al., 2011b),

and because EC/BC and OC emissions range widely between different cooking

technologies and biofuels (Grieshop et al., 2011; Guofeng et al., 2012; MacCarty et al.,

2008b; Venkataraman et al., 2005)

As a consequence, further studies should be conducted to provide concrete scientific

data allowing estimation of the net climate effect of different improved and clean

cooking solutions in different regions worldwide.

Page 33: INTO THE SMOKE: A RESEARCH ON HOUSEHOLD AIR ...oa.upm.es/49583/1/CANDELA_DE_LA_SOTA_SANDEZ.pdfCada año, la contaminación del aire interior provoca la muerte prematura de 4,3 millones

17

2.3. Household energy for cooking in sub-Saharan and Western Africa

In many parts of the world, the use of biomass as energy source for cooking has

generally peaked or will do so in the near future (Sander et al., 2011a). In sub-Saharan

Africa, however, biomass is the main fuel used for cooking for 80% of the population

(i.e. 700 million people) and will likely be for 880 million by 2020 (IEA, 2014).

Electricity and LPG are used exclusively for cooking by less than 10% in most African

countries, less than 1% of households cook with kerosene and the vast majority of

countries do not use coal at all (IEA, 2016; Rysankova et al., 2014; WHO, 2016).

When considering rural areas, almost 95% of the population use fuelwood and

charcoal, and just 2% and 3% of households report use of electricity and gas as their

primary cooking fuel (WHO, 2016). In urban areas there is also an important share of

biomass consumption, much as charcoal, combined with LPG, natural gas, and

electricity.

Around 7.5 million Tonnes (Mt) of PM2.5 are emitted annually in Africa, of which almost

three-quarters is from the indoor burning of biomass (IEA, 2016). The inefficient

combustion of solid biomass in traditional stoves makes the residential sector in Africa

the major emitter of PM2.5 worldwide, accounting for more than 15% of all energy-

related PM2.5 emitted today (IEA, 2016).

In sub-Saharan Africa, IAP exposure is the single greatest health risk for women and

girls (WHO, 2016), causing around half a million of premature deaths annually (IEA,

2016), half of them being children under five (Dherani et al., 2008). If no action is

taken, by 2030, 870,000 people will die annually from diseases linked to solid fuel

cooking (Rysankova et al., 2014).

More than 300 Mt of wood are consumed each year in Sub-Saharan Africa for cooking,

including 130-180 Mt of wood for charcoal production, mostly conducted with

traditional technologies with low wood-to-charcoal conversion efficiencies (Liyama,

2013; Rysankova et al., 2014). Recent data confirmed that woodfuel use in certain

parts of sub-Saharan Africa, Eritrea, western Ethiopia, Kenya, Uganda, Rwanda and

Burundi and some parts of West Africa, is unsustainable (Bailis et al., 2015). In some

African countries, women spend as many as five hours per day gathering fuelwood for

cooking (OECD/IEA, 2014).

Solid fuel use and charcoal production in sub-Saharan Africa generates 120–380 Mt

CO2-eq of Kyoto Protocol greenhouse gases (0.4–1.2% of global CO2 emissions) and up

to 600 Mt CO2-eq when PM is included (Lambe et al., 2015b). Moreover, the change in

rainfall patterns in sub-Saharan Africa has been associated with the increasing burning

of biomass for cooking and land clearing (Ichoku et al., 2016; Ramanahatan, 2007).

Page 34: INTO THE SMOKE: A RESEARCH ON HOUSEHOLD AIR ...oa.upm.es/49583/1/CANDELA_DE_LA_SOTA_SANDEZ.pdfCada año, la contaminación del aire interior provoca la muerte prematura de 4,3 millones

18

According to IEA projections for 2040 (Figure 6) (OECD/IEA, 2014), the mix of fuels

used by household for cooking energy is relatively inelastic. This is in part due to the

rapid population growth expected in this region of the world, which can outstrip the

moderate incremental progress in increasing access to clean, modern energy systems.

Moreover, a transition to cleaner cooking fuels and appliances is not straightforward,

as people who have access to fuels, such as LPG, natural gas, biogas or electricity, may

also continue using solid biomass due to cultural or affordability reasons (fuel stacking)

(OECD/IEA, 2014). Indeed, historical trends over the last 25 years show that population

relying on traditional use of solid biomass has tracked population growth fairly closely,

despite increasing incomes (Rysankova et al., 2014).

Meanwhile, one of the major changes within IEA projections is not captured in Figure

6, because it involves not a fuel switch but a change in the way that solid biomass is

used (OECD/IEA, 2014). Access to more efficient stoves is growing rapidly across sub-

Saharan Africa. For example, penetration rate of rocket stoves and similar improved

stoves doubled from 4 million in 2011 to 8 million by the end of 2013 and advanced

biomass cookstoves and clean renewable cooking alternatives such as biogas, solar and

liquid biofuels cumulatively reached 1.3 million African families by late 2013

(Rysankova et al., 2014). Many improved cookstove businesses are already operating,

and their numbers are growing as cookstove technologies advance and innovations in

end-user finance make stoves more affordable (Lambe et al., 2015b).

But, despite these progress, the overall penetration of clean cookstoves and fuels in

sub-Saharan Africa is still low, with only one in six households using clean cooking

energy for most cooking needs (Lambe et al., 2015b), in partly due to the lack of

finance resources in African households. Many cookstove programme implementers

have seen carbon finance as an interesting revenue option to overcome the finance

barriers in sub-Saharan Africa. Indeed, in 2013, 43% of new funding was attributed to

Figure 6. Household energy consumption for cooking by fuel in Sub-saharan Africa. Source: IEA, 2014

Page 35: INTO THE SMOKE: A RESEARCH ON HOUSEHOLD AIR ...oa.upm.es/49583/1/CANDELA_DE_LA_SOTA_SANDEZ.pdfCada año, la contaminación del aire interior provoca la muerte prematura de 4,3 millones

19

carbon offset sales ($45 million) or carbon fund investments ($3.6 million) and the sale

of carbon offsets drove the distribution of 1 million cookstoves (GACC, 2014; Lambe et

al., 2015b).

In addition to economic revenues, one of the most recognized benefits of carbon

finance is that, as it requires rigorous monitoring and evaluation, it helps to follow up

the cookstove use, and enhance quality assurance on the production line (Lambe et al.,

2015b). However, demand for carbon credits is currently minimal, so cookstove

projects have to develop sustainable business models or other alternative source of

revenue to support their operations (Lambe et al., 2015a).

Within the Economic Community of West African States (ECOWAS), which represents

approximately one third of sub-Saharan Africa’s total population (Auth and Musolino,

2014), around 75% of population still uses wood and charcoal for cooking, often in

inefficient stoves (Saho and Reiss, 2013). It is estimated that quite minor shares of the

population (e.g. Sierra Leone (10%), Senegal (16%) and the Gambia (20%)) are using

improved biomass cookstoves and less than 25% of inhabitants rely on electricity,

kerosene and LPG for cooking (Auth and Musolino, 2014). As a consequence, IAP

affects 257.8 million people and cause the prematurely death of 174,000 people each

year (Auth and Musolino, 2014).

The fact that West Africa is one of the regions with a largest expected population

increase (OECD/IEA, 2014) magnifies common challenges existing in other sub-Saharan

African countries to achieve modern energy access for cooking. But, at the same time,

West Africa is experiencing the most rapid economic growth of sub-Saharan regions

(OECD/IEA, 2014), bringing potential opportunities for the development of clean

cooking sector.

Indeed, a wide variety of initiatives in the region are promoting access to cooking

solutions. In 2012, the ECOWAS Centre for Renewable Energy and Energy Efficiency

63% wood

13% charcoal

23% LPG, kerosene and electricity

2% others

Figure 7.Share of cooking fuel use in West African countries. Source: own elaboration with data from (Saho and Reiss, 2013)

Page 36: INTO THE SMOKE: A RESEARCH ON HOUSEHOLD AIR ...oa.upm.es/49583/1/CANDELA_DE_LA_SOTA_SANDEZ.pdfCada año, la contaminación del aire interior provoca la muerte prematura de 4,3 millones

20

(ECREEE) initiated a regional Cooking Energy initiative called West African Clean

Cooking Alliance (WACCA) with the objective of promoting the adoption of clean and

efficient cooking devices for all ECOWAS households, through policy and regulatory

design, capacity building, and technology promotion. Other initiatives in the region

include the World Bank-led African Clean Cooking Energy Solution Initiative (ACCES),

which aims to promote market-based, large scale dissemination of clean cooking

solutions in sub-Saharan Africa.

Page 37: INTO THE SMOKE: A RESEARCH ON HOUSEHOLD AIR ...oa.upm.es/49583/1/CANDELA_DE_LA_SOTA_SANDEZ.pdfCada año, la contaminación del aire interior provoca la muerte prematura de 4,3 millones

21

3. Inter-comparison of methods to measure black carbon from

cookstoves

Page 38: INTO THE SMOKE: A RESEARCH ON HOUSEHOLD AIR ...oa.upm.es/49583/1/CANDELA_DE_LA_SOTA_SANDEZ.pdfCada año, la contaminación del aire interior provoca la muerte prematura de 4,3 millones

22

3.1. Background

As discussed in previous chapters, emissions of BC resulting from residential burning of

biofuels is a climate and health global concern (Arora and Jain, 2015), and its

quantification is critical to understand and evaluate the effectiveness of BC mitigation

actions, such as the introduction of cleaner and more efficient cooking technologies.

However, BC emissions from cookstoves have been less characterized than those from

other common sources, such as diesel engines (Johnson et al., 2008a). In recent years,

the number of studies measuring biofuel combustion emissions increased

(Bhattacharya et al., 2002; Fan and Zhang, 2001; MacCarty et al., 2008b; Zhang et al.,

2000) but only a limited number of them analyzed BC cookstove emissions using

optical and thermal-optical methods (Johnson et al., 2008a; Patange et al., 2015;

Ramanathan et al., 2011a; Rehman et al., 2011; Venkataraman et al., 2005).

The scarcity of studies on biofuel emissions from cookstoves and, more specifically, on

BC emissions, is due to the complexity related to measuring emissions from such a

heterogeneous and diverse activity (MacCarty et al., 2008b). In general, pollutant

measurements (including BC) in rural and resource-constrained areas are limited by

the lack of adequate instrumentation, power supply and reproducible environmental

conditions. Other causes are the wide distribution and remote location of these stoves,

as well as the relatively invasive and complex assessment methods available (Johnson

et al., 2008a). Very often, BC measurement techniques pose a challenge in terms of

their high upfront cost and level of technical expertise required for operating them

(Rehman et al., 2011).

In this chapter, the performance of two low-cost and easy to use methods for BC

estimation were assessed: reflectometry, and a cell-phone-based system (in general,

any camera-based system). These methods have the potential to be used for studies in

resource-constrained locations due to their low cost, portability and low technical

requirements. Moreover, their easy use can open up possibilities of involving

cookstove users in the data collection process, which would result in awareness raising

on this topic (Nelms et al., 2016). Although there are several examples of citizen

science projects (Castell et al., 2015; Kaufman et al., 2014), this approach has rarely

been used in cookstoves studies. The increased use of mobile phones can bridge the

gap in resource constrained locations, where other infrastructure is lacking.

The performance of the lower-cost methods was assessed by comparison with the

European reference method for EC quantification on filter samples (CEN TC264-WG35),

thermo-optical transmission (TOT) analysis (Cavalli et al., 2010) which, due to its

relative complexity and cost, is in general not applicable for field work in situations

with limited means. The results obtained with the three techniques were compared for

different aerosol types and loads, as well as for different filter substrates.

Page 39: INTO THE SMOKE: A RESEARCH ON HOUSEHOLD AIR ...oa.upm.es/49583/1/CANDELA_DE_LA_SOTA_SANDEZ.pdfCada año, la contaminación del aire interior provoca la muerte prematura de 4,3 millones

23

The aim was to assess the performance of the reflectometer and cellphone optical

analysis when compared to the thermal-optical transmission (TOT) analyzer. If proved

to be comparable within a given uncertainty range, these methods could be

recommended as lower-cost tools for estimation of BC from cookstove emissions in

resource-constrained locations.

3.2. Methods

3.2.1. Sample collection

Aerosol samples were collected in two locations, Dakar (Senegal) and Barcelona

(Spain). Samples were collected to obtain a range of BC and EC loads (from low to high)

and to assess the influence of different aerosol types on measured BC and EC

concentrations. This enables a more robust comparison of the three techniques

(Ahmed et al., 2009). Samples were collected on quartz fibre filters (Pallflex,

tissuquartz, 2500 QAT-UP) and glass fibre filters (Hi-Q environmental products, FPAE-

102, 0.3 μm aerodynamic equivalent diameter).

In Dakar, samples were collected at the Research and Studies Centre on Renewable

Energies of the University of Dakar (CERER, UCAD) using a Laboratory Emissions

Monitoring System (LEMS), developed by the Aprovecho Research Centre (ARC) (Figure

8). This system was designed to sample cookstove emissions after dilution. In this way,

the aerosol concentrations sampled may be considered as a good approximation of

indoor emissions in a typical rural household in Senegal during the cooking hours. In

Barcelona, on the other hand, aerosol concentrations were sampled in outdoor air at

an urban background air quality monitoring station located in the vicinity of traffic

emissions (IDAEA-CSIC Station).

Both types of samples were representative of markedly different aerosol types:

biomass burning aerosols for indoor air quality in households using solid fuels as

primary source of energy in Senegal, and urban aerosol dominated mainly by traffic

emissions, characteristic of ambient air in a typical European city, Barcelona.

In the LEMS in Senegal, the stove was placed under a hood structure to collect exhaust

flows produced during the combustion. The LEMS was equipped to measure CO, CO2,

and PM2.5 in real-time. It had also a gravimetric system to measure PM2.5 using filter-

based sampling, which consists of a vacuum pump that draws a sample through a

sample line and a critical orifice at a steady flow of 16.7 L/min. A cyclone particle

separator was used so that PM2.5 was collected on a filter while the pump was

switched on.

Page 40: INTO THE SMOKE: A RESEARCH ON HOUSEHOLD AIR ...oa.upm.es/49583/1/CANDELA_DE_LA_SOTA_SANDEZ.pdfCada año, la contaminación del aire interior provoca la muerte prematura de 4,3 millones

24

Filters for PM collection were regularly changed as they were quickly saturated with

PM2.5. To solve this, the ARC designed a modification of the gravimetric system by

adding a second sample train in parallel with the gravimetric (PM2.5) sample train, with

a lower flow rating of 3 L/min (Figure 9). This separate BC sample train allowed a

simultaneous collection of PM2.5 and BC samples, as well as the use of different types

of filter media.

The Water Boiling Test (WBT) protocol (version 4.2.3) was applied. It consists of a

standardized test, commonly used in the laboratory, which is a simplified simulation of

the cooking process intending to measure how efficiently a stove uses fuel to heat

water in a cooking pot and the emissions produced while cooking (EPA et al., 2014). In

this test, 5 litres of water are brought to boil (cold start phase) and then simmered for

45 min (simmer phase). Although WBT has recognized limitations to reproduce actual

random conditions in the field (Johnson et al., 2008a; Roden et al., 2006; Shen et al.,

Figure 8. Laboratory Emission Monitoring System (LEMS). Source: Aprovecho Research Centre

Figure 9. Schematic of the gravimetric system of the LEMS. Source: Aprovecho Research Center

Page 41: INTO THE SMOKE: A RESEARCH ON HOUSEHOLD AIR ...oa.upm.es/49583/1/CANDELA_DE_LA_SOTA_SANDEZ.pdfCada año, la contaminación del aire interior provoca la muerte prematura de 4,3 millones

25

2013), its repeatability allows to study the effect of design on the performance of

cookstoves, as well as the emission process influencing factors, mechanisms, kinetics,

etc. (Arora and Jain, 2015; Chen et al., 2012; Shen et al., 2013).

In Barcelona, ambient air PM2.5 samples were collected by means of a MCV CAV-A

(MCV S.A.) high-volume sampler (30m3/h). This type of sampler is certified to be

equivalent to the EU reference for PM2.5 monitoring. The sampler was located at the

IDAEA-CSIC reference station (41°23′14″ N, 02°06′56″E), and samples were collected

over 24h periods.

In total, 102 quartz fibre filters and 81 glass fibre filters were collected in Senegal, and

222 quartz fibre filters were collected in Barcelona. Some of these samples were

removed because BC deposition was below or above analytical limits of detection.

Finally, 73 quartz fibre filters collected in Senegal and 76 in Barcelona were analysed

with the three analytical methods, and 52 glass fibre filters sampled in Senegal were

analysed with the cell-phone system and the reflectometer, given that glass fibre

filters are not recommended for analysis by TOT. Glass fibre filters are generally

preferred to quartz ones in resources-constrained areas due to their lower cost.

3.2.2. Determination of BC and EC on filters

There are different methods to determine the concentration of BC (EC) based on the

chemical, physical and light absorption properties of the particles (Petzold et al., 2013).

The optical methods quantify the light absorbing component of the particles, and are

based on the fact that BC absorbs light due to its colour. The thermal-optical methods

determine the quantity EC based on its stability at high temperatures in an inert

atmosphere. In this work, we used two optical filter-based techniques (reflectometry

and cell-phone based system) and one thermal-optical transmission (TOT) filter-based

technique, all of them offline. The latter determines EC concentrations, whereas the

former two estimate BC concentrations.

Cell-phone based system (Nexleaf). It is an optical technique in which a photograph of

the filter is captured with a cell-phone or a camera and transmitted to a server where

an algorithm compares the image (the colour of the filter) with a calibrated scale

(Figure 10). The BC load is estimated by measuring its reflectance in the red

wavelength, based on the “blackness” of the photograph.

Page 42: INTO THE SMOKE: A RESEARCH ON HOUSEHOLD AIR ...oa.upm.es/49583/1/CANDELA_DE_LA_SOTA_SANDEZ.pdfCada año, la contaminación del aire interior provoca la muerte prematura de 4,3 millones

26

This system is easy to use, cheap and rapid and its performance was already assessed

by its developers (Patange et al., 2015; Ramanathan et al., 2011a). According to the

manufacturer specifications, the measurement range is 5-25 µg BC/cm2, with an

accuracy of 20% (Ramanathan et al., 2011a). It is important that the filter colours are

not in the limits of the reference scale (too white or too black), and photos of the

filters and optical reference cards need to be taken under relatively uniform lighting.

Within a citizen science approach, cookstove users may use the camera on their

mobile phones to capture and transmit the images containing the filter and colour

chart to the server, to automatically compute the daily BC concentrations for each

household. This would only be possible if comparable sampling methods and durations

are used, in order to allow for comparison of BC loadings (µg/cm2) between different

users.

Smoke stain reflectometer (Model 43D, Diffusion Systems Ltd.). The reflectometer is

a non-destructive and portable system, commonly used to determine BC

concentrations (Begum et al., 2007; Biswas et al., 2003; Downward et al., 2015). It has

already been used in several cookstove BC emission studies (Huboyo et al., 2009;

Johnson et al., 2008a). A high-performance LED with maximum emission at 650 nm

shines on the filter, and the reflected light is measured by a photo-sensitive element

(Safo-Adu et al., 2014).

Then, the electrical response is amplified to produce a reflectance reading (RR). The

darker the filters, the lower the amount of reflected light, in such a way that low RR

values correspond to high BC loads and vice-versa. The reflectometer reads on a scale

of 0 (black) to 100 (white), although RR<10 and >90 suffer from major uncertainty

(Adeti, P.J, 2012).

Sunset Laboratory Carbon Analyzer. This instrument determines the elemental carbon

(EC) content on quartz fibre filters by using a thermal-optical method. Carbonaceous

aerosols are thermally desorbed from the filter substrate under an inert helium

atmosphere followed by an oxidizing atmosphere, using controlled heating ramps. The

carbon contained in the sample is detected as CH4 by a flame-ionization detector

Figure 10. Nexleaf BC reference card. Source: Nexleaf Ltd.

Page 43: INTO THE SMOKE: A RESEARCH ON HOUSEHOLD AIR ...oa.upm.es/49583/1/CANDELA_DE_LA_SOTA_SANDEZ.pdfCada año, la contaminación del aire interior provoca la muerte prematura de 4,3 millones

27

(Petzold et al., 2013) while OC and EC carbon are discriminated by monitoring the

optical transmittance through the filter sample. In this work, the EUSAAR2 thermal

protocol was used (Cavalli et al., 2010). Even though differences have been reported

with the use of different temperature protocols (EUSAAR2, NIOSH, IMPROVE) (Cavalli

et al., 2010; Chow et al., 2004) these differences are systematic and may be corrected

when comparing results from studies using different protocols for the same types of

aerosols.

In this work, the thermal-optical analysis was considered as the reference system for

comparison with the two lower-cost optical methods described above. This technique

was recently established as the reference EU method for OC and EC quantification on

filter samples by the CEN standardization working group WG35 (CENTC264-WG35).

3.2.3. Method limitations

A number of limitations must be taken into account in this assessment:

a) Regarding the optical methods, the extent of filter loading can influence particle,

thus biasing the results (EPA, 2012).

b) When using the reflectometer, the device is in direct contact with the filter surface.

This may lead to potential mass loss and/or cross contamination (Yan et al., 2011). The

use of a teflon ring is recommended to protect the filter and mitigate potential impacts

of contamination.

c) The dependence of aerosol absorption on different wavelengths. The reflectometer

works in the visible spectrum, and therefore BC measurements may be affected by

interference with other light absorbing components. In the cell-phone system, BC load

is estimated by measuring its reflectance in the red wavelength, where interference of

other light absorbing components, like brown carbon, should be small or within the

instrument’s uncertainty range (Ramanathan et al., 2011a).

d) For all three methods, results are influenced by the chemical composition and

emissions sources of the aerosol, filter loading and uniformity of the particle deposit

(Watson et al., 2005). BC/EC measurements from biomass smoke may be more

strongly affected than diesel engine samples, in part because of their higher levels of

inorganic components and brown carbon (Novakov and Corrigan, 1995).

e) The methods assessed quantify BC (EC) concentrations, but do not measure organic

carbon (OC). This should be considered a limitation from the point of view of climate

studies, where both OC and BC (EC) data are needed to obtain a reasonable

characterisation of aerosols warming and cooling implications.

Page 44: INTO THE SMOKE: A RESEARCH ON HOUSEHOLD AIR ...oa.upm.es/49583/1/CANDELA_DE_LA_SOTA_SANDEZ.pdfCada año, la contaminación del aire interior provoca la muerte prematura de 4,3 millones

28

3.2.4. Technical requirements and features of the methods

In addition to the performance of the methods regarding their comparability with the

reference, data quality and uncertainty of the measurements, the selection of a

specific method for measuring BC (EC) emissions from cookstoves should also consider

technical requirements. Table 1 includes the main parameters to be evaluated.

(*Prices don’t take into account costs of sampling equipment; the only include costs of

the equipment for filter analysis).

.

Method Robustness Difficulty

of use

Requires a

computer

for

operation

May

operate on

battery

Types of filter

applicable

Accuracy (according

to manufacturers) Price*

Cell-phone

based system

(Nexleaf)

High (only

requires

normal

maintenance

of a cell-

phone or a

camera)

Low

Yes, or a

smartphone,

both with

internet

access

Yes

Paper, Teflon,

glass fibre,

quartz fibre

Low-Medium

(20% agreement with

thermos-optical

method)(Ramanathan

et al., 2011a)

Card with

reference

scale (free)

2 euro/filter

analyzed

Cell-phone

costs

Smoke Stain

Reflectometer

(Model 43D,

Diffusion

Systems Ltd.)

Medium

(requires

easy

maintenance,

cleaning)

Medium

No,

device

readings are

displayed in

the screen

No.

Power

requiremen

ts: 110-

230VAC

50/60HZ

Paper

(recommende

d by the

manufactured

), Teflon, glass

fibre and

quartz fibre

filters

Medium-High

5%

<4.000 euros

(approx.)

Sunset

Laboratory

Carbon

Analyser

Medium-Low

(requires

frequent

maintenance,

laboratory-

based)

Medium-

High

Yes

No.

Power

requiremen

ts:

110-

230VAC

50/60HZ

Quartz fibre

filters

High

(±0,1 µg/cm2 for EC)

>40.000

euros

(approx.)

Table 1. Technical requirements and features of the methods evaluated during the intercomparison. Source: own elaboration

Page 45: INTO THE SMOKE: A RESEARCH ON HOUSEHOLD AIR ...oa.upm.es/49583/1/CANDELA_DE_LA_SOTA_SANDEZ.pdfCada año, la contaminación del aire interior provoca la muerte prematura de 4,3 millones

29

3.2.5. Data analysis

The statistical program STATGRAPHICS (StatPoint Technologies, Inc., 2014) was used

for data analysis. The relationship between the analytical methods was established

using the coefficient of determination (R2), and an analysis of variance (ANOVA) was

used to compare regression lines from the two types of aerosols. Further, the

Wilcoxon signed-rank test (Woolson, 2008) a non-parametric test procedure for the

analysis of matched-pair data, was conducted in order to assess the differences

between results from the cell-phone and the TOT system.

3.3. Results and discussion

3.3.1. BC analysis by the cell-phone system compared with EC by thermo-

optical analysis

Figure 11 presents results for BC load using quartz fibre filters with two different types

of aerosols: biomass aerosols from cookstove emissions, and urban aerosols. For both

types of aerosols, a high degree of correlation between the cell-phone and TOT

methods may be observed (R2>0.80).

Figure 11. BC determinations (µg/cm2) by cell-phone system plotted against EC loads (µg/cm

2) by TOT. Source:

own elaboration

For biomass aerosol samples, linear regression resulted in a slope of 1.35 with a

correlation coefficient (R2) equal to 0.84 (p<0.05), after exclusion of two outliers. In the

case of urban aerosols, the correlation was slightly higher (R2=0.88), with a regression

line slope=0.71 (p<0.05). The systematic error of the methods and the experimental

Biomass cookstove aerosol y = 1.3504x + 0.1897

R² = 0.8356

Urban aerosol y = 0.7066x + 0.594

R² = 0.8812

0

5

10

15

20

25

0 5 10 15 20 25

BC

(ce

ll-p

ho

ne

sys

tem

) (

µg/

cm2 )

EC (Thermal-optical method) (µg/cm2)

Page 46: INTO THE SMOKE: A RESEARCH ON HOUSEHOLD AIR ...oa.upm.es/49583/1/CANDELA_DE_LA_SOTA_SANDEZ.pdfCada año, la contaminación del aire interior provoca la muerte prematura de 4,3 millones

30

error may explain the intercept of the regression curve, although the value is very

close to zero. These results show that, for each different type of aerosol, BC

concentrations measured with the optical cell-phone method are well correlated with

the reference EC concentration range shown in Figure 11 (2-20 µg/cm2).

This good correlation would suggest that the optical cell-phone method could be

considered an option for the estimation of BC concentrations in cookstove studies. It is

essential to note that an initial calibration of the method with a reference instrument

and using EC from local aerosol samples as reference would be an absolute pre-

requisite for this. Any study attempting to use these methods must previously calibrate

them using the same type of aerosol (under the same real-world conditions) as will be

targeted.

ANOVA results showed that slopes obtained from cookstove and urban aerosols were

statistically different, with a confidence level of 95%. The different slopes obtained

result from differences in the physical and chemical composition of the aerosols

studied, mainly their mass absorption cross-section (MAC) (Zanatta et al., 2016). The

MAC is defined as the light absorption coefficient (σap) divided by elemental carbon

mass concentration (mEC), and expressed in units of m2/g (Zanatta et al., 2016).

Different types of aerosols are characterised by different MAC values, especially when

dealing with the combustion of different fuels (Ahmed et al., 2009; Watson et al.,

2005). In particular, wood burning emits large amounts of organic gases along with BC

aerosols during smoldering (Bond et al., 2007; Petzold et al., 2013; WHO and

Scovronik, 2015). The coating of non-absorbing organic (OC) layers on BC particles

enhances the absorption (increases the MAC) (Zanatta et al., 2016), resulting in higher

BC loads by the optical method (Ahmed et al., 2009).

Furthermore, light-absorbing organic matter, also known as brown carbon (Andreae

and Gelencsér, 2006), may have influenced BC measurements. As a reference, in the

present study urban aerosol samples had OC loadings between 5.6 and 22.9 µg/cm2,

and an average OC/EC of 2.45 (±1.14) (for a representative subset of samples). For

biomass aerosols, OC filters loadings were 3.8 - 53.7, and the average OC/EC was 2.61

(±1.78). Thus, OC/EC ratios were similar for both types of aerosols in this study,

suggesting that the relative OC load did not explain the differences in response to

optical methods for the urban and biomass aerosols. However, one very relevant

limitation is that a number of the biomass aerosol filters were not preserved cold

during transport and until analysis, which may have accounted for probable OC losses.

Studies have shown that biomass aerosols have higher light absorption than traffic-

generated particles (Salako, 2012), while others had shown an opposite trend (Jeong et

al., 2004). The discrepancies could be due to the complex interaction of physical

aerosol properties and the filter matrix (Andreae and Gelencsér, 2006).

Page 47: INTO THE SMOKE: A RESEARCH ON HOUSEHOLD AIR ...oa.upm.es/49583/1/CANDELA_DE_LA_SOTA_SANDEZ.pdfCada año, la contaminación del aire interior provoca la muerte prematura de 4,3 millones

31

Another factor possibly affecting the different slopes obtained in Figure 11 could be

particle size, given that the cookstove emissions were fresh particles directly emitted

by the source, whereas the urban particles underwent a certain ageing during

transport from the source (vehicular traffic) to the urban background station.

However, further studies would be necessary to evaluate the potential influence of

particle size.

To assess the statistical differences between both methods as a function of the aerosol

load, the Wilcoxon signed-rank test, a non-parametric statistical hypothesis test

recommended for small sample sizes, was applied. Differences were quantified by

calculating the absolute relative differences between the cell-phone and the TOT

results, for each data point and then averaged (in absolute value) across samples. An

analysis of the differences between the TOT and the cell-phone systems at different

loadings was carried out in order to understand the potential influence of the filter

loading on the optical methods response (Weingartner et al., 2003).

The Wilcoxon signed-rank test indicated that results from both methodologies were,

with 95% significance level, slightly different except for urban aerosols with low EC

load. Table 2 shows that these differences were not large, and that they varied with

aerosol load: larger similarities were observed at higher aerosol loads (with decreasing

relative differences, from 43% to 36% in cookstove aerosols, with increasing particle

load for biomass aerosols, Table 2). The only exception to this was observed for urban

aerosols at low concentrations (non-statistically significant differences). This result

highlights the overall need to take into account the aerosol load, as well as the aerosol

type, when calibrating the cell-phone method.

Filter loading produces two effects on the optical methods response. On the one

hand, the mass loading increases the absorption coefficient (Presser et al., 2014). At

the same time, shadowing effects may occur on the filter surface with heavier filter

mass loading, resulting in an apparent reduced optical path length through the filter

EC (BC)

micrograms/cm2

Wilcoxon test (p<0.05) Absolute value relative differences

Biomass cookstove

aerosol Urban aerosol

Biomass

cookstove

aerosol

Urban aerosol

1 to 5 (N=13) Statistically significantly

differences (Z=3.146 )

Non statistically

significantly differences

(Z= 1.470)

43 % 23 %

5 to 10 (N=40) Statistically significantly

differences (Z=4.631 )

Statistically significantly

differences (Z= 5.520) 40 % 26 %

10 to 20 (N=20) Statistically significantly

differences (Z=3.864 )

Statistically significantly

differences (Z=2.097 ) 36 % 23 %

Table 2. Wilcoxon test results and absolute value relative difference between results from the Nexleaf system and TOT, as a function of aerosol type and EC load. N =number of samples, Z=test statistic. Source: own

elaboration

Page 48: INTO THE SMOKE: A RESEARCH ON HOUSEHOLD AIR ...oa.upm.es/49583/1/CANDELA_DE_LA_SOTA_SANDEZ.pdfCada año, la contaminación del aire interior provoca la muerte prematura de 4,3 millones

32

(Presser et al., 2014; Weingartner et al., 2003) and leading to the underestimation of

the measured optical signals with high filter loads.

In general, in this kind of studies (e.g., cookstove studies) measured concentrations

remain within a given (high) concentration range. Therefore carrying out the

calibrations proposed in this work, tailored to the specific aerosol under study and at

the concentrations expected, should be feasible. In our study, because the differences

are relatively small (between 36-43%) one single correction equation is proposed.

3.3.2. Reflectance analysis by smoke stain reflectometer compared with EC

by thermo-optical analysis

Data are already available in the literature correlating reflectance measurements with

EC values (Adeti, P.J, 2012). However, due to the variability in carbonaceous aerosol

levels and composition associated with different sources (Hinds, 1999), including fuel-

stove combinations, local calibration for the specific type of aerosol under study of

reflectance-based EC measurements is still necessary (Johnson et al., 2008a).

By regressing the EC load (µg/cm2) against the reflectance values from the

reflectometer two calculation equations were obtained (p<0.05). After following a

regression analysis, the best regression model to describe the relationship between RR

values and EC was found to be a reciprocal lineal model. The equation to calculate EC

from RR values determined with the reflectometer in the case of urban aerosol studied

is 1/RR = 0.0034EC+0.0116 (R²=0.9) and, in the case of biomass cookstove aerosol

studied, 1/RR= 0.0081EC-0.006 (R²=0.9). The results in Figure 12 show a clearly low

dispersion of the data points, indicating a high accuracy in the estimation of EC using

RR values.

Page 49: INTO THE SMOKE: A RESEARCH ON HOUSEHOLD AIR ...oa.upm.es/49583/1/CANDELA_DE_LA_SOTA_SANDEZ.pdfCada año, la contaminación del aire interior provoca la muerte prematura de 4,3 millones

33

3.3.3. Analysis of filter substrate on BC and EC quantification

To evaluate the influence of the filter substrate on the performance of the optical

methods evaluated in this study, 16 paired filters (quartz and glass fibre) were

identically loaded using the double-filter system of the LEMS (see Methods section).

The BC load for each type of filter substrate was then analysed using reflectometry and

cell-phone methods, the only two methods which are able to analyse both quartz and

glass fibre filters. The comparison between filter substrates is especially relevant due

to the significantly lower cost of glass fibre filters when compared to quartz fibre ones,

in addition to their lower dependence on ambient humidity (for mass determination

purposes by gravimetry). As a result, glass fibre filters are much more appropriate for

resource-constrained sampling locations.

Biomass cookstove aerosol y = 0.0081x - 0.006

R² = 0.9216

Urban aerosol y = 0.0034x + 0.0116

R² = 0.9293

0,00

0,02

0,04

0,06

0,08

0,10

0,12

0,14

0,16

0,18

0 5 10 15 20 25

1/R

efl

ect

ance

Re

adin

gs (

Smo

ke s

tain

re

fle

cto

me

ter)

(%)

EC (Thermal-optical method) (µg/cm2)

y = 0.9346x - 1.1783 R² = 0.9894

0

10

20

30

40

50

60

70

0 20 40 60 80

Re

fle

ctan

ce R

ead

ings

gla

ss f

ibe

r fi

lte

rs (

%)

Reflectance Readings quartz fiber filters (%)

y = 1.1177x + 0.3463 R² = 0.9541

0

5

10

15

20

25

0 5 10 15 20 25

BC

(N

exl

eaf

sys

tem

) (µ

g/cm

2 )

Qu

artz

fib

re f

ilte

rs

BC (Nexleaf system) (µg/cm2) Quartz fibre filters

Figure 12. EC load (µg/cm2) determined by TOT plotted against reflectance readings from the smoke stain

reflectometer. Source: own elaboration

Figure 13. Comparison of BC concentrations determined by the reflectometer (right) and Nexleaf system (left) on quartz filters and co-located glass fiber filters. Source: own elaboration

Page 50: INTO THE SMOKE: A RESEARCH ON HOUSEHOLD AIR ...oa.upm.es/49583/1/CANDELA_DE_LA_SOTA_SANDEZ.pdfCada año, la contaminación del aire interior provoca la muerte prematura de 4,3 millones

34

Figure 13 shows that optical BC measurements were moderately affected by the filter

substrate. For identically-charged filters, reflectance values on glass fibre filters were

7% lower than for quartz fibre filters, with a significantly low dispersion of the data

(R2=0.99). Regarding the cell-phone system results, the influence of the filter substrate

tended to be larger and with an opposite trend, with BC concentrations on glass fibre

filters being approximately 12% higher than those determined on quartz fibre filters.

Data dispersion, even being still low, was higher (R2=0.94) than by reflectometry.

There is visual evidence that filter characteristics are different. Glass fibre filters are

matted, while quartz fibre ones are fibrous (Figure 15 and Figure 14). Also, they have

different pore size. Previous studies have explained that fibrous filters allow particles

to become partly or completely embedded in the filter (Bond et al., 1999). The

penetration of particles into the filters causes multiple scattering within the filter (Davy

et al., 2017; Presser et al., 2014) and may cause filters being more reflective for a given

aerosol loading than if particles were retained on their surface (Edwards et al., 1983).

This occurs on matted filters, where particles appear to be distributed discretely over

the uneven topography of the matted surface (Presser et al., 2014).

Although a more extensive empirical study could be useful to confirm the influence of

filter substrate, results suggest that, in addition to taking into account the BC source

(fuel and emission process), when calibrating optical methods it is also necessary to

consider the filter substrate. ¡Error! No se encuentra el origen de la referencia.

summarizes calculation and correction equations presented in Figure 11, Figure 12,

and Figure 13.

An external verification of the correlations summarized in ¡Error! No se encuentra el

origen de la referencia. would be recommended. In general, calculation and

subsequent testing of these correction coefficients against a reference system is

recommended for each specific study, given that each specific population is affected

by EC and BC emissions from emission processes, which determine the type of aerosol

Figure 15. Quartz fibre filters Pallflex, tissuquartz, 2500 QAT-UP. Source: own elaboration

Figure 14. Glass fiber filters Hi-Q environmental products, FPAE-102, 0.3 μm aerodynamic equivalent diameter.

Source: own elaboration

Page 51: INTO THE SMOKE: A RESEARCH ON HOUSEHOLD AIR ...oa.upm.es/49583/1/CANDELA_DE_LA_SOTA_SANDEZ.pdfCada año, la contaminación del aire interior provoca la muerte prematura de 4,3 millones

35

generated (Bond, 2004) and thus the correction coefficients. Statistical

representativeness should also be taken into account.

Correction

equation

(TOT vs Optical

method with

quartz fibre filters)

Cell-phone system Smoke stain reflectometer

Biomass cookstove

aerosol Urban aerosol

Biomass cookstove

aerosol Urban aerosol

BC= 1.3504EC+ 0.1897

R² = 0.8356

BC= 0.7066EC+ 0.594

R² = 0.8812

1/RR= 0.0081EC- 0.006

R² = 0.9216

1/RR = 0.0034EC+ 0.0116

R² = 0.9293

Filter material

adjustment

(quartz-glass fibre)

BCglass = 1.1177BCquartz + 0.3463

R² = 0.9541

RRglass = 0.9346RRquartz - 1.1783

R² = 0.9894

Table 3. Summary of correction equations obtained as a function of analytical method, aerosol and filter type. BC and EC are expressed in μg/cm2; RR=Reflectance readings in %. Source: own elaboration.

Page 52: INTO THE SMOKE: A RESEARCH ON HOUSEHOLD AIR ...oa.upm.es/49583/1/CANDELA_DE_LA_SOTA_SANDEZ.pdfCada año, la contaminación del aire interior provoca la muerte prematura de 4,3 millones

36

4. Quantification of carbonaceous aerosol emissions from

cookstoves in Senegal

Page 53: INTO THE SMOKE: A RESEARCH ON HOUSEHOLD AIR ...oa.upm.es/49583/1/CANDELA_DE_LA_SOTA_SANDEZ.pdfCada año, la contaminación del aire interior provoca la muerte prematura de 4,3 millones

37

4.1. Background

As previously discussed in section 2.2, cooking-related aerosols have been much less

studied than those of other common sources (Roden et al., 2006; Soneja et al., 2015),

leading to an important underestimation of climate impacts from traditional burning of

biomass and a lack of knowledge of emission effects due to the switch from traditional

to improved cookstoves (Grieshop et al., 2011; Lee and Chandler, 2013).

Some studies determining BC (EC) and OC EFs from cooking technologies are available

for the Asian Region (Arora and Jain, 2015; Chen et al., 2012; Guofeng et al., 2012; Li et

al., 2009; Shen et al., 2013, 2010; Venkataraman et al., 2005; Zhang et al., 2000); and

some contributions have been made from Latin America (Johnson et al., 2008b; Roden

et al., 2009, 2006). However, only one study has performed in sub-Saharan Africa

(Malawi) (Wathore et al., 2017) and, to our knowledge, there are no studies regarding

the characterization of aerosol emissions from cooking in the West African Region.

This chapter presents a laboratory characterization of aerosol emissions from four

stove-biofuel combinations widely used in Senegal, where 83% percent of households

depend on biomass fuels to cover their daily cooking energy needs (WB, 2013). Two of

these stove-fuel combinations were, in addition, tested in a rural village of Senegal

under uncontrolled conditions to understand how real cooking practices in this region

influence carbonaceous aerosol emissions. Finally, total carbonaceous aerosol

emissions and the net climate effect from fuelwood burning at household level in

Senegal and West Africa were estimated.

A better knowledge of aerosol emissions from residential biofuels in West Africa

should contribute to reduce the uncertainties on emission inventories and climate

prediction models at regional level, and should derive in a more representative and

robust estimation of the climatic impacts of this source. Moreover, this information

should help improved cookstove initiatives to be considered within the national and

regional SLCP’s Reduction Strategies, for addressing both near-term and long-term

climate change impact and, simultaneously, promoting health and social benefits

(Shoemaker et al., 2013).

4.2. Methods

4.2.1. Laboratory sampling

Laboratory tests were conducted at the Centre de Études et Recherches sur les Énergies

Renouvelables (CERER) of the University of Dakar, using a LEMS. Three replicates of

WBT were conducted to test each cooking system. Both LEMS and WBT were already

explained in detail in section 3.2.1.

Page 54: INTO THE SMOKE: A RESEARCH ON HOUSEHOLD AIR ...oa.upm.es/49583/1/CANDELA_DE_LA_SOTA_SANDEZ.pdfCada año, la contaminación del aire interior provoca la muerte prematura de 4,3 millones

38

4.2.2. Field sampling

The study was conducted in March 2016 in Bibane, a rural village with approximately

650 inhabitants (74 families) in the commune of Niakhar, in the region of Fatick

(Senegal) (Figure 16). Typical home structures are made of earth bricks and thatched

roof. The kitchen is physically separated from the bedroom area, and very poorly

ventilated (Figure 16). This type of household structure is very common in rural areas

of Senegal and other Sub-Saharan West African countries (Ochieng et al., 2013).

In Bibane, like most rural sub-Saharan villages, women are responsible for cooking, as

well as for the rest of the household chores. Cooking is among the most time-

consuming of women’s responsibilities (Wodon and Blackden, 2006). They spend much

of their time near the stove and performing other tasks related to food preparation,

such as pounding the millet by hand, fetching water, etc. They usually cook inside the

kitchen, to protect the food and fire from animals, wind and children, and to be

protected from the sun. Fuelwood is scarce in Bibane, so its collection requires long

journeys to the forest. As a result, fuel used to cook is heterogeneous, with a high

percentage of families frequently using crop residues and dung.

All of the families in the village used traditional three-stone fires before improved cook

stoves were introduced in 2012, within the framework of a program to optimize the

use of forest resources in the region. In total, 2,500 improved cookstoves (known as

Noflaye Jegg) were distributed in the region, 50 of them in Bibane. However, the

research team discovered that the majority of the improved cookstoves were in

disuse, completely or partially damaged, and only 15 out of 50 were in a good or fair

state, and considered adequate to be included in the study.

Before starting measurements, a communal meeting was conducted to explain the

study purpose and activities, where women provided oral consent to enrol in the

study. Finally, the 15 households with improved cookstoves were included, together

with other 14 using the traditional three stone fire.

SENEGAL FATICK

Figure 16. Typical kitchen in Bibane, rural village in the region of Fatick, Senegal. Source:own elaboration.

Page 55: INTO THE SMOKE: A RESEARCH ON HOUSEHOLD AIR ...oa.upm.es/49583/1/CANDELA_DE_LA_SOTA_SANDEZ.pdfCada año, la contaminación del aire interior provoca la muerte prematura de 4,3 millones

39

Emissions testing was done using a Portable Emission monitoring system (PEMS), with

the same design of the LEMS (section 3.2.1), but with a portable hood made from

fireproof fabric. The portable hood was placed before starting the meal preparation

directly over the centre of the combustion zone (see Figure 17) at 50-80 cm above the

cooking surface, to allow cookers to have an acceptable distance for the stove use.

Background concentrations were sampled and then subtracted from the emission

sample concentrations to account for background contributions to emissions samples.

Emission sampling was started once the pot was put over the cookstove and start time

was recorded, including the ignition phase emissions. At the end of the food

preparation stage, emissions sampling was terminated, cooked food weighed and end

time recorded using a digital scale of 0−30 kg range with 1 g resolution. Fuelwood was

measured before and after cooking sessions, as well as the charcoal formed. The goal

was to minimize interference with normal cooking practices, so the food prepared

during the tests was freely chosen by families

An important limitation of the PEMS sampling in remote areas where lack of electrical

power. We connected it to a diesel generator, placed far away enough to not disturb

women while cooking and to avoid emissions interferences.

4.2.3. Stoves and fuels

Four stoves were selected between the most commonly used in rural Senegal, each

one representing a method of fuelwood combustion: i) Three-stone (open burning

Figure 17. Portable Emission Monitoring System (PEMS), designed by the Aprovecho Research Centre. Source: own elaboration.

Page 56: INTO THE SMOKE: A RESEARCH ON HOUSEHOLD AIR ...oa.upm.es/49583/1/CANDELA_DE_LA_SOTA_SANDEZ.pdfCada año, la contaminación del aire interior provoca la muerte prematura de 4,3 millones

40

traditional stove); ii) Noflaye Jegg (rocket combustion chamber-type locally produced);

iii) Jambaar bois (basic improved ceramic stove, widely distributed in the Dakar region

by the German cooperation PERACOD program); iv) Prime Square Stove (a top-lit

updraft gasifier, TLUD, with a natural draft, with no fan). Photos of the stoves are

shown in Figure 18.

Each stove was analysed in the laboratory using two wood species very commonly

used in Senegal as residential fuel for cooking: Casuarina Equisetifolia (common named

Filao) and Cordyla Pinnata (common named Dimb) (Ndiaye et al., 2015).

Both were split in regular pieces of 15x5x3cm. Moisture content in wet basis for dimb

and filao was 8% and 8.8%, respectively, and calorific value was 21.79 MJ/Kg and 19.0

MJ/Kg, respectively.

Three stones and Noflaye Jegg were the stove types analysed in the field. Three stones

is still the most commonly used biomass stove in Senegal, and Noflaye Jegg stove has

been widely implemented in the Casamance Region, (25,000 units between 2012 and

2016, (Sota et al., 2014) and in Fatick (2500 units in 2012). Rocket stove design is the

most commonly used type of improved cookstove in sub-Saharan Africa (Lambe et al.,

2015b).

As fuel used in Bibane was quite heterogeneous due to fuelwood scarcity, the same

wood was distributed to every family participating in the study. The test performance

with uniform wood type allowed to reduce variability between households, and thus

the sample size needed to have statistically significant results, and to focus the study in

the effect of stove type on carbonaceous aerosol emissions. Wood used during field

testing (dimb) was purchased to a local vendor locally, split in bigger and more

irregular pieces than those used in the laboratory, and distributed to households.

Purchased dimb had a moisture content of 5% on a wet basis and 21.69 MJ/Kg of

calorific value.

Figure 18. Cookstoves tested in the study (from left to right: three stones, Noflaye Jegg, Jambaar Bois and gasifier Prime Square. Source: own elaboration.

Page 57: INTO THE SMOKE: A RESEARCH ON HOUSEHOLD AIR ...oa.upm.es/49583/1/CANDELA_DE_LA_SOTA_SANDEZ.pdfCada año, la contaminación del aire interior provoca la muerte prematura de 4,3 millones

41

4.2.4. EC and OC analysis

After sampling, particle-loaded filters were packed with aluminium foil and stored in a

freezer until analysis. Gravimetric measurements of glass fibre filters were conducted

using a high precision digital balance at the CERER, and OC/EC analyses on quartz

filters were conducted on a Sunset Laboratory Carbon Analyser using the EUSAAR2

Protocol (see section 3.2.2.) in the in the Institute of Environmental Assessment and

Water Research (IDÆA-CSIC) in Barcelona. The Sunset Laboratory carbon Analyser was

chosen between the methods presented in chapter 2 because it is the most accurate

method (see Table 1), and because it is the only one which can analyse OC content. For

this reason, results presented in this chapter are termed as EC.

4.2.5. Data Analysis

Kruskal-Wallis one-way analysis of variance was used to identify significant differences

of OC and EC EFs with different cookstoves, fuels, and during different WBT phase.

Wilcoxon t-test was conducted in order to identify pairs with statistically significant

difference in EF values. Non parametric tests were used because observations were

not normally distributed, due to the relative small sample size and the presence of

outliers (mainly in results from the field study).

The inter quartile range (IQR), difference between the 75th percentile and the 25th

percentile, was used to determine the test-to-test variability. This measure is robust

against outliers and non-normal data. All statistical analyse were conducted at a

significance level of 0.05 and performed using IBM SPSS Statistics Base software and

STATGRAPHICS (StatPoint Technologies, Inc., 2014).

4.3. Results and discussion

4.3.1. Laboratory estimation of EC and OC emission using the WBT

Figure 19 presents EC EFs (in grams per MJ, for comparison on the same scale), total

fuel consumption per WBT, and EC total emissions per WBT completed.

Results on Figure 19a show that the type of stove had an effect on EC EFs. The highest

average EC EF per WBT was found with the Noflaye Jegg for both fuels (dimb and filao),

while EC EFs were fairly uniform across the other two improved cookstoves and the

three stones.

Differences of EC EF’s between stoves are related to the nature of combustion (Arora

and Jain, 2015; MacCarty et al., 2007a), which is, in turn, affected by a number of

factors (e.g. air-fuel ratio, burn rate, combustion temperature, combustion efficiency,

thermal efficiency, residence time in the combustion chamber, flame turbulence, etc.)

(MacCarty et al., 2008b; Venkataraman et al., 2005).

Page 58: INTO THE SMOKE: A RESEARCH ON HOUSEHOLD AIR ...oa.upm.es/49583/1/CANDELA_DE_LA_SOTA_SANDEZ.pdfCada año, la contaminación del aire interior provoca la muerte prematura de 4,3 millones

42

Figure 19. Laboratory estimation of EC emission using the WBT. Figure 19a: Elemental carbon (EC) emission

factors per Water Boiling Test (WBT); Figure 19b: EC emission factors per WBT phase (cold-start and simmer);

Figure 19c: Fuel consumption per WBT; Figure 19d: Total EC Emissions per WBT. Source: own elaboration

In this study, the CO/CO2 ratio was used as a benchmark for stove combustion

efficiency (Guofeng et al., 2012; Li et al., 2009) (Table 4). The Noflaye Jegg performed

with the lowest CO/CO2 ratios (3.4±1.1% for dimb and 5.3±0.1% for filao), indicating

efficient combustion, and the highest EC emissions (0.18±0.06 g/MJ and 0.06±0.01

g/MJ) for dimb and filao, respectively.

CO/CO2 ratio (%)

Noflaye Jegg Three stones Jambaar bois Gasifier

Dimb 3.4±1.1% 8.7±0.5% 7.8±2.4% 3.6±0.6%

Filao 5.3±0.1% 8.3±1.8 % 6.7±1.9 % 4.0± 0.4%

Table 4. CO/CO2 ratio (expressed in %) for the stove-fuel combinations analysed in the laboratory. Source: own elaboration

Rocket stoves have a strong draft which enhances air flow through the fire, improving

the combustion efficiency and resulting in higher combustion temperatures (Li et al.,

2009; MacCarty et al., 2008b), which typically results in higher EC emissions (Bond,

2004; Cachier et al., 1996; MacCarty et al., 2008a; Rau, 1989).

Page 59: INTO THE SMOKE: A RESEARCH ON HOUSEHOLD AIR ...oa.upm.es/49583/1/CANDELA_DE_LA_SOTA_SANDEZ.pdfCada año, la contaminación del aire interior provoca la muerte prematura de 4,3 millones

43

The three stones showed the highest CO/CO2 ratios (8.7±0.5% with dimb and 8.3±1.8

% with filao), indicating inefficient combustion, and low values of EC EF (0.09±0.01

g/MJ for dimb and 0.04±0.01 g/MJ for filao).

Jambaar bois showed very similar values to the three stones fire, both in combustion

efficiency and EC EFs values. CO/CO2 ratio was 7.8±2.4% and 6.7±1.9 %; and EC EFs

were 0.09±0.01 g/MJ and 0.05±0.01 g/MJ, with dimb and filao, respectively. Previous

studies have also found EFs for basic ceramic stove similar to traditional stoves

(Wathore et al., 2017). Table 6 summarises EC and OC EFs determined in previous

laboratory and field studies.

The gasifier showed the lowest CO/CO2 ratio (i.e. high efficiency), 3.6±0.6% and 4.0±

0.4%, with dimb and filao, respectively; and low EC EF’s: 0.09±0.02 g/MJ when burning

dimb (equal to EF of jambaar bois and three stones), and 0.05±0.01 g/MJ with filao.

This is coherent with previous findings showing the lowest EF for gasifiers in

comparison with traditional and other basic improved stoves (Arora and Jain, 2015;

MacCarty et al., 2008b; Wathore et al., 2017), (Table 6).

In addition to the cookstove technology, the biofuel characteristics also determine

combustion conditions, and thus, aerosol emissions (Li et al., 2009; Venkataraman et

al., 2005; Venkataraman and Rao, 2001). In this study, EC EF’s of all the stoves tested

showed higher EC EF’s when burning dimb over filao (64.2%, 51.7%, 49.7% and 56.7%

for the Noflaye Jegg, Jambaar bois, gasifier and three-stones, respectively), although

not statistically significant at 95% confidence level.

A number of characteristics of fuel wood could affect pollutant emissions: i) the

geometrical shape and size of the fuel; ii) the density (Atiku et al., 2016; Mitchell et al.,

2016); iii) the lignin content (Atiku et al., 2016); (v) the heat value and (vi) the moisture

(Shen et al., 2010), among others.

Dimb and filao used in this study had similar moisture fraction (8% and 8.8%, on a wet

basis) and calorific value (21.79 MJ/Kg and 19.0 MJ/Kg), and both were split in stick of

15x3x4 cm. Typical lignin content is around 26% for casuarina esquetiofila (Filao)

(Mahmood, 1993), and 40% for cordyla pinnata (dimb) (Samba, 2001). Previous studies

show that lignin content promotes elemental carbon formation (Atiku et al., 2016;

Tillman, 1987), so this may partially explain the higher EC EF’s of wood dimb.

Results suggest that emission factors of EC are dependent on the type of fuel burned,

but further studies could study the effect of physical and chemical characteristics of

filao and dimb in EC EF’s, which cannot be explained at this stage.

Figure 19b presents EC EFs for the stove-fuel combinations by WBT phase, except in

the case of the gasifier, where only the EF of the whole WBT is presented, because its

cylindrical combustion chamber cannot be emptied during the burning cycle. EC EFs

Page 60: INTO THE SMOKE: A RESEARCH ON HOUSEHOLD AIR ...oa.upm.es/49583/1/CANDELA_DE_LA_SOTA_SANDEZ.pdfCada año, la contaminación del aire interior provoca la muerte prematura de 4,3 millones

44

were found to be higher during the cold start phase for every stove-fuel combination

analysed. This can be explained because cold start phase is characterized by strong

flames and significant amounts of elemental carbon (Atiku et al., 2016; Roden et al.,

2009), especially shortly after fire ignition (Roden et al., 2006), whereas the simmer

phase is characterized by smaller flames, with lower emission of EC particles (Roden et

al., 2006).

EC EFs were 20%, 10%, 29% higher for the cold start phase with regard to simmer, for

the dimb with three stone, jambaar bois and noflaye jegg, respectively; and 25%, 20%

and 14% higher for filao with three stone, Jambaar Bois and Noflaye Jegg, though not

statistically significant (p>0.05). Previous studies also found that changes in

combustion conditions during the cooking cycle induced variation in EC EFs (Arora and

Jain, 2015; Jetter et al., 2012; Roden et al., 2006).

In terms of total EC emission per WBT completed, Figure 19d shows that results

followed almost the same trend as in the case of average EC EFs per WBT. Noflaye Jegg

showed the highest EC EF and did not show a reduction in fuel wood consumption per

task completed with respect to the three stones fire (Figure 19c). Therefore, this type

of stove has the highest EC emission per WBT (3.75 g/WBT with dimb and 1.90 g/WBT

with filao). On the other side, the gasifier showed the highest fuel savings and low EC

EFs, so this type of stove induced to the smallest total EC emissions per WBT (1.35

g/WBT with dimb and 0.57 g/WBTwith filao). An intermediate situation was found for

the three stones and the Jambaar bois, which show very similar results in terms of EC

total emissions (three stones, 2.00 g/WBT with dimb and 0.67 g/WBT with filao and

Jambaar bois (1.94 g/WBT with dimb and 0.89 g/WBT with filao).

Figure 20 presents OC EFs for three stones+filao, jambaar bois+filao, gasifier+filao and

gasifier+dimb. OC EFs for the rest of stove-fuel combinations are not presented due to

problems in filter conservations between WBT and EC/OC analysis.

When burning wood filao, the average OC EF for the whole WBT was found to be the

lowest for the gasifier (0.08±0.01 g OC/MJ), followed by three stones (0.18±0.03 g

OC/MJ) and then Jambaar bois (0.21±0.08 g OC/MJ). Low OC and EC emissions of

gasifier are explained because this technology emits very low levels of particulate

matter when compared to other biomass stoves (MacCarty et al., 2008b). As already

observed in previous studies, in the case of the three-stone fire and jambaar bois, a

large bed of charcoal under the flaming fuel resulted in smoldering conditions, with

high OC particles emissions (MacCarty et al., 2007a). Overall, the EF OC results appear

to be consistent with prior studies (Table 6).

Page 61: INTO THE SMOKE: A RESEARCH ON HOUSEHOLD AIR ...oa.upm.es/49583/1/CANDELA_DE_LA_SOTA_SANDEZ.pdfCada año, la contaminación del aire interior provoca la muerte prematura de 4,3 millones

45

Figure 20. Laboratory estimation of OC emission using the WBT Figure 20a: OC emission factors per WBT; Figure

20b: OC emission factors per WBT phase (cold-start and simmer); Figure 20c: Total OC emissions per WBT; Figure

20d: OC/EC ratio per WBT. Source: own elaboration.

Phases that were associated with higher OC emissions were different depending on

the cookstove tested, as occurred in previous studies (Arora and Jain, 2015). For the

traditional stove, OC EF was found to be higher during simmer (0.19±0.05 g OC/MJ)

than during cold start (0.16±0.03 g OC/MJ), whereas Jambaar bois showed highest OC

EF during cold start (0.22±0.09 g OC/MJ) when compared to simmer phase (0.19±0.05

g OC/MJ), although none of these differences between WBT phases were statistically

significant (Wilcoxon, p<0.05).

The higher OC emissions during simmer phase can be attributed to high-carbon

content and large surface area per unit mass of charcoal, that results in the formation

of organic compounds (Akagi et al., 2010). On the other hand, higher OC emissions

during flaming conditions (cold start phase) with the Jambaar Bois, could be attributed

to low flame temperatures due to heat losses through the cookstove walls (Arora and

Jain, 2015).

Regarding the influence of fuelwood species on OC EFs, the gasifier showed higher OC

EFs when burning dimb (0.46±0.05 g OC/MJ) than with filao (0.08±0.01 g OC/MJ).

Higher OC emissions with dimb might be related with the ash content of this fuelwood

specie (Guofeng et al., 2012), but further chemical composition analysis should be

d)

Page 62: INTO THE SMOKE: A RESEARCH ON HOUSEHOLD AIR ...oa.upm.es/49583/1/CANDELA_DE_LA_SOTA_SANDEZ.pdfCada año, la contaminación del aire interior provoca la muerte prematura de 4,3 millones

46

done to verify this assumption Therefore, gasifier results would suggest that OC EF’s

are dependent on the type of fuel burned, although emission tests with more types of

stove should be performed.

Figure 20d presents results of OC/EC ratio, considered an environmentally relevant

property with good predictive capabilities, even for fuels in which brown carbon

absorption is significant (Pokhrel et al., 2015). Overall, all stove fuel combinations

showed OC/EC ratios higher than one, indicating the dominance of OC and suggesting

that the smoldering combustion was dominant during the WBT.

Figures 19d and 20c show that OC and EC total emissions had a relatively large degree

of test-to-test variability, and boxes overlapped considerably. For this reason,

differences in OC and EC EFs across stove types, phase of WBT and type of fuel were

not found to be statistically significant in most cases (p>0.05).

In terms of EC total emissions, the highest variability was observed for Noflaye Jegg

(IQR=2.8 with dimb and IQR= 1.03 with filao). This is because flaming conditions are

more present in the cooking cycle for this stove, characterized by more frequent peaks

of emissions, which significantly affect total EC. For OC total emissions, Jambaar bois

showed the highest IQR, equal to 5.8.

In this study three replicates of WBT were conducted because it was commonly used in

the literature (Wang et al., 2014) and because of time and resource limitations.

However, number of replicates is linked to the stove-fuel combination being tested

and sometimes three are not enough to ensure confidence in the results (Wang et al.,

2014). These findings suggest that in forthcoming emissions testing of the stoves

included in this study, a larger number of WBT replicates, especially for the higher

emitters such as the rocket stove, should be analysed.

Table 5 summarizes EC and OC emission factors, EC/OC ratios and total emissions per

WBT from the stove-fuel combinations tested under laboratory.

Page 63: INTO THE SMOKE: A RESEARCH ON HOUSEHOLD AIR ...oa.upm.es/49583/1/CANDELA_DE_LA_SOTA_SANDEZ.pdfCada año, la contaminación del aire interior provoca la muerte prematura de 4,3 millones

47

Stove type Fuel

type

EC

(g/MJ)

EC

(g/kg)

EC

(g/L of

water)

OC

(g/MJ)

OC

(g/kg)

OC

(g/L of water) OC/EC

Fuel consumed

per WBT

Total BC emited

per WBT

Total OC

emitted per

WBT

Three

stones

Dimb 0.09±0.01 1.98±0.17 0.20±0.01 No data No data No data No data 1.00±0.02 2.00±0.20 No data

Filao 0.04±0.01 0.75±0.20 0.07±0.02 0.18±0.03 3.33± 0.61 0.29±0.05 4.67±0.85 0.89±0.02 0.67±0.16 2.92±0.46

Jambaar

bois

Dimb 0.09±0.01 2.05±0.16 0.19±0.02 No data No data No data No data 0.94±0.10 1.94±0.34 No data

Filao 0.05±0.01 0.87±0.23 0.08±0.02 0.21±0.08 4.04± 1.50 0.42±0.28 4.61±0.40 0.98±0.29 0.89±0.49 4.22±2.80

Noflaye

Jegg

Dimb 0.18±0.06 3.93±1.20 0.36±0.11 No data No data No data No data 0.96±0.06 3.75±1.37 No data

Filao 0.06±0.01 1.23±0.20 0.19±0.02 No data No data No data No data 1.63±0.12 1.90±0.31 No data

Gasifier

Dimb 0.09±0.02 1.95±0.52 0.27±0.07 0.46±0.05 9.91± 1.11 1.38±0.15 5.40±1.71 0.70±0.02 1.35±0.35 4.55±3.80

Filao 0.05±0.01 0.86±0.18 0.11±0.02 0.08±0.01 1.49± 0.28 0.20±0.04 1.76±0.24 0.67±0.01 0.57±0.12 0.63±0.48

Table 5. Summary of EC and OC emission factors, EC/OC ratios and total emissions per WBT from the stove-fuel combinations tested in the laboratory

Page 64: INTO THE SMOKE: A RESEARCH ON HOUSEHOLD AIR ...oa.upm.es/49583/1/CANDELA_DE_LA_SOTA_SANDEZ.pdfCada año, la contaminación del aire interior provoca la muerte prematura de 4,3 millones

48

4.3.2. Field estimation of EC and OC emissions

Emission tests were conducted in 29 households in the rural village of Bibane: 15 with

the Noflaye Jegg and 14 with the three-stone fire. Most families cooked inside the

kitchen space during the tests (separated from the rest of the house and very poorly

ventilated), although few of them cooked outdoors.

All families prepared typical meals of rural households in the region. They consisted of

rice with vegetables, fish, or meat for lunch, whereas for dinner water was boiled to

steam cous-cous and a sauce was prepared. Time of preparation ranged from 38 to

111 minutes (average=75 minutes), with no significant differences between lunch and

dinner and between the type of stove used. Likewise, no differences were found in the

quantity of fuel used per kg of food prepared between the three stones and the

Noflaye Jegg (0.6 ±0.4 kg of fuel/kg food), consistent with previous laboratory results.

As fuelwood type was the same for every household, field results focused on the

analysis of the influence of stove type on EC, OC and PM emissions. PM EFs (Figure

21a) and OC EFs (Figure 21c) were 30.3% and 35.7% lower for Noflaye Jegg when

compared to the three stones, although this difference was not statistically significant

(p>0.05). On the other hand, EC EFs (Figure 21b), increased a 75% (p<0.05) with

Noflaye Jegg in comparison with the three stones stove. The ratio OC/EC (Figure 21d)

was 1.77±0.56 for the three stones, indicating the dominance of OC, which is caused

by the smoldering combustion (Arora et al., 2013). OC/EC ratio < 1 (0.67±0.28) for the

Noflaye Jegg is showing the dominance of flaming conditions, typical in the rocket

stoves (MacCarty et al., 2008a)

Page 65: INTO THE SMOKE: A RESEARCH ON HOUSEHOLD AIR ...oa.upm.es/49583/1/CANDELA_DE_LA_SOTA_SANDEZ.pdfCada año, la contaminación del aire interior provoca la muerte prematura de 4,3 millones

49

Therefore, while the Noflaye Jegg reduces PM EF, the lower OC/EC ratio implies that

carbonaceous aerosol emissions with this stove are more warming (from a climate

perspective) when compared to the three stones, which could lessen some of the

impact of reducing overall aerosol emissions (Johnson et al., 2008b). However, the

comparison of both stoves should be done in terms of total EC and OC emissions,

taking into account their fuel-efficiency (section 4.3.4). Overall, EC and OC EFs

obtained agree well with previous studies (Table 6), although many differences were

observed due to differences on stove type, fuel parameters, operation conditions, etc.

Nonetheless, most publications present EF in g/kg instead of g/MJ, which prevents

comparison between different fuel types at same level. Table 7 presents a summary of

EF’s, ratios and total emissions from in-field emission testing for each household

included in the study.

a) b)

c) d)

Figure 21. Field estimation of EC and OC emissions. Figure 21a: Particulate matter emission factors; Figure 21b: Elemental carbon emission factors; Figure 21c: Organic carbon) Emission factors; Figure 21d:OC/EC ratio. Source: own elaboration.

Page 66: INTO THE SMOKE: A RESEARCH ON HOUSEHOLD AIR ...oa.upm.es/49583/1/CANDELA_DE_LA_SOTA_SANDEZ.pdfCada año, la contaminación del aire interior provoca la muerte prematura de 4,3 millones

50

Type of Study Country Type of stove and fuel EC EF

(g/Kg)

OC EF

(g/Kg) Study

Field UCT and

laboratory WBT Senegal

Fuel: wood (Cordyla Pinnata and Casuarina Equisetifolia )

Stoves:

-Three stones (lab WBT)

-Rocket stove (lab WBT)

-Ceramic improved stove (lab WBT)

-Gasifier (lab WBT)

-Three stones (in-field UCT)

-Rocket stove (in-field UCT)

(a1.98±0.17;

b0.75±0.20)

(a3.93±1.20;

b 1.23±0.20)

(a2.05±0.16;

b 0.87±0.23)

(a1.95±0.52;

b 0.86±0.18)

(a1.81±1.22)

(a3.14±1.51)

(b3.33±0.61)

(b4.04±1.50)

(a9.91±1.11;

b1.49±0.28)

(a3.02±2.27)

(a2.06±1.22)

This study

Field UCT Malawi

Fuel: wood

Stove:

-Traditional

-Ceramic improved stove

-Forced-draft Philips

-Forced-draft ACE

2.07±1.24

1.51±0.63

0.81±0.26

1.30±0.69

2.09±0.75

2.43±1.62

0.96±0.12

1.27±0.96

(Wathore et al.,

2017)

Laboratory

WBT USA

Fuel: wood (Douglas fir)

Stoves:

-Three stones

-Rocket stove

-Gasifier

-Fan stove

0.88

1.16

0.28

0.06

1.45

0.55

0.82

0.14

(MacCarty et al.,

2008b)

Field Honduras Fuel: wood (oak and pine)

Stove: Traditional woodstove

1.50±0.3

4.00±0.9 (Roden et al., 2006)

Page 67: INTO THE SMOKE: A RESEARCH ON HOUSEHOLD AIR ...oa.upm.es/49583/1/CANDELA_DE_LA_SOTA_SANDEZ.pdfCada año, la contaminación del aire interior provoca la muerte prematura de 4,3 millones

51

Simulated kitchen

WBT and in-field

WBT and UCT

Mexico

Fuel: wood

Stove:

-Open fire simulated kitchen WBT

-Mud–cement Patsari simulated-kitchen WBT

-Open fire in-home WBT

-Mud–cement Patsari in-home WBT

-Brick Patsari in-home WBT

-Open fire in-home

-Mud–cement Patsari in-home

-Brick Patsari in-home

1.1±0.5

1.0±0. 5

1.1±0.1

1.0±0.6

0.8±1.1

0.3±0.1

0.8±0.4

0.1±0.2

2.5±0.4

2.4±0.9

1.8±1.1

2.6±1.2

1.1±1.3

4.4±1.6

2.7±1.1

0.8±0.9

(Johnson et al.,

2008b)

Field China Fuel: wood

Stove: improved metal stove without chimney

0.91-1.60

2.20-3.60 (Shen et al., 2013)

Field China

Fuel: wood (17 tree species widely used as biofuel in

China)

Stove: improved chimney brick stove

0.83±0.69

0.62±0.64

(Guofeng et al.,

2012)

Laboratory WBT India

Fuel: wood (4 species widely used as biofuel in India)

Stove: Traditional stove

0.38-0.62

0.17-4.69 (Venkataraman et

al., 2005)

Laboratory

Controlled Cooking

Test

India

Fuel: Acacia Nilotica

Stoves:

-Traditional

-Natural-draft front feed

-Natural-draft top feed

-Forced-draft

d1.47±0.15

d1.26±0.11

d 1.89±0.06

d 0.42±0.06

d 1.47±0.21

d 1.05±0.13

d 0.84±0.04

d 0.84±0.13

(Arora and Jain,

2015)

Table 6. Comparison of Measured EC and OC EF for fuelwood combustion with previous studies. Source: own elaboration.

Page 68: INTO THE SMOKE: A RESEARCH ON HOUSEHOLD AIR ...oa.upm.es/49583/1/CANDELA_DE_LA_SOTA_SANDEZ.pdfCada año, la contaminación del aire interior provoca la muerte prematura de 4,3 millones

52

HH

CODE

STOVE

TYPE

TYPE OF

MEAL

TYPE OF

WOOD

DURATION

COOKING

SESSION

(min)

TOTAL

FUEL USED

(kg)

TOTAL FOOD

PREPARED(k

g)

fuel

used/kg of

food

PM

(g/MJ)

PM

(g/Kg)

EC

(g/MJ) EC (g/Kg)

OC

(g/MJ)

OC

(g/Kg) OC/EC EC/PM OC/PM

1NJ Noflaye

Jegg Rice Dimb 83,7 2,9 5,9 0,5 0,18 3,97 0,10 2,18 0,05 1,04 0,48 0,55 0,26

2NJ Noflaye

Jegg Rice Dimb 58,8 2,2 3,1 0,7 0,22 4,72 0,08 1,80 0,08 1,84 1,02 0,38 0,39

3NJ Noflaye

Jegg Rice Dimb 110,9 2,1 8,5 0,2 0,19 4,15 0,12 2,61 0,07 1,43 0,55 0,63 0,34

4NJ Noflaye

Jegg Rice Dimb 67,1 2,3 5,7 0,4 0,17 3,60 0,08 1,69 0,04 0,91 0,54 0,47 0,25

5NJ Noflaye

Jegg Sauce Dimb 104,7 2,1 8,2 0,3 0,23 5,09 0,13 2,78 0,06 1,37 0,49 0,55 0,27

6NJ Noflaye

Jegg Sauce Dimb 64,2 1,7 2,2 0,8 0,20 4,37 0,12 2,60 0,06 1,38 0,53 0,59 0,32

7NJ Noflaye

Jegg Rice Dimb 82,7 1,8 5,5 0,3 0,19 4,05 0,07 1,57 0,05 1,01 0,64 0,39 0,25

8NJ Noflaye

Jegg Rice Dimb 38,0 1,3 0,8 1,6 0,16 3,55 0,08 1,76 0,12 2,67 1,52 0,50 0,75

9NJ Noflaye

Jegg Rice Dimb 92,3 1,7 6,0 0,3 0,61 13,12 0,28 6,14 0,19 4,13 0,67 0,47 0,31

10NJ Noflaye

Jegg

Cous-

cous Dimb 157,2 4,5 6,2 0,7 0,11 2,49 0,19 4,04 0,09 1,88 0,46 1,62 0,75

11NJ Noflaye

Jegg Sauce Dimb 80,4 2,4 11,2 0,2 0,30 6,43 0,16 3,45 0,08 1,77 0,51 0,54 0,27

12NJ Noflaye

Jegg Rice Dimb 80,2 3,1 4,6 0,7 0,34 7,27 0,27 5,90 0,23 5,02 0,85 0,81 0,69

13NJ Noflaye

Jegg Sauce Dimb 82,6 2,2 4,0 0,6 0,16 3,38 0,23 5,03 0,15 3,31 0,66 1,49 0,98

14NJ Noflaye

Jegg Sauce Dimb 81,4 2,7 2,5 1,1 0,14 2,99 0,10 2,19 0,05 1,05 0,48 0,73 0,35

Page 69: INTO THE SMOKE: A RESEARCH ON HOUSEHOLD AIR ...oa.upm.es/49583/1/CANDELA_DE_LA_SOTA_SANDEZ.pdfCada año, la contaminación del aire interior provoca la muerte prematura de 4,3 millones

53

15NJ Noflaye

Jegg Rice Dimb 77,4 2,4 10,7 0,2 0,33 7,10 0,16 3,38 0,09 2,03 0,60 0,48 0,29

1T Three

stones Rice Dimb 56,6 5,3 19,0 0,3 0,08 1,72 0,01 0,30 0,04 0,83 2,81 0,17 0,49

2T Three

stones Sauce Dimb 79,7 4,6 5,9 0,8 0,24 5,21 0,07 1,44 0,09 1,94 1,35 0,28 0,37

3T Three

stones Sauce Dimb 54,1 4,3 10,2 0,4 0,19 4,08 0,05 1,05 0,07 1,57 1,49 0,26 0,38

4T Three

stones Rice Dimb 54,1 2,4 3,3 0,7 0,25 5,46 0,05 1,05 0,09 1,95 1,85 0,19 0,36

5T Three

stones Rice Dimb 48,2 1,0 2,2 0,4 0,28 6,05 0,09 1,90 0,11 2,41 1,27 0,31 0,40

6T Three

stones Rice Dimb 74,2 2,3 2,2 1,0 0,99 21,55 0,21 4,59 0,46 10,05 2,19 0,21 0,47

7T Three

stones Sauce Dimb 71,8 1,6 3,8 0,4 0,38 8,20 0,07 1,62 0,11 2,44 1,50 0,20 0,30

8T Three

stones Sauce Dimb 96,5 3,6 4,5 0,8 0,20 4,32 0,19 4,19 0,17 3,60 0,86 0,97 0,83

9T Three

stones Rice Dimb 48,1 1,4 4,8 0,3 0,26 5,65 0,08 1,72 0,16 3,52 2,05 0,30 0,62

10T Three

stones Sauce Dimb 95,6 2,2 4,1 0,5 0,30 6,55 0,08 1,64 0,14 2,96 1,81 0,25 0,45

11T Three

stones Rice Dimb 64,3 2,0 3,6 0,5 0,40 8,66 0,04 0,96 0,13 2,73 2,83 0,11 0,31

12T Three

stones Rice Dimb 65,8 1,6 8,7 0,2 0,28 6,09 0,08 1,80 0,13 2,77 1,53 0,30 0,45

13T Three

stones Sauce Dimb 101,1 4,1 2,1 1,9 0,46 9,96 0,11 2,38 0,21 4,58 1,92 0,24 0,46

14T Three

stones

Cous-

cous Dimb 156,2 5,5 12,2

0,11 2,47 0,03 0,72 0,04 0,94 1,31 0,29 0,38

Table 7. Summary of EC, OC and PM Emission Factors, ratios and total emissions from in-field emission testing for each household included in the study. Source:own elaboration

Page 70: INTO THE SMOKE: A RESEARCH ON HOUSEHOLD AIR ...oa.upm.es/49583/1/CANDELA_DE_LA_SOTA_SANDEZ.pdfCada año, la contaminación del aire interior provoca la muerte prematura de 4,3 millones

54

4.3.3. Laboratory vs. field results

EC EFs from WBT were compared with results obtained under real world conditions for

the three stones and the Noflaye Jegg stoves when burning dimb (Figure 22. Boxplots

of EC EF (g/MJ) for each type of stove and test. Source: own elaboration.).

EC EFs estimated at laboratory scale with the WBT were 18.0% higher than EF

determined during daily cooking activities for the Noflaye Jegg, and 14.5% higher in the

case of the three stones (not significant, p-value>0.05). Previous studies also found

differences between results from laboratory tests and those obtained during

uncontrolled field conditions (Arora and Jain, 2015; Chen et al., 2012; Johnson et al.,

2008a; Roden et al., 2009, 2006; Shen et al., 2013), although differences found in these

studies were higher than in this study. Further studies with a higher number of

laboratory and field repetitions would be necessary to obtain a more representative

comparison.

Differences between laboratory and field results can be explained by several factors.

Firstly, the fire management behaviour in the laboratory was essentially normalized;

cooking fire was constantly tended, and fuel was added more frequently, whereas the

process observed in the field was rather random, including some low combustion

efficiency situations. On the other hand, wood pieces used in the laboratory had

regular shape and dimensions, while in the field they were less uniform in shape and

with larger dimensions. Smaller wood has been found to yield lower PM emissions and

a higher BC fraction emissions (Bond, 2004) (Li et al., 2009). Finally, minor fugitive

emissions may have escaped through the fabric hood during field testing.

Differences were lower when comparing EF from field tests with those obtained during

the simmer phase of WBT: 3.8% for the three stones and 4.2% for the Noflaye Jegg

(p>0.05) (Table 8).

Figure 22. Boxplots of EC EF (g/MJ) for each type of stove and test. Source: own elaboration.

Page 71: INTO THE SMOKE: A RESEARCH ON HOUSEHOLD AIR ...oa.upm.es/49583/1/CANDELA_DE_LA_SOTA_SANDEZ.pdfCada año, la contaminación del aire interior provoca la muerte prematura de 4,3 millones

55

EC EF (g/MJ)

% difference*

Stove-fuel

WBT

Field test

Cold start Simmer Average WBT Cold start Simmer Average

WBT

Three stone-dimb 0.10±0.01 0.08±0.01 0.09±0.01 0.08±0.06 20% 3.8% 14.5%

Noflaye Jegg-dimb 0.21±0.07 0.15±0.04 0.18±0.06 0.14±0.07 33.3% 4.2% 18%

Table 8. EC EFs and % difference between laboratory and field data sets.*differences were not significant (p>0.05). Source:own elaboration

This is coherent with field observations during testing activities, which showed that

during meal preparations (rice and cous-cous), cookers maintained a constant and low-

medium flame during almost all the time, except when frying vegetables, fish or meat

for the first few minutes, which needed stronger flames.

Considering that, although user practices are highly individual, they may have regional

averages governed by common customs (Chen et al., 2012), simmer phase of WBT is

likely quite illustrative of EC emissions in homes of rural villages of Senegal. Further

field measurements are needed as a reality check for each stove-testing and region.

4.3.4. Estimation of total emissions and net climate effect of cooking-related

carbonaceous aerosol in Senegal

Total annual EC and OC emission per year of use of the Noflaye Jegg and the three

stones were calculated with EF values and data of daily consumption (9.9±5.1 kg/day

for the Noflaye Jegg and 9.6±5.5 Kg/day and for the three stones), both determined

during the field study. As in this region cooking activities occur most often in indoor

environments, it was necessary to take into account the fraction of emissions released

to ambient air (exfiltration rate), which is dependent on the ventilation and particle

deposition rates in kitchens (Soneja et al., 2015). Exfiltration rate for typical household

characteristics of this study was not available, so total emissions have been

determined using two values of exfiltration rate from previous studies (26% and 80%)

(Soneja et al., 2015 and Venkataraman et al., 2005, respectively). Annual EC and OC

emissions per stove were then factored by the Global Warming Potential (GWP)

recommended by the IPCC, 900 for EC (Bond et al., 2013) and -46 for OC (Bond et al.,

2011), for a time horizon of 100 years. Figure 23a presents the net climatic effect

estimated with regard to EC and OC emissions for the three stones and Noflaye Jegg

(in tons of CO2-eq per year).

Page 72: INTO THE SMOKE: A RESEARCH ON HOUSEHOLD AIR ...oa.upm.es/49583/1/CANDELA_DE_LA_SOTA_SANDEZ.pdfCada año, la contaminación del aire interior provoca la muerte prematura de 4,3 millones

56

Across a 100-year horizon, Noflaye Jegg showed an overall increase of 88% global

warming potential in comparison to the traditional stove. In consequence, while the

use of Noflaye Jegg has clear benefits for improving indoor air quality (Sota et al., 2014

and chapter 5 of this thesis), results suggest that this type of rocket stove, under

typical cooking practices and fuelwood used Senegal, leads to negative climate effects

with regard to carbonaceous aerosol emissions.

The fact that rocket and other types of natural draft stoves increase EC EFs is not new

(Arora and Jain, 2015; MacCarty et al., 2008a). The unexpected point is that Noflaye

Jegg stove did not provide fuel saving benefits, neither in the laboratory study, nor in

the field. For this reason, this work emphasizes the importance of taking into account

total EC and OC emissions in addition to EFs.

Besides this, total annual emissions of EC and OC from residential burning of fuel wood

in Senegal were estimated using the average value of EC and OC EFs determined in this

study (1.85 g EC/kg and 3.97 g OC/kg) and the total annual firewood used in

Senegalese households (450 ktoe/year, for the period 2000-2005) (Dafrallah, 2009).

As previously discussed, EC and OC EFs vary significantly across different types of

stove-fuel combinations. However, average EFs from this study could be considered as

representative, as they were estimated from a set of biomass cookstoves and wood

fuels widely used in Senegal. Fuel consumption is also highly dependent on type of

stove used, but data on the fractions of different stoves used in Senegal are

unavailable, so the total annually firewood used in Senegalese households was the

best information available.

Total emissions of EC and OC from residential fuelwood combustion in Senegal were

estimated to be 421.1±213.0 and 561.4±411.4 t/year, respectively, with 26% of

exfiltration rate, and 1295.6±655.3 t/year and 1727.5±1265.8 t/year with 80% of

Figure 23. Figure 23a: 100 years Global Warming Potential (from Elemental carbon and Organic Carbon emitted) per year of use of Noflaye Jeg stove and Three stones; Figure23b: 100 years Global Warming Potential (from Elemental

carbon and Organic Carbon emitted) per year from household fuelwood cookstoves in Senegal. Results are presented for a rate exfiltration of 26% and 80%. Source: own elaboration.

Page 73: INTO THE SMOKE: A RESEARCH ON HOUSEHOLD AIR ...oa.upm.es/49583/1/CANDELA_DE_LA_SOTA_SANDEZ.pdfCada año, la contaminación del aire interior provoca la muerte prematura de 4,3 millones

57

exfiltration rate (Table 9). Factoring by GWP100 values, net effect of EC and OC

emissions from household fuelwood consumption in Senegal were 353.1 kton of CO2-eq

with 26% of exfiltration rate and 1086.6 ton of CO2-eq with 80 % of exfiltration (Figure

23b).

Using the 5th percentile of the EC and OC EFs data set and 26% of exfiltration rate, the

result was 173.4 ton of CO2-eq, and, when using the 95th percentile of EC and OC EFs

and 80% of exfiltration rate, calculations show a result of 2031.9 ton of CO2-eq. This

range (173.4-2031.9 ton of CO2-eq) constitutes a rough estimation of the net effect on

climate of carbonaceous aerosols emitted into the atmosphere from household use of

fuelwood for cooking in Senegal.

It is important to highlight that a complete assessment of the global warming potential

should include other species co-emitted from residential cookstove use, such as SO2,

NOx ,CO, CO2, CH4, N2O (Bhattacharya et al., 2002; Roden et al., 2006; Smith et al.,

2000). However, the objective at this stage was to assess the net climate effect of

cooking-related carbonaceous aerosol emissions in Senegal, which is still uncertain.

Total EC and OC emissions from fuelwood consumption in Sub-Saharan and West

Africa were also calculated with the same EF and exfiltration rates used for Senegal,

and using data on personal consumption of energy for cooking in Sub-Saharan and

West Africa (IEA, 2006; OECD/IEA, 2014). Results presented in Table 9, together with

results of EC and OC emissions from other regions of the world where residential

biofuel combustion is prevalent, show that carbonaceous emissions from cooking in

Sub-Sarahan Africa are relevant and should be taken into account within national and

regional climate mitigation strategies.

Base year Fuel Type Region EC emissions (Gg/year) OC emission (Gg/year) Study

2000-05 Fuelwood Senegal 26%

a:

b 0.4 (0.2-0.8)

80%a:b1.3( 0.6-2.4)

26%: 0.6 (0.2-1.1)

80%: 1.7(0.6-3.3) This study

2005 Biomassc sub-Saharan

Africa

26%: 190.6(92.0-357.1)

80%: 586.5(283.1-1098.8)

26%: 254.1(94.2-488.6)

80%: 782.0(289.9-1503.3) This study

2012 Fuelwoodd West Africa

26%: 118.8(51.3-229.9)

80%: 365.6(157.8-707.4)

26%: 254.9(104.0-542.7)

80%: 784.5(320.1-1669.7) This study

2007 Fuelwood Rural China 92 75.7 (Guofeng et al.,

2012)

2000 Fuelwood China 109 547.8 (Cao et al., 2006)

2000-01 Fuelwood India e165(50-530) No data (Habib et al., 2004)

2000

Fossil fuels, biofuels,

open biomass burning

and urban waste

Global 7500 (2000-29,000) 47,000 (18,000-180,000) (Bond et al., 2013)

Table 9. BC and OC emissions in different regions of the world determined in this study and in previous publications. Source: own

elaboration. a % of total emissions released to the atmosphere (exfiltration rate);

b Average, 5

th percentile and 95

th percentile (in parentheses;

c

fuelwood, charcoal, agricultural waste and animal dung, dIncludes fuelwood directly consumed by households and fuelwood to produce

charcoal; e

Central value and uncertainty range at 95%.

Page 74: INTO THE SMOKE: A RESEARCH ON HOUSEHOLD AIR ...oa.upm.es/49583/1/CANDELA_DE_LA_SOTA_SANDEZ.pdfCada año, la contaminación del aire interior provoca la muerte prematura de 4,3 millones

58

5. Indoor air pollution from biomass cookstoves in Senegal

Page 75: INTO THE SMOKE: A RESEARCH ON HOUSEHOLD AIR ...oa.upm.es/49583/1/CANDELA_DE_LA_SOTA_SANDEZ.pdfCada año, la contaminación del aire interior provoca la muerte prematura de 4,3 millones

59

5.1. Background

In Senegal, IAP accounts for 4.8% of the national burden of disease (WHO, 2007), and

cause the prematurely death of 6,300 people every year (WHO, 2009). Although an

important increase of LPG use at household level in urban areas occurred in 2009, LPG

subsidies were eliminated and consumers returned in large numbers to wood-based

biomass for cooking (Sander et al., 2011b). Nowadays, biomass accounts for 83% of

the demand for domestic fuel use in rural areas and 58% in urban areas (WB, 2013).

The international donor community and national governments increased efforts since

the 1980’s to disseminate cleaner and more efficient burning stove designs in Senegal,

as in many LMI countries. A number of studies have analysed impacts on household air

pollution and personal exposure from cooking activities in Asia (Balakrishnan et al.,

2015; Bartington et al., 2017; Chengappa et al., 2007; Chowdhury et al., 2013;

Dasgupta et al., 2006; Dutta et al., 2007; Gao et al., 2009; Hu et al., 2014; Kar et al.,

2012; Siddiqui et al., 2009), Latin America (Albalak et al., 2001; Armendáriz et al.,

2008; Commodore et al., 2013; Fitzgerald et al., 2012; Naeher et al., 2000; Northcross

et al., 2010; Park and Lee, 2003; Riojas-Rodriguez et al., 2011; Smith et al., 2010; Zuk et

al., 2007) and Africa (Ezzati et al., 2000; Ochieng et al., 2017, 2013; Pennise et al.,

2009). However, almost no studies exist evaluating indoor air quality impacts from

biomass combustion with traditional stoves and indoor air quality improvements

derived from the use of improved cookstoves in Senegal.

This study provides a comprehensive real-world assessment of household pollutant

concentrations from the combustion of biomass fuels for cooking in Senegal. It focuses

on the quantification and characterization of BC, PM, UFP and CO indoor

concentrations from burning wood fuel during cooking activities using traditional and

improved cookstoves in a Senegalese rural village, which may be considered

representative of rural areas in the country.

UFP refers to the particles with an aerodynamic diameter less than 0.1 µm (Pope and

Dockery, 2006), characterized by low mass and large surface area. Their concentrations

are not accurately reflected by particle mass concentrations, so they are monitored in

terms of particle number concentrations (Sahu et al., 2011). In addition, the lung-

deposited surface area (LDSA) concentration is increasingly emphasized as relevant

parameter to estimate health effects of UFPs in urban areas (Ntziachristos et al., 2007;

Reche et al., 2015). Recent studies show that most particles generated from biomass

combustion are typically between 0.05 µm and 0.2 µm (Tiwari et al., 2014). However,

to date, research on this parameter with regard to cookstove emissions is still

relatively scarce (Hosgood et al., 2012; Leavey et al., 2015; Patel et al., 2016; Sahu et

al., 2011), especially when compared to studies focused on CO or PM mass.

Page 76: INTO THE SMOKE: A RESEARCH ON HOUSEHOLD AIR ...oa.upm.es/49583/1/CANDELA_DE_LA_SOTA_SANDEZ.pdfCada año, la contaminación del aire interior provoca la muerte prematura de 4,3 millones

60

Moreover, the study identifies parameters such as type of wood and household

characteristics (i.e. construction materials, ventilation and kitchen size), which

influence indoor air pollutant concentrations, in addition to the type of stove.

5.2. Methods

5.2.1. Household selection and study design

A cross-sectional study (Edwards et al., 2007) was conducted to evaluate the

comparative impact of improved and traditional stoves on indoor air pollutant

concentrations (PM2.5, UFP, BC and CO) in the rural village of Bibane (described in

detail in section 4.2.2), during the same period in which the quantification study of

carbonaceous aerosols emissions was conducted. Even though the air quality monitors

were not worn by women, it is assumed that indoor air pollutant concentrations may

be considered as a proxy for personal exposure during cooking hours due to the small

size of the kitchens and their low ventilation rates.

The sample size advisable in cross-sectional sample designs to evaluate changes in

indoor air pollution due to improved stoves depends on the detectable difference in

means and the covariance (COV) of measurements. For improved cookstoves without a

chimney, it is realistic to expect a difference of 60%, and COV of 0.7. Therefore, the

advisable sample size per stove type for this study was estimated to be 22 (Edwards et

al., 2007).

The largest possible sample size of improved cookstoves in Bibane was 15 (number of

improved stoves in good or fair good state and available for the study), so the same

wood specie (dimb) was distributed to all families to reduce the sample size advisable,

as it was done for the quantification study of carbonaceous aerosol emissions (chapter

4).

A total of 22 households was selected, 12 with the traditional stove and 10 with the

improved rocket stove. The 15 improved stoves available were not included in the

study due to time and resource constraints. Information on general kitchen

characteristics was annotated, and the daily fuel wood was measured using a digital

scale of 0−30 kg range with 1 g resolution. General features of study kitchens are

summarized in Table 10, while and Table 11 and Table 12 provide individual

information of each household.

Page 77: INTO THE SMOKE: A RESEARCH ON HOUSEHOLD AIR ...oa.upm.es/49583/1/CANDELA_DE_LA_SOTA_SANDEZ.pdfCada año, la contaminación del aire interior provoca la muerte prematura de 4,3 millones

61

Table 10. Summary of participant kitchen characteristics. Source: own elaboration

Kitchen characteristics N

Stove type

Improved 10

Traditional 12

Wall materials

Earth bricks 16

Tin 1

Thatched 1

Missing 4

Roof materials

Thatched 17

Tin 1

Missing data 4

Kitchen characteristics Mean±sd

Kitchen volume (m3) 22.0±9.3

Free area of inlet/outlet opening (windows+doors+holes) (m2) 2.1±1.6

Fuel consumption (kg) 9.8±5.2

Page 78: INTO THE SMOKE: A RESEARCH ON HOUSEHOLD AIR ...oa.upm.es/49583/1/CANDELA_DE_LA_SOTA_SANDEZ.pdfCada año, la contaminación del aire interior provoca la muerte prematura de 4,3 millones

62

Table 11. Participant kitchen characteristics. Improved cookstoves. Source: own elaboration

Household code BI-01-A BI-02-A BI-03-A BI-04-A BI-05-A BI-06-A BI-07-A BI-08-A BI-09-A BI-10-IA

nº of windows and holes 0 6 1 1 2 1 3 1 0 3

nº of doors 1 1 1 1 1 1 1 1 1 1

Area of windows and holes m2 0 0.082 0.24 0.2 0.072 0.35 0.86 0.58 0 0.32

Area of doors m2 1.65 1.20 1.30 1.26 0.97 2.16 2.026 1.65 0.86 1.21

free area of inlet opening

(windows+holes+

doors) m2

1.65 1.29 1.54 1.46 1.039 2.51 2.89 2.23 0.86 1.54

Kitchen volume (m3) 23.43 16.37 18.28 15.19 21.77 20.38 19.23 48.70 21.70 13.91

Walls material earth bricks earth bricks earth bricks earth bricks earth bricks thatched earth bricks earth bricks tin earth bricks

Roof material thatched thatched thatched thatched thatched thatched thatched tin thatched thatched

Household population

(fraction of standard adult

child of 0-14 years=0.5)

9.50 4.50 9.50 8 7 17 6.50

8.50 16

Wood consumption (kg/day) 5.40 4.60

11.80 20 15.80 6.40 8.27 7.40 10

Page 79: INTO THE SMOKE: A RESEARCH ON HOUSEHOLD AIR ...oa.upm.es/49583/1/CANDELA_DE_LA_SOTA_SANDEZ.pdfCada año, la contaminación del aire interior provoca la muerte prematura de 4,3 millones

63

Household code BI-01-T BI-02-T BI-03-TT BI-04-T BI-05-T BI-06-T BI-07-T BI-08-T BI-09-T BI-10-T BI-11-T BI-12-T

nº of windows and holes 4 2 3 1 1 2 no data 1 no data no data no data 3

nº of doors 1 1 1 1 1 2 no data 1 no data no data no data 1

Area of windows and holes m2 0.15 4.30 0.098 5.21 0.046 0.11 no data 0.038 no data no data no data 0.21

Area of doors m2 1.28 1.22 0.87 1.50 1.09 2.67 no data 1.43 no data no data no data 1.024

free area of inlet opening

(windows+holes+

doors) m2

1.43 5.52 0.97 6.70 1.14 2.78 no data 1.47 no data no data no data 1.23

Kitchen volume (m3) 18.99 30.29 14.49 21.79 10.44 39.28 no data 18.92 no data no data no data 17.32

Walls material earth

bricks

earth

bricks

earth

bricks

earth

bricks

earth

bricks

earth

bricks no data

earth

bricks no data no data no data

earth

bricks

Roof material thatched thatched thatched thatched thatched thatched no data thatched no data no data no data thatched

Household population

(fraction of standard adult

child of 0-14 years=0.5)

4.50 7 4 13 19 9 4.5 7 5 4.5 5.5 9

Wood consumption (kg/day) 9.32 4.66 1.24 13.50 20 15.80 15.54 9.33 6.86 4.3 6.68 8.49

Table 12. Participant kitchen characteristics. Traditional cookstoves. Source:own elaboration

Page 80: INTO THE SMOKE: A RESEARCH ON HOUSEHOLD AIR ...oa.upm.es/49583/1/CANDELA_DE_LA_SOTA_SANDEZ.pdfCada año, la contaminación del aire interior provoca la muerte prematura de 4,3 millones

64

5.2.2. Indoor Air Pollution monitoring

Indoor air pollutant (PM2.5, UFP, BC, CO and LDSA) concentrations were monitored in

all kitchens using the same instrumentation. The instrumentation deployed was a

function of its portability, autonomy, range of concentration, budget and sensitivity.

One household with a traditional stove and one with improved stove were always

monitored simultaneously in order to account for meteorological variability.

Dusttrak (TSI Inc., Shoreview, MN, Model DRX). Aerosol particulate monitor which

monitors real time particle concentrations (PM1, PM2.5, PM10) using a laser photometer

with a reading range of 0–100 mg/m3 and a resolution of 0.001 mg/m3. It was installed

in 21 households during the main cooking periods (lunch and dinner) and during part

of non-cooking periods (several minutes before and after cooking). DustTrak was used

for studies in similar contexts (Chowdhury et al., 2013; Ochieng et al., 2017) and its

performance has been already assessed (Chowdhury et al., 2007; Rivas et al., 2017;

Viana et al., 2015).

DiSCmini (Matter Aerosol). Portable monitor (Testo (Fierz et al., 2011) determining

particle number concentration and mean diameter in the range 10-700 nm, connected

to an impactor with a cutoff at 700 nm to prevent interference with coarse particles.

Anti-static tubing was used during all intercomparison exercises (Asbach et al., 2017;

Todea et al., 2016; Viana et al., 2015). The DiSCmini compares reasonably well with

reference instruments under laboratory and field settings in urban and industrial

environments (Koehler and Peters, 2015; Viana et al., 2015), with an uncertainty of

30% in terms of particle number concentration. However, to this author's knowledge,

no field study of indoor air pollution from stoves used this device. It was installed in 6

households, during the main cooking periods and part of non-cooking periods. This

instrument suffered frequent failures during the study due to the fact that UFP

concentrations monitored were frequently outside its monitoring range.

MicroAeth AE-51 (AethLabs). Portable black carbon monitor with no cyclone at the

inlet. The flow rate of AE-51 was set at the minimum, i.e. 50 mL/min, but due to very

high BC loading in cookstove emissions, the filter strip had to be changed with a high

frequency (every 5 minutes approximately) to prevent saturation (Rehman et al.,

2011). This fact complicated the sampling procedure and resulted in data losses, and

therefore the use of a diluter is advised for future studies. Previous studies have used

this device for indoor BC monitoring studies from cooking activities (Kar et al., 2012;

Rehman et al., 2011). Microaeth was installed in 15 households, during the main

cooking periods and during part of non-cooking periods. The uncertainty of this

instrument was quantified for outdoor air stations as 10% (Viana et al., 2015).

Indoor Air Pollution Meter 5000 Series (IAP meter; Aprovecho Research Center, OR,

USA). Red light scattering photometer to measure PM2.5 concentration, with a reading

Page 81: INTO THE SMOKE: A RESEARCH ON HOUSEHOLD AIR ...oa.upm.es/49583/1/CANDELA_DE_LA_SOTA_SANDEZ.pdfCada año, la contaminación del aire interior provoca la muerte prematura de 4,3 millones

65

range of 0-60 mg/m3 and a resolution of 0.025 mg/m3. It also includes an

electrochemical sensor to measure CO concentration. It was installed in 22

households, during 24-hour periods. This device has already been used in certain

indoor air pollution studies from cookstoves, (Grabow et al., 2013; Sota et al., 2014)

but its uncertainty has not been reported in the scientific literature.

Monitoring devices were deployed according to standard protocols, 1 m away from the

emission source and 1.45 m above the ground to represent breathing position of

cooks, on the side of the stove opposite to open doors and windows (Chengappa et al.,

2007). Inlets were hung approximately 10 cm apart from each other. The sampling

interval was set to 1 minute for all instruments.

Devices were not installed in the same number of households because: i) some of

them failed throughout the campaign (e.g. DisCMini), ii) there were insufficient filter

tickets available, due to the high replacement frequency (Microaeth) and iii) some

battery failed (DustTrak). Differences in the sampling period duration were due to the

devices’ autonomy. While IAP meters had autonomy enough to measure during 24h

periods, DustTrak, DiSCmini and Microaeth had not. Therefore, the latter were

deployed during the main cooking periods (lunch and dinner), and during part of the

background (non-cooking) period. Table 13 and Table 14 describe the devices installed

in each household.

Page 82: INTO THE SMOKE: A RESEARCH ON HOUSEHOLD AIR ...oa.upm.es/49583/1/CANDELA_DE_LA_SOTA_SANDEZ.pdfCada año, la contaminación del aire interior provoca la muerte prematura de 4,3 millones

66

Household code BI-01-T BI-02-T BI-03-TT BI-04-T BI-05-T BI-06-T BI-07-T BI-08-T BI-09-T BI-10-T BI-11-T BI-12-T

IAP meter 5014 5014 5014 5014 5014 5025 5025 5025 5025 5014 5014 5014

DustTrak 2 NO 3 3 3 3 2 3 NO 3 2 2

Microaeth 5 787 5 5 ACS 5 5 5 NO NO NO NO

Discmini MDS IMPACT IMPACT NO NO NO NO NO NO NO NO NO

Table 13. Devices installed per household. Traditional cookstove. Source: own elaboration

Household code BI-01-A BI-02-A BI-03-A BI-04-A BI-05-A BI-06-A BI-07-A BI-08-A BI-09-A BI-10-IA

IAP meter code 5025 5025 5025 5025 5025 5014 5014 5014 5025 5025

DustTrak code 3 3 2 2 3 3 2 2 2 3

Microaeth code 787 5 ACS and 787 787 ACS ACS 5 NO NO NO

Discmini code IMPACT MDS MDS NO NO NO NO NO NO NO

Table 14. Devices installed per household. Improved cookstove. Source: own elaboration

Page 83: INTO THE SMOKE: A RESEARCH ON HOUSEHOLD AIR ...oa.upm.es/49583/1/CANDELA_DE_LA_SOTA_SANDEZ.pdfCada año, la contaminación del aire interior provoca la muerte prematura de 4,3 millones

67

5.2.3. Instrument quality control

Microaeth, Discmini, and DustTrak monitors were previously intercompared for a 24 h

period with their stationary counterparts (Multi-angle absorption photometer (MAAP),

condensation particle counter (CPC) and optical particle counter (OPC), respectively)

(Viana et al., 2015) in an urban air quality monitoring station in Barcelona (Spain).

Detailed results are presented in Figure 24, Figure 25, Figure 26 and Figure 27.

Comparisons were carried out for quality control purposes, i.e. correction equations

were not applied, given the differences between the test (urban) and field (cookstove)

in aerosol load and composition. In addition, the first day of the campaign all of the

monitors were collocated in a chosen kitchen for field validation, evaluating inter-

instrument variability, sensitivity, and consistency (Chowdhury et al., 2013). Detailed

results are presented in Figure 28, Figure 29 and

Figure 30. DustTrak was used for the field-validation of the IAP meter, as explained in

section 5.3.4.

Figure 24. Laboratory comparison. Relationship between DustTrak monitors (code: DustTrak 2 ) and OPC (GRIMM) concentrations (µg/m3) for the period 12/02/2016-26/02/2016 at the Barcelona (BCN) site. Source: own elaboration

Figure 25. Laboratory comparison. Relationship between DustTrak monitors (code: DustTrak 3) and OPC (GRIMM) concentrations (µg/m3) for the period 12/02/2016-26/02/2016 at the Barcelona (BCN) site. Source: own elaboration

Page 84: INTO THE SMOKE: A RESEARCH ON HOUSEHOLD AIR ...oa.upm.es/49583/1/CANDELA_DE_LA_SOTA_SANDEZ.pdfCada año, la contaminación del aire interior provoca la muerte prematura de 4,3 millones

68

Figure 26. Laboratory comparison. Relationship between Microaeth (AE51_787, AE51_ACS and AE51_5) and MAAP concentrations (µg/m3) obtained for the period 12/02/2016-26/02/2016 at the BCN site. Source:own elaboration

Figure 27. Laboratory comparison. Relationship between DISCmini (MDS, MD impact) and CPC concentrations (pt/m3) obtained for the period 12/02/2016-26/02/2016 at the BCN site. Source: own

elaboration

Figure 28. Field intercomparison. Relationship of DustTrak monitors (DST2 and DST3) concentrations obtained at the household BI-01-A. Source: own elaboration.

Page 85: INTO THE SMOKE: A RESEARCH ON HOUSEHOLD AIR ...oa.upm.es/49583/1/CANDELA_DE_LA_SOTA_SANDEZ.pdfCada año, la contaminación del aire interior provoca la muerte prematura de 4,3 millones

69

Figure 30. Field intercomparison. Relationship of Discmini monitors (Impact and MDS) concentrations obtained at the household BI-01-A. Source: own elaboration

Figure 29. Field intercomparison. Relationship of Microaeths monitors (AE51_5, AE51_ACS and AE51_787) concentrations obtained at the household BI-01-A. Source: own elaboration.

Page 86: INTO THE SMOKE: A RESEARCH ON HOUSEHOLD AIR ...oa.upm.es/49583/1/CANDELA_DE_LA_SOTA_SANDEZ.pdfCada año, la contaminación del aire interior provoca la muerte prematura de 4,3 millones

70

5.2.4. Data Analysis

Concentration levels of the different pollutants with the two types of stoves were

processed and arithmetic means (AM), standard deviations (SD), medians, and range

were calculated. The Kruskal–Wallis test was used to test for differences in pollutants

concentration between stove types. This test was selected because indoor air pollution

data from this study doesn’t fit a normal distribution.

In addition, a linear mixed-effects model was constructed to assess the effect of

factors at household-level on indoor air pollution. This model includes fixed and

random effects. The fixed effects were: stove type, kitchen volume, free area of

inlet/outlet opening and the kitchen construction materials. In addition, the random

effects characterize variation due to individual differences. In this study, households

were considered as random effect (Hu et al., 2014).

The formula that describes the model is: yit = β0 + β1x1 + β2x2...βnxn + ui + εit (1)

where yit represents the pollutant concentration value modelled for household i at day

t; β0 represents the intercept value; β1 through to βn represent the fixed effect variable

coefficients for variables x1 through xn; µi are random effects at household level,

assumed normally distributed with mean 0 and variance su2, and εit are the model

residuals assumed normally distributed with mean 0 and variance su2 (Rivas et al.,

2015). Data analysis was performed using IBM SPSS Statistics Base and Python

software.

Page 87: INTO THE SMOKE: A RESEARCH ON HOUSEHOLD AIR ...oa.upm.es/49583/1/CANDELA_DE_LA_SOTA_SANDEZ.pdfCada año, la contaminación del aire interior provoca la muerte prematura de 4,3 millones

71

5.3. Results and discussion

5.3.1. Indoor air pollutant concentrations during cooking activities

Figure 31 shows a representative example of particulate and gaseous indoor

concentrations for a household using the rocket stove (BI-02-A), covering the two main

daily cooking activities (lunch and dinner) and a portion of non-cooking period. This

example was selected based on data availability for all the parameters under study:

particle number concentration, mass (PM10, PM2.5, PM1) and mean diameter, as well as

BC and CO concentrations. LDSA concentrations were also available (not shown

graphically, for clarity).

The pollutant time series shown in Figure 31, point to cooking as the main emission

source of indoor air pollution and personal exposure (especially for women and

children). During cooking periods, number concentrations (N) and particle mass (PM2.5)

reached up to 3106 #/cm3 and 2000 µg/m3, respectively. LDSA and CO concentrations

climb to 4800 μm2/cm3and 105 ppm, respectively. As previously found in past studies

using solid fuel in biomass stoves, results presented in Figure 31 support the evidence

that particulate emissions (PM, N and BC) and CO are co-emitted (Chengappa et al.,

2007; Naeher et al., 2000; Northcross et al., 2010).

Figure 31. PM1, PM2.5, PM10 (µg/m3), BC (µg/m3), UFP (/m

3), mean diameter (nm) and CO (ppm) concentrations measured

on March 3, 2016 in a household with rocket cookstove Noflaye Jegg, during lunch and dinner preparation at household BI-02-A . Source: own elaboration

Page 88: INTO THE SMOKE: A RESEARCH ON HOUSEHOLD AIR ...oa.upm.es/49583/1/CANDELA_DE_LA_SOTA_SANDEZ.pdfCada año, la contaminación del aire interior provoca la muerte prematura de 4,3 millones

72

In the example shown in Figure 31, during non-cooking activities, average indoor CO

and BC concentrations were close to zero, 1.28 ppm and 1.33 µg/m3 respectively.

Another study conducted in rural villages in Northwest Bangladesh also showed very

low indoor CO concentrations (0.2 ppm) during non-cooking activities (Chowdhury,

2012), whereas slightly higher BC concentrations (3.7 μg/m3 ±0.9,) were monitored in

villages of northern India surrounded by traffic (Kar et al., 2012).

Average indoor background concentrations were PM10=159.9 µg/m3, PM2.5=99.1µg/m3

and PM1=88.5 µg/m3, similar to other studies in rural areas in India (PM2.5 from 45 to

207 µg/m3) (Sambandam et al., 2015). High PM10/PM2.5 ratios (1.6 on average) were

monitored during non-cooking periods due to indoor and outdoor dust, which was re-

suspended by wind, movement of inhabitants, and animals in the kitchen, among

other causes. In the present study, during cooking periods PM10/PM2.5 ratios were on

average 1.1. This highlights that particles are coarser during non-cooking periods and

the important contribution of fine particles to indoor air during cooking periods.

Regarding UFPs, mean background N concentrations during non-cooking hours were

5,568 #/cm3, 99% lower than during cooking periods. These results are lower than

those reported by Chowdhury et al. (2012) in Bangladesh (Chowdhury, 2012), where

mean UFP concentrations during non-cooking hours was 15,000 ± 7,200 #/cm3. This

may be due to differences in ventilation rates and/or lack of enough time between

cooking periods to ventilate the room. Finally, in this example, LDSA mean

concentration during non-cooking periods was 50,5 μm2/cm3, similar to rural

background sites measurements found in other studies (Casino, Italy, 69 μm2/cm3)

(Buonanno et al., 2012).

Although ambient air quality data was not monitored simultaneously with indoor air

pollution, contributions to household pollution from other combustion sources such as

transport or industry were not considered significant in Bibane. Households were

sufficiently dispersed so that neighbourhood cooking activities did not affect other

households, and no significant additional combustion sources (e.g., forest fires) were

recorded during the study period. A study conducted in rural villages of Western Africa

showed that 74−87% of total PM2.5 mass concentrations in cooking areas was from

biomass burning particles, together with a small percent of crustal material and soil

dust, from the Sahara desert (Zhou et al., 2014).

Figure 32 shows, for the same representative household (BI-02-A), the detailed time-

series of pollutant concentrations during a single combustion episode (lunch

preparation). The aim of this analysis was to assess emission patterns for different

pollutants as a function of the cooking cycle.

Page 89: INTO THE SMOKE: A RESEARCH ON HOUSEHOLD AIR ...oa.upm.es/49583/1/CANDELA_DE_LA_SOTA_SANDEZ.pdfCada año, la contaminación del aire interior provoca la muerte prematura de 4,3 millones

73

Fire was lit at 11.31h am (marked with a red star in Figure 32) and a peak of all

pollutants (particles and CO) was observed immediately after, showing the fast impact

of combustion on indoor concentrations (10,000 µg/m3 for PM2.5, 1,000 µg/m3 for BC,

2.5106/cm3, 4,800 μm2/cm3 for LDSA, 30 ppm for CO). Mean particle diameter

decreased with increasing particle number concentrations (from 62 nm during non-

cooking to 51 during cooking hours), although it did not show such a fast evolution,

probably linked to instrumental limitations (particle diameter measurements were

close to the instrument’s detection limit). In this study, the high emissions during the

initial few minutes of the combustion phase were due to the use of straw as accelerant

to start the fire. Other authors have reported similar results with the use of ignition

products (Arora et al., 2013; Roden et al., 2006).

After the initial increase in concentrations (recorded for all pollutants), PM

concentrations decreased (down to 2,000 µg/m3), while the impact of emissions on

indoor air were more persistent over time with regard to particle number

(concentrations remaining at 1.5106/cm3). This has implications with regard to

population (mainly women and children) exposure to UFPs (Ko et al., 2000).

Conversely, PM mass concentrations progressively accumulated and increased over

time, reaching at the end of the cooking period similar concentrations to those

monitored during its initial phase (8,000 µg/m3).

As shown in Figure 32, PM concentrations showed a larger variability throughout the

combustion period than other parameters such as UFP or CO concentrations. Pollutant

Figure 32. PM1, PM2.5, PM10, BC, UFP, Size, LDSA and CO concentrations measured during lunch preparation on March 3 2016 in a household with improved cookstove Noflaye Jegg at household BI-02-A. Source: own

elaboration

Page 90: INTO THE SMOKE: A RESEARCH ON HOUSEHOLD AIR ...oa.upm.es/49583/1/CANDELA_DE_LA_SOTA_SANDEZ.pdfCada año, la contaminación del aire interior provoca la muerte prematura de 4,3 millones

74

peaks during cooking occurred when fuel was added or moved, due to activities such

as placing or removing the cooking pot on the fire, stirring the food, or as a result of

external factors (e.g., changing wind speed). The high variability in mean particle

diameter during cooking periods is also noteworthy, as it indicates that different types

of emissions were generated during the different types of combustion stages

(influenced by fuel feeding, strength of the flame, ventilation, etc.).

As expected, PM10, PM2.5 and PM1 concentrations were highly similar during the

cooking period, indicating that biomass combustion was dominated by fine and

ultrafine particles. Indeed, other works analysed particle size distribution during

biomass burning showing that PM1 was the main contributor to PM2.5 and PM10

fractions (Arora et al., 2013; Park and Lee, 2003).

Although fire was completely extinguished at 13.03h, PM and BC concentrations began

to decrease toward the end of the food preparation period (12.52 h), when the cooker

stopped feeding the stove and used the residual heat to finish the cooking. PM and BC

levels returned to concentrations closer to those prior to the cooking period (117

µg/m3 and 1.2 µg/m3 for BC) around 10 minutes after the fire was extinguished. CO

and N concentrations increased during the final period of cooking, when using residual

heat. Once the fire was extinguished, N showed a similar trend to PM and BC, while

concentrations of CO showed a gradual decrease over 25 minutes. The same behaviour

for CO was found in previous studies (Chowdhury, 2012), and it is indicative of the

remaining charcoal smouldering once the fire was extinguished.

5.3.2. Indoor air pollutant concentrations during cooking periods as a

function of the type of stove

Table 15 presents the descriptive statistics for pollutant concentrations during cooking

periods in the studied households, as a function of the type of stove used: traditional

or improved. In total, 8 households were available with traditional stoves, and 9 for

improved ones. However, not all pollutants were available for both types of

households (irrespective of the use of traditional or improved cookstoves) given that

instrumental failures were frequent due to the high concentrations monitored. The

instrument which was most affected by technical failures was the DiscMini, and thus

particle number, LDSA concentrations and mean particle diameter were only available

for 3 improved cookstove and 3 traditional cookstove households. Results shown refer

only to cooking periods. PM1, PM2.5 and PM10 values during combustion showed less

than 10% of difference for both types of stoves pointing to a predominance of fine

particles. For clarity reasons, PM2.5 was selected for presentation in Table 15 as an

indicator for PM10 and PM1 mass fractions, as it is most commonly used in indoor air

pollution studies and international standards.

Page 91: INTO THE SMOKE: A RESEARCH ON HOUSEHOLD AIR ...oa.upm.es/49583/1/CANDELA_DE_LA_SOTA_SANDEZ.pdfCada año, la contaminación del aire interior provoca la muerte prematura de 4,3 millones

75

During cooking hours, results show that PM, UFP, LDSA and CO concentrations in

homes using improved stoves were statistically lower than in those with traditional

stoves. Conversely, BC concentrations were higher in improved cookstove households.

All of these differences were statistically significant according to the Kruskal-Wallis test

(p<0.05).

Based on this analysis, reductions (or increases) in pollutant concentrations during

cooking as a function of the stove were pollutant-dependent. The largest reductions

due to improved cookstove implementation were observed for PM2.5, with a 75.4%

reduction in median concentrations. CO concentrations decreased by 54.3%, followed

by particle number concentrations (30.0%) and LDSA (14.2%). As a result, the use of

improved cookstoves in the households of this study (between 3 and 9, depending on

the pollutant) showed clear and statistically significant improvements for indoor air

quality with regard to particle mass (PM2.5), particle number, LDSA and CO

concentrations.

Other studies also found CO and PM2.5 concentration reductions with the rocket stove,

when compared to a traditional three stone fire: De la Sota et al. found 36% and 33%

of 24-hr mean percent change for PM2.5 and CO respectively in rural Senegal (Sota et

al., 2014); and median percent reductions in 24-h PM2.5 and CO concentrations ranged

from 2 to 71% and 10–66% respectively in southern India (Sambandam et al., 2015).

Even though this is a positive result in relative terms, it should be noted that, in

absolute terms, median concentrations during cooking periods remained high even

when improved cookstoves were used: 1,780 µg/m3 for PM2.5, 1.5*106/cm3 for UFPs,

4,014.2 μm2 /cm3 for LDSA, and 34.75 ppm for CO.

Following an opposite behaviour, median BC concentrations increased by 36.1% (also

statistically significant), from 497.7 µg/m3 with traditional stoves to 677.6 µg/m3 with

improved cookstoves. This result highlights the complexity linked to the estimation of

indoor air pollution impact of cookstove interventions, as improved stoves do not

perform similarly for different pollutants (Arora and Jain, 2015; Jetter and Kariher,

2009; Kar et al., 2012; MacCarty et al., 2010, 2008b; Roden et al., 2009).

These findings are consistent with results presented in chapter 4, showing that the

Noflaye Jegg rocket stove emitted more BC (EC) than the traditional stove. The

consequence of this is that negative health effects from BC exposure can still be

expected after installation of improved stoves, in addition to climate warming effects

(Baumgartner et al., 2014).

.

Page 92: INTO THE SMOKE: A RESEARCH ON HOUSEHOLD AIR ...oa.upm.es/49583/1/CANDELA_DE_LA_SOTA_SANDEZ.pdfCada año, la contaminación del aire interior provoca la muerte prematura de 4,3 millones

76

Type of

stove

PM2.5(µg/m3) Particle Number Concentration (N) (pt/cm

3) Mean diameter (nm)

n Am ± sd Median Range % of

change in

median

n Am ± sd Median Range % of

change in

median

n Am ± sd Median Range % of

change in

median Improved

cookstove 9 4,621.9±8,387.1 1,780 20-

84,200 -75.4%*

3 1,713,620±

1,079,780 1,529,080 12,869-

6,721,660 -30.01%*

3 50.4±18.4 46 24.4-

244.6 20.7%*

Traditional

cookstove 8 10,563.9±10,328.5 7,240 62-

54,600 3 2,576,060±

1,913,410 2,184,800 372-

7,594,540 3 41.8±24.5 38.1 10-299

Type of

stove

Lung Deposited Surface Area (LDSA) (μm2 /cm

3) Black Carbon (BC) (µg/m

3) Carbon monoxide (CO) (ppm)

n Am ± sd Median Range % of

change in

median

n Am ± sd Median Range % of

change in

median

n Am ± sd Median Range % of

change in

median Improved

cookstove 3 4,201.3± 1,962.6 4,014.2 30.9-

10,886.5 -14.2%*

7 1,042.8±

1213.5 677.6 0.2-

11,806.6 36.1%*

8 23.8±26.5 15.9 0.01-

285.8 -54.3%*

Traditional

cookstove 3 5,085.9± 2,980.1 4,680.2 6.5-

11,432.3 8 767.9±874.3 497.7 0.02-

5,778.9 8 41.3±26.8 34.8 0.004-

166.3

Table 15. Descriptive statistics of indoor air pollutants concentrations by stove type. n=number of households sampled. Am=arithmetic mean. Sd= standard deviation. *significant variation (p<0.05; Kruskal-Wallis tests). % of change in median compared to the traditional stove. Source: own elaboration

Page 93: INTO THE SMOKE: A RESEARCH ON HOUSEHOLD AIR ...oa.upm.es/49583/1/CANDELA_DE_LA_SOTA_SANDEZ.pdfCada año, la contaminación del aire interior provoca la muerte prematura de 4,3 millones

77

Finally, mean particle diameter increased from 42 to 50 nm (20.7%, Table 15) when

rocket cookstoves were used. Previous studies showed a shift toward finer particles

with some types of improved stoves (Arora et al., 2013; Just et al., 2013). For example,

Just et al., 2013, found a mean particle diameter of 35 nm with the rocket stove and 61

nm with the three stone fire. However, such studies were conducted under controlled

conditions and may not represent field conditions. Moreover, particle size distributions

are significantly affected by type of cookstoves and fuels (Just et al., 2013; Tiwari et al.,

2014). Finally, higher LDSA values for three stone can be attributed to larger emissions

of ultrafine particles (Sahu et al., 2011).

With regard to previous studies, although several authors assessed indoor air pollution

from cookstoves use in the field (Armendáriz et al., 2010, 2008; Chengappa et al.,

2007; Chowdhury et al., 2013; Hu et al., 2014; MacCarty et al., 2007b; Ochieng et al.,

2013; Park and Lee, 2003; Ruth et al., 2014; Sambandam et al., 2015; Siddiqui et al.,

2009; Sota et al., 2014; Zuk et al., 2007), most of them present 24h or 48h average

data, which do not reflect acute exposure levels observed during cooking periods.

There are few field studies differentiating between cooking and non-cooking periods,

whose results are summarized in Table 16, together with those obtained in this work.

LDSA values are not included since, to this author's knowledge, none of the field

studies that evaluated this parameter (Hosgood et al., 2012; Sahu et al., 2011) showed

indoor concentration values separately for both cooking and non-cooking periods.

Page 94: INTO THE SMOKE: A RESEARCH ON HOUSEHOLD AIR ...oa.upm.es/49583/1/CANDELA_DE_LA_SOTA_SANDEZ.pdfCada año, la contaminación del aire interior provoca la muerte prematura de 4,3 millones

78

Pollutant concentration during cooking activities

This study Other studies

CO (ppm)

Traditional:

Median: 34.8

Mean (SD): 41.3 (26.8)

Rocket stove:

Median: 15.9

Mean (SD): 23.8 (26.5)

(Balakrishnan et al., 2015), India

Traditional:

Median: 57.6

Mean (SD): 56.2 (13.8)

Forced-draft advanced cookstove:

Median: 20

Mean (SD): 21.5 (4.8)

PM2.5 (µg/m3)

Traditional:

Median: 7,240

Mean (SD): 10,563.9 (10,328.5)

Rocket stove:

Median: 1,780

Mean (SD): 4,621.9 ( 8,387.1)

(Balakrishnan et al., 2015), India

Traditional:

Median: 1,451

Mean (SD): 1,991 (2060)

Forced-draft advanced combustion stove:

Median:1,046

Mean (SD): 1,813 (2,686)

N (pt/cm3)

Traditional:

Median: 2,184,800

Mean(SD):2,576,060(1,913,410)

Rocket stove:

Median: 1,529,080

Mean(SD):1,713,620±1,079,780

(Zhang et al., 2012), China

Traditional:

Mean: 290,000

(Chowdhury, 2012), Bangladesh

Chimney biomass improved stove:

Mean (SD):75,000 (31,000)

Mean diameter

(nm)

Traditional:

Median: 38.1

Mean (SD): 41.8 (24.5)

Rocket stove:

Median: 46

Mean (SD): 50.4 (18.3)

(Zhang et al., 2012), China

Traditional:

Mean (start combustion-10min): 40-

50

Mean (10-15 min): 63

BC (µg/m3)

Traditional:

Median: 497.7

Mean (SD): 767.9 (874.3)

Rocket stove:

Median: 677.6

Mean (SD): 1,042.8 (1,213.5)

(Kar et al., 2012) , India

Traditional:

Mean (SD):127.6 (103.5)

Natural draft gasifier:

Mean (SD):75.5(59.4)

Table 16. Summary of pollutant concentration during cooking activities reported for previous studies and within this study. Source: own elaboration

Table 16 shows that results were generally higher in comparison with other field

studies, with the exception of CO and mean particle diameter, which are indeed

comparable. This significant variation might be explained by the differences between

studies regarding fuels and stoves characteristics, variations in cooking habits, stove

location, ventilation, etc. (Just et al., 2013; Siddiqui et al., 2009).

Page 95: INTO THE SMOKE: A RESEARCH ON HOUSEHOLD AIR ...oa.upm.es/49583/1/CANDELA_DE_LA_SOTA_SANDEZ.pdfCada año, la contaminación del aire interior provoca la muerte prematura de 4,3 millones

79

Figure 33. Box-plots of pollutants concentrations for three-stone fire and the rocket stove. The boxplot presents minimum, first quartile, median, (middle dark line), third quartile and maximum values. Source: own elaboration.

Page 96: INTO THE SMOKE: A RESEARCH ON HOUSEHOLD AIR ...oa.upm.es/49583/1/CANDELA_DE_LA_SOTA_SANDEZ.pdfCada año, la contaminación del aire interior provoca la muerte prematura de 4,3 millones

80

Moreover, the range of pollutant concentrations during cooking periods is also highly

variable within stove types (Commodore et al., 2013). Figure 33 presents box-plots of

pollutant concentrations for the three-stone fire and the rocket stoves.

The large variability observed in Figure 33 within stove type can be attributed to a

myriad of factors (e.g., variability in cookstove use and time, activity patterns, weather

conditions, household room configuration and ventilation, cooking behaviours, etc.)

(Clark et al., 2013). It highlights that averaging concentrations across households or

experiments is a major limitation when addressing cookstove studies, as opposed to

other types of air quality studies (e.g., ambient air). As shown in Figure 33¡Error! No se

encuentra el origen de la referencia., the mean/median values should not be

considered fully representative of the datasets, and other metrics of data variability

(e.g., standard deviation, interquartile range) are essential to understand the data

distribution.

5.3.3. Effect of household-level determinants on indoor air pollution

A Linear mixed effect modelling was performed to identify factors at household-level

that might have a significant role in determining PM2.5 concentrations during cooking

periods (Hu et al., 2014).

Factors considered for inclusion as fixed effects were: stove type (traditional,

improved), walls material (earth bricks, thatched, tin), roof material (thatched, tin),

kitchen volume (m3), free area of inlet/outlet opening (m2), fuel consumption

(continuous, kg/day) and elapsed time (minutes). Elapsed time was included as a

continuous variable to examine temporal changes in concentrations (Albalak et al.,

2001). Fuel characteristics were not considered as a factor, since the same type of

wood was distributed between participants during the sampling period.

Other factors, such as type of meal prepared, meteorological conditions or cooker

behaviours were not measured, so their influences were included in the model as

random effect at household level, with a scalar (variance component) covariance

structure (Hu et al., 2014).

Table 17 provides estimates of individual parameters (β), as well as their standard

errors and confidence intervals. The variables selected for inclusion in the final model

were those which significantly impacted PM2.5 measurements (p<0.05) and

contributed to the lowest Akaike information criterion (a measure of the relative

quality of statistical models for a given set of data) score (Hu et al., 2014). To improve

normality and variance homogeneity, log-transformed concentrations were used.

Page 97: INTO THE SMOKE: A RESEARCH ON HOUSEHOLD AIR ...oa.upm.es/49583/1/CANDELA_DE_LA_SOTA_SANDEZ.pdfCada año, la contaminación del aire interior provoca la muerte prematura de 4,3 millones

81

Parameter Coefficient (β)

(log-µg/m3)

Standard error 95 % confidence interval

Type of stove

Improved

Traditional

3.61

3.89

0.28

0.27

3.06;4.16

3.36;4.42

Roof material

Thatched

Tin

-0.69

0*

0.16

0

-1.0034;-0.37

-

Wall material

Earth bricks

Thatched

Tin

0.49

1.02

0*

0.047

0.076

0

0.40;0.58

0.87;1.17

-

Kitchen volume (m3) -0.019 0.0054 -0.029;0.0081

Free area of inlet opening (m2) -0.12 0.015 -0.15;-0.092

Daily Fuel consumption (kg/day) 0.040 0.0030 0.034-0.045

Variation explained between households (%) 35

Table 17. Linear Mixed Effect Modelling of log-transformed PM2.5. *This parameter is set to zero because it is the reference level for the model Source: own elaboration

The estimate of the “Elapsed time” parameter was not significant (p>0.05), and

therefore it was removed from the model. Coefficients indicate an increase in PM2.5

concentration (in log-µg/m3) for traditional cookstoves with respect to improved

cookstoves.

For the roof material variable, PM2.5 concentration decreased with thatch roofing with

respect to tin, which is the reference format (β=0). This is due to the fact that thatched

roofs enhance ventilation and further decrease indoor air pollutants compared to

homes with metal roofs, also observed in previous studies (Ruth et al., 2014).

Regarding wall materials, thatched walls resulted in the highest predicted PM2.5.

However, as only one house had tin walls (see Table 17), this material was not

considered representative and therefore more samples should be included to analyse

its effect on indoor air pollutant concentrations. Results evidenced that kitchen

volume has a significant negative association with PM2.5, with an approximately 1.9%

decrease in log-PM2.5 concentrations for every unit increase in volume. This is

consistent with other studies (Albalak et al., 2001).

Finally, the free area of inlet/outlet opening (sum of areas of windows, doors and

other holes in the kitchen) (Iyengar, 2015) was correlated with PM2.5 concentrations,

with a 12% decrease in log-PM2.5 concentrations for every unit increase in free area of

inlet opening. Results are coherent with other studies showing that particulate levels

can significantly decrease by the existence of opening structures (Baumgartner et al.,

2011; Bruce et al., 2004; Dasgupta et al., 2006; Park and Lee, 2003; Ruth et al., 2014).

Page 98: INTO THE SMOKE: A RESEARCH ON HOUSEHOLD AIR ...oa.upm.es/49583/1/CANDELA_DE_LA_SOTA_SANDEZ.pdfCada año, la contaminación del aire interior provoca la muerte prematura de 4,3 millones

82

As shown in Table 17, the linear mixed effect model explained 35% of the variation of

indoor PM2.5 concentration. The remaining variation may reasonably be due to

imperfect specification of variables, such as cooking time, other household combustion

sources, etc. (Pokhrel et al., 2015). It should be noted that this study aimed at

evaluating the association of PM2.5 concentrations with factors at household level

rather than obtaining a predictive model.

PM2.5 was selected to perform the linear mixed effect model because it is the pollutant

analysed in a greater number of households (8 traditional and 9 improved). Since the

other pollutants are simultaneously produced during biomass combustion (Chengappa

et al., 2007), factors at household level could be assumed to affect in a similar way,

although additional experiments would be necessary to confirm this hypothesis.

5.3.4. Comparison between IAP meter and DustTrak units

As previously explained, IAP meter units were co-located with DustTrak monitors. The

aim of this comparison was to assess the performance of this lower cost instrument

(de la Sota et al., 2017). It is relevant to present the correlation between these two

methods because although the IAP meter has been used in a number of studies

(Gorritty et al., 2011; Grabow et al., 2013; Sota et al., 2014), there are no publications

covering field assessments of its performance compared to other commercial light-

scattering measuring instruments.

Figure 34 shows the correlation of one DustTrak (code 3) with one IAP meter (code

5025) during measurements for the same household previously presented in Figure 31

and Figure 32, including both cooking and non-cooking periods to cover the entire

range of typical daily variation of household PM2.5 concentrations.

y = 0,1751x + 10,708 R² = 0,7314

0

500

1000

1500

2000

2500

3000

3500

4000

4500

0 2000 4000 6000 8000 10000 12000 14000 16000 18000 20000

PM

2.5 (

µg/

m3 )

, me

asu

red

by

IAP

me

ter

PM2.5 (µg/m3), measured by DustTrak

Figure 34. Correlation of DustTrak response with IAP meter for PM2.5 concentrations (µg/m3) at

1 min resolution. Source: own elaboration

Page 99: INTO THE SMOKE: A RESEARCH ON HOUSEHOLD AIR ...oa.upm.es/49583/1/CANDELA_DE_LA_SOTA_SANDEZ.pdfCada año, la contaminación del aire interior provoca la muerte prematura de 4,3 millones

83

The response of the IAP meter showed a relatively good linearity with the DustTrak on

a minute by minute basis (R2 =0.73). However, the correlation equations showed a

large inter-variability across the co-location tests, and in certain households there was

no such a good fit. Table 18 shows the results for all the co-location tests. R2 values

ranged between 0.12 and 0.87 in all of the households where IAP meters and

DustTraks were co-located. In total, a 66.7% of the households (10 households)

exhibited R2 values higher than 0.75, while in the remaining 5 cases the dispersion of

the data was considered high (R2<0.75).

Household code. Improved cookstove

DustTrak vs IAP meter

BI-01-I (DST 3 vs IAP meter 5025)

BI-02-I (DST 3 vs IAP meter 5025)

BI-03-I (DST 2 vs IAP meter 5025)

BI-06-I (DST 3 vs IAP meter

5014)

BI-07-I (DST 2 vs IAP meter 5014)

BI-08-I (DST 2 vs IAP meter 5014)

BI-09-I (DST 2 vs IAP meter 5025)

y = 0.25x + 8.15 R² = 0.8143

y = 0.18x + 10.71 R² = 0.73

y = 0.21x - 4.01 R² = 0.77

y = 0.11x - 39.71 R² = 0.78

y = 0.32x + 39.11 R² = 0.63

y = 0.31x + 47.97 R² = 0.87

y = 0.22x + 9.89 R² = 0.56

Household code. Traditional cookstove

BI-01-T (DST 2 vs IAP meter 5014)

BI-03-T (DST 3 vs IAP meter 5014)

BI-05-T (DST 3 vs IAP meter 5014)

BI-06-T (DST 3 vs IAP meter

5025)

BI-07-T (DST 2 vs IAP meter 5025)

BI-11-T (DST 2 vs IAP meter 5014)

BI-12-T (DST2 vs IAP meter

5014)

y = 0.21x + 673.74 R² = 0.12

y = 0.22x - 6.14 R² = 0.79

y = 0.20x + 13.52 R² = 0.70

y = 0.12x - 91.17 R² = 0.80

y = 0.15x + 57.47 R² = 0.82

y = 0.16x + 88.42 R² = 0.78

y = 0.17x - 111.90 R² = 0.89

This variability in the responses of the IAP meter with regard to the DustTrak

(considered here to be the internal reference) may be explained by different reasons:

i) the fact that they are different instruments operating on different principles

(different laser characteristics); ii) DustTrak and IAP meter were not placed at the exact

same point in the kitchen (the inlets of the devices were hung approximately 10-20 cm

apart from each other, so it may occur that devices were not equally affected by

factors such as ventilation; iii) DustTrak zero was set with an ambient free of pollution

(using a zero filter), whereas the IAP meter background was PM2.5 from ambient air.

Therefore, results from IAP meter presented in this study should only be interpreted as

relative measures, since background concentrations were not considered in the

reported PM2.5 values; and to estimate the additional pollution directly caused by

cooking, it would be necessary to measure background concentrations in the area; and

iv) Two different IAP meters were used during the study. Although both IAP meters

belong to the 5000 series, previous studies showed that using inexpensive off-the-shelf

smoke detector technology may suffer from substantial intra-unit variability

(Chowdhury et al., 2007). This result indicates the need of individual instrument

calibrations, which is also necessary when using other more costly air pollution

monitors (Chowdhury, 2012; de la Sota et al., 2017; Viana et al., 2015).

Table 18. Field intercomparison. Relationship of IAP meter(5014 and 5025) and DustTrak (2 and 3) concentrations. DustTrak vs IAP meter

Page 100: INTO THE SMOKE: A RESEARCH ON HOUSEHOLD AIR ...oa.upm.es/49583/1/CANDELA_DE_LA_SOTA_SANDEZ.pdfCada año, la contaminación del aire interior provoca la muerte prematura de 4,3 millones

84

6. Conclusions

Page 101: INTO THE SMOKE: A RESEARCH ON HOUSEHOLD AIR ...oa.upm.es/49583/1/CANDELA_DE_LA_SOTA_SANDEZ.pdfCada año, la contaminación del aire interior provoca la muerte prematura de 4,3 millones

85

In recent years, addressing cookstove black carbon emissions has aroused interest as a

mean to mitigate this potent short-lived forcing agent (Grieshop et al., 2011). Among

other advantages, the inclusion of black carbon (BC) co-benefits in cookstoves projects

may provide an opportunity to attract new additional funding for further scaling up of

these activities (Hamrick, 2015). However, most stoves implemented worldwide have

not yet been tested with regard to their BC emissions (Jetter, 2015). Moreover, new

improved and clean cooking solutions (including both fuels and stoves) are being

developed continuously around the globe, each one with very different health, social,

environment, and economic related potential impacts, which in turn can sharply vary

from region to region.

This thesis furthers the body of knowledge in the field of climate impacts as a result of

cooking with traditional biomass stoves and the role played by alternative cooking

solutions, with a particular emphasis on black carbon and co-pollutants emissions, and

focusing in Western Africa. The findings of this study also add some insights to the

small number of current quantitative studies on indoor air pollution from cooking in

this region.

In chapter 3, two low cost optical techniques to determine black carbon emissions, a

cell-phone based system and a reflectometer, were tested against a more

sophisticated reference system, based on thermo-optical analysis. Results show that,

for the aerosol types and concentrations tested, the lower-cost methods may

constitute good alternatives for the estimation of black carbon concentrations on filter

substrates in cookstove studies under resource-constrained conditions. This finding

can help to leverage important methodological barriers, as black carbon measurement

techniques usually pose a challenge in terms of their high upfront cost and level of

technical expertise required for their operation (Ramanathan et al., 2011).

Prior to the application of either of the low cost methods, locally-determined

correction coefficients must be calculated by comparison with a reference method

based on a statistically significant number of samples for each specific study. The

calibration should be carried out using filters collected in the field and under the same

conditions expected in the field, given that laboratory conditions will not yield the

same type of emissions. Results also highlight the need to take into account the

aerosol load on the filter and the filter substrate when calibrating lower cost methods.

The easy use of the cell-phone system opens up options for its application in science

projects which actively involve citizens in scientific endeavours, also known as citizen

science. This methodology could be useful for the process of data collection, which

would increase the amount and frequency of data being collected, reducing the time

and cost of sampling. Moreover, this would rise awareness regarding the

environmental and health issues related to cookstoves and may lead to positive

behavioural changes (Wyles et al., 2016).

Page 102: INTO THE SMOKE: A RESEARCH ON HOUSEHOLD AIR ...oa.upm.es/49583/1/CANDELA_DE_LA_SOTA_SANDEZ.pdfCada año, la contaminación del aire interior provoca la muerte prematura de 4,3 millones

86

Chapter 4 provided original cooking-related carbonaceous (elemental and organic

carbon, EC and OC) aerosols emissions estimates in Senegal, where previous

information was not available. Four biomass stoves, each one representing a type of

combustion, were tested in the laboratory using the standardized water boiling test

and filter samples analysed with a thermal-optical method.

The highest average elemental carbon emission factor (EF) (g/MJ) was found for the

rocket stove. This was also found in previous studies, and can be explained by the fact

that this type of stove has a strong draft which typically results in higher flame

temperatures and elemental carbon emissions. What is remarkable is the fact that the

rocket stove analysed in this study did not show a reduction in fuel wood consumption

per task completed (i.e., Water Boiling Test, WBT) with respect to the three stones fire.

As a result, this type of stove was the one with the highest EC total emissions. On the

other side, the gasifier showed the highest fuel savings and very low elemental and

organic carbon emission factors, so this type of stove had the smallest total emissions

per WBT. Similarly to other studies, the basic improved ceramic stove showed similar

results to the three stones fire, both in emission factors and in fuel consumed per

WBT.

Results also showed that emission factors of EC and OC were also dependent on the

wood specie burned, but further analysis are needed to study the effect of physical

and chemical characteristics on carbonaceous aerosol emissions.

Likewise, the three stones and the rocket stove were tested under uncontrolled

conditions in a rural village of Senegal, to understand how real cooking practices in this

region influenced EC and OC emissions. The rocket stove had significantly lower PM2.5

EF when compared with the three stones, but, similarly to laboratory results, it showed

higher black EC EFs and did not reduced the daily fuelwood consumption. Moreover,

OC, which has been associated with cooling properties, was reduced with the rocket

stove. Therefore, carbonaceous emissions of the rocket stove produces a net positive

warming effect in comparison with the traditional stove.

Differences between laboratory and field results were found, although less than

expected based on previous studies. Field results were found to be more similar to

results from the simmer phase of the WBT. This is coherent with field observations

during testing activities, which showed that during meal preparations, cookers

maintained a constant and low-medium flame almost all the time. This finding could

be very useful to understand how representative are the standardized studies of actual

cooking practices in Senegal, and other West African countries.

Finally, total EC and OC emissions and the net climate effect from fuelwood burning at

household level in Senegal and West Africa were estimated. Quantifying the

contribution that traditional and alternative cookstoves play in the emissions of

Page 103: INTO THE SMOKE: A RESEARCH ON HOUSEHOLD AIR ...oa.upm.es/49583/1/CANDELA_DE_LA_SOTA_SANDEZ.pdfCada año, la contaminación del aire interior provoca la muerte prematura de 4,3 millones

87

carbonaceous aerosol provides more accurate data that may contribute to reduce

uncertainties of emission inventories and climate prediction models at regional level.

This is especially important in the case of black carbon, as its climate impacts are highly

dependent on the place where it is emitted (Hamrick, 2015).

This information should give additional impetus for improved and clean cookstove

dissemination actions to be considered by governments and decision makers in West

African countries in the Nationally Appropriate Mitigation Actions (NAMAs) and other

funds and strategies targeted to climate change mitigation.

Chapter 5 presented the first real-world assessment of household concentrations of

PM2.5, ultrafine particles, black carbon and carbon monoxide from the combustion of

fuelwood for cooking in Senegal. Results confirmed that household air pollution in this

area is mainly due to cooking activities and showed that the installation of the rocket

stove contributed to a significant reduction of total fine and ultrafine particulate

number and CO concentration with respect to traditional stoves, but increased indoor

BC concentrations. This increase of indoor BC concentrations with the rocket stove

despite the fact that total fine particle concentration decrease is coherent with results

presented in chapter 4.

The study also identified factors at household-level with a significant role in

determining pollutant concentrations during cooking periods. Thatched roofs were

found to conduct to lower IAP levels, when compared to homes with metal roofs.

Higher values of the free area of inlet/outlet opening (i.e. sum of areas of windows,

doors and other holes in the kitchen) also conducted to a reduction of indoor air

pollution. Finally, though to a lesser extent, kitchen volume was also found to affect

indoor concentration pollutant levels. In this region kitchens are generally very poorly

ventilated for a number of cultural and lack of awareness reasons, and this is the first

study in Senegal to investigate in detail how and to what extent different household

and ventilation characteristics finally affects IAP.

Findings evidence that, in addition to a switch in the emission source (i.e. cookstove

and/or fuel), successful strategies focused on the improvement of household air

quality in Senegal (and other populations of Sub-Saharan Africa) should improve

ventilation practices, appropriate construction materials and education on the

importance of having a healthy cooking environment.

According to results presented in chapters 4 and 5, rocket stoves would have a positive

effect with regard to IAP reduction, but more uncertain effects on climate, as they

emit more warming (BC) particles. This proves that the climate and health-relevant

properties of stoves do not always scale together and highlights that both dimensions

should be considered in technological decisions (Grieshop et al., 2011).

Page 104: INTO THE SMOKE: A RESEARCH ON HOUSEHOLD AIR ...oa.upm.es/49583/1/CANDELA_DE_LA_SOTA_SANDEZ.pdfCada año, la contaminación del aire interior provoca la muerte prematura de 4,3 millones

88

It is important to note that the observation of no change in fuel use with the rocket

stove was not found in other studies. For example, (Mazorra, 2017) found substantial

fuel savings with the same type of rocket stove in other West-African regions, and this

most likely would imply a reduction of total BC emissions. Moreover, this research

work is focused on black and organic carbon emissions from biomass stoves, which

have an important effect on climate change at both regional and global levels, but it

does not include climate impacts from other pollutants emitted (such as methane,

carbon monoxide, etc.) or factors affecting forestry, such as biomass renewability.

Therefore, results presented in this work do not mean that rocket stove use does not

have climate benefits, as the full picture is much more complicated and uncertain (see

Section 7. Limitations and further research). Rocket stoves, if properly designed, used

and maintained, might produce substantial benefits, such as reduce cooking time,

decrease Products of Incomplete Combustion (PIC’s) production, lower fuel wood

consumption or reduce danger of burns, as previously found in other studies.

Therefore, the types of improved biomass stoves found in this study are not the end

point, but the first step towards the clean cooking.

On this path towards clean and sustainable cooking, findings from this research may be

useful in the development of effective technologies promoting both climate and health

co-benefits and may give useful insights for the selection of more appropriate

technological solutions in each specific context.

However, it should be noted that an effective technology design by itself is not a

solution (Ezzati et al., 2000; Grieshop et al., 2011). Indeed, the high dropout rate found

in the rural village of Senegal were this research study was conducted, together with

the numerous failed cookstoves initiatives around the world, confirm that adoption

and sustained use of stoves by families is plausibly the most important and challenging

consideration.

In addition to provide technical advances, improved stoves have to cover most of the

desirable stove characteristics desired by users, and this is not an easy task. For

instance, gasifiers may fail to cover the household’s preferences and needs as they are

often top-loading only, meaning that the fuel must be fed into the chamber from

above; they require much more frequent re-loading of fuel during cooking sessions;

fan gasifiers may need electricity, which is not available in many regions, and fuelwood

pieces need to be quite small, adding extra work to the cooking-related activities.

Moreover, gasifiers are usually significantly more expensive when compared to other

basic biomass improved stoves.

Therefore, in sub-Saharan Africa, where LPG and other modern fuel cooking appliances

would be out of reach for the majority of population in the medium term, efforts

should be focused on the design of gasifiers and other types of advanced biomass

stoves more affordable and suitable for real cooking practices. Moreover, it is

Page 105: INTO THE SMOKE: A RESEARCH ON HOUSEHOLD AIR ...oa.upm.es/49583/1/CANDELA_DE_LA_SOTA_SANDEZ.pdfCada año, la contaminación del aire interior provoca la muerte prematura de 4,3 millones

89

necessary to continue working to improve the performance (e.g. reduce black carbon

emissions, fuel consumption, etc.) of already available biomass stoves, such as the

rocket stove.

Finally, once technical and usability requirements are achieved, the key point is to

enable an appropriate environment that assures a long-term use and maintenance of

the cooking solutions. If not achieved, all the efforts made to value the potential health

and climate benefits of improved kitchens, such as this thesis, will be futile.

The keys to adoption and sustained use are so complex and broad that they are being

the object of numerous research studies, and are beyond the scope of this thesis.

However, one of the most important that has been traditionally forgotten should be

stressed: If women cannot make independent choices about household resource use,

public policies and other efforts may not be able to promote the cooking technology

adoption (Miller and Mobarak, 2013). A broader social understanding and actions to

fight gender inequalities should accompany every effort of clean cooking promotion.

Page 106: INTO THE SMOKE: A RESEARCH ON HOUSEHOLD AIR ...oa.upm.es/49583/1/CANDELA_DE_LA_SOTA_SANDEZ.pdfCada año, la contaminación del aire interior provoca la muerte prematura de 4,3 millones

90

7. Limitations and further research

Page 107: INTO THE SMOKE: A RESEARCH ON HOUSEHOLD AIR ...oa.upm.es/49583/1/CANDELA_DE_LA_SOTA_SANDEZ.pdfCada año, la contaminación del aire interior provoca la muerte prematura de 4,3 millones

91

While this study provides valuable insights in the field of climate and indoor air

pollution impacts from biomass stoves, a number of limitations should be considered

in interpreting the results. Moreover, several new questions and research needs

emerged during this work, that would need to be covered in further studies. Both

limitations and further research are outlined below:

Given that both laboratory and field testing of stoves are rather costly and time

and resource intensive, this research faced some sample size limitations. In the

laboratory, three replicates of WBT were conducted. However, findings suggested that

in forthcoming emissions testing it would be desirable to include more replicates,

especially for the more emitting stoves, as the rocket stove types. In field studies, the

sample size was rather limited and the variability of results between households was

considerable. Further studies should include, to the possible extent within available

resources, a larger sample size to provide more statistically significant insights. At this

point, it would be very useful to study innovative approaches of data collection in a

way that more information was collected without a high increase of resources.

With regards to methodological limitations, due to very high levels of cookstove

emissions, the filter strip of the portable BC monitor (MicroAeth AE-51, AethLabs) had

to be changed with a high frequency to prevent saturation, complicating the sampling

procedure and resulting in data losses. The monitor used to determine particle number

concentration and mean diameter (DiSCmini, Matter Aerosol) failed throughout the

campaign, also due to the high levels of emissions. Therefore, a use of diluters is

recommended for future studies. Although PM2.5 concentrations were performed with

two devices based in laser photometers previously used in indoor air pollution studies

from cookstoves, further studies should preferably include an in-situ gravimetric

calibration of these devices.

EFs determined in the laboratory vary significantly across type of fuel-stove

combinations tested. Moreover, the field test focused on a single village and only one

type of improved stove, so great caution must be taken before extrapolating the

results to other locations or cooking solutions used in Senegal and other West African

countries. More laboratory and field studies of different cooking technologies and

fuels are needed, covering different areas of the region and studying the seasonality

effect on emissions. This would derive in more representative EFs data to be used in

emission inventories and climate prediction models at regional level.

Further studies to estimate the impacts of cooking activities in West Africa

should include the impact of other health and climate relevant pollutants (e.g. brown

carbon, methane, carbon monoxide, carbon dioxide, polycyclic aromatic

hydrocarbons (PAHs), etc.), as well as other climate-relevant aspects, such as fuelwood

Page 108: INTO THE SMOKE: A RESEARCH ON HOUSEHOLD AIR ...oa.upm.es/49583/1/CANDELA_DE_LA_SOTA_SANDEZ.pdfCada año, la contaminación del aire interior provoca la muerte prematura de 4,3 millones

92

renewability. These studies should be especially focused on biomass-based advanced

technologies and fuels, such as gasifiers and pellets, because they will probably be the

more accessible and affordable clean cooking solutions in sub-Saharan Africa, and to

date very few impact measurements have been carried out on these stoves and fuels.

Future research on cooking testing should involve a larger system, including

technology, fuel, users, cooking environment and the surrounding ambient air. In this

regard, studies should include personal exposure monitoring and consider the stove-

use behaviours (i.e. cooking tasks, work schedules, etc.) Cooking environment

characteristics, such as ventilation, construction materials or stove orientations, should

also be systematically studied to better understand the role they play in stove

emissions production, efficiency, safety parameters and Indoor air pollution. It would

also be very interesting to analyse the contribution of household air pollution to

outdoor (ambient) air pollution, which remains very poorly characterized.

Page 109: INTO THE SMOKE: A RESEARCH ON HOUSEHOLD AIR ...oa.upm.es/49583/1/CANDELA_DE_LA_SOTA_SANDEZ.pdfCada año, la contaminación del aire interior provoca la muerte prematura de 4,3 millones

93

8. References

Page 110: INTO THE SMOKE: A RESEARCH ON HOUSEHOLD AIR ...oa.upm.es/49583/1/CANDELA_DE_LA_SOTA_SANDEZ.pdfCada año, la contaminación del aire interior provoca la muerte prematura de 4,3 millones

94

Adeti, P.J, A., H., 2012. Comparison of Cohen, British and Gagel Method of Approach to the

Determination of Black Carbon in Nucleapore and Teflon Filters Using the M43D

Smokestain Reflectometer. Elixir Pollut. 52, 11258–11260.

Ahmed, T., Dutkiewicz, V.A., Shareef, A., Tuncel, G., Tuncel, S., Husain, L., 2009. Measurement

of black carbon (BC) by an optical method and a thermal-optical method:

Intercomparison for four sites. Atmos. Environ. 43, 6305–6311.

doi:10.1016/j.atmosenv.2009.09.031

Akagi, S.K., Yokelson, R.J., Wiedinmyer, C., Alvarado, M.J., Reid, J.S., Karl, T., Crounse, J.D.,

Wennberg, P.O., 2010. Emission factors for open and domestic biomass burning for

use in atmospheric models. Atmospheric Chem. Phys. Discuss. 10, 27523–27602.

doi:10.5194/acpd-10-27523-2010

Albalak, R., Bruce, N., McCracken, J.P., Smith, K.R., de Gallardo, T., 2001. Indoor Respirable

Particulate Matter Concentrations from an Open Fire, Improved Cookstove, and

LPG/Open Fire Combination in a Rural Guatemalan Community. Environ. Sci. Technol.

35, 2650–2655. doi:10.1021/es001940m

Andreae, M.O., Gelencsér, A., 2006. Black carbon or brown carbon? The nature of light-

absorbing carbonaceous aerosols. Atmospheric Chem. Phys. 6, 3131–3148.

Anenberg, S.C., Balakrishnan, K., Jetter, J., Masera, O., Mehta, S., Moss, J., Ramanathan, V.,

2013. Cleaner Cooking Solutions to Achieve Health, Climate, and Economic Cobenefits.

Environ. Sci. Technol. 47, 3944–3952. doi:10.1021/es304942e

Armendáriz, C., Edwards, R.D., Johnson, M., Rosas, I.A., Espinosa, F., Masera, O.R., 2010.

Indoor particle size distributions in homes with open fires and improved Patsari cook

stoves☆. Atmos. Environ. 44, 2881–2886. doi:10.1016/j.atmosenv.2010.04.049

Armendáriz, C., Edwards, R.D., Johnson, M., Zuk, M., Rojas, L., Jiménez, R.D., Riojas-Rodriguez,

H., Masera, O., 2008. Reduction in personal exposures to particulate matter and

carbon monoxide as a result of the installation of a Patsari improved cook stove in

Michoacan Mexico. Indoor Air 18, 93–105. doi:10.1111/j.1600-0668.2007.00509.x

Arora, P., Jain, S., 2015. Estimation of Organic and Elemental Carbon Emitted from Wood

Burning in Traditional and Improved Cookstoves Using Controlled Cooking Test.

Environ. Sci. Technol. 49, 3958–3965. doi:10.1021/es504012v

Arora, P., Jain, S., Sachdeva, K., 2013. Physical characterization of particulate matter emitted

from wood combustion in improved and traditional cookstoves. Energy Sustain. Dev.

17, 497–503. doi:10.1016/j.esd.2013.06.003

Asbach, C., Alexander, C., Clavaguera, S., Dahmann, D., Dozol, H., Faure, B., Fierz, M., Fontana,

L., Iavicoli, I., Kaminski, H., MacCalman, L., Meyer-Plath, A., Simonow, B., van

Tongeren, M., Todea, A.M., 2017. Review of measurement techniques and methods

for assessing personal exposure to airborne nanomaterials in workplaces. Sci. Total

Environ. doi:10.1016/j.scitotenv.2017.03.049

Atiku, F.A., Mitchell, E.J.S., Lea-Langton, A.R., Jones, J.M., Williams, A., Bartle, K.D., 2016. The

Impact of Fuel Properties on the Composition of Soot Produced by the Combustion of

Residential Solid Fuels in a Domestic Stove. Fuel Process. Technol. 151, 117–125.

doi:10.1016/j.fuproc.2016.05.032

Austin, K.F., Mejia, M.T., 2017. Household air pollution as a silent killer: women’s status and

solid fuel use in developing nations. Popul. Environ. doi:10.1007/s11111-017-0269-z

Page 111: INTO THE SMOKE: A RESEARCH ON HOUSEHOLD AIR ...oa.upm.es/49583/1/CANDELA_DE_LA_SOTA_SANDEZ.pdfCada año, la contaminación del aire interior provoca la muerte prematura de 4,3 millones

95

Auth, K., Musolino, E., 2014. ECOWAS RENEWABLE ENERGY AND ENERGY EFFICIENCY STATUS

REPORT. ECREEE, Paris, France.

Bailis, R., Drigo, R., Ghilardi, A., Masera, O., 2015. The carbon footprint of traditional

woodfuels. Nat. Clim. Change 5, 266–272. doi:10.1038/nclimate2491

Balakrishnan, K., Sambandam, S., Ghosh, S., Mukhopadhyay, K., Vaswani, M., Arora, N.K., Jack,

D., Pillariseti, A., Bates, M.N., Smith, K.R., 2015. Household Air Pollution Exposures of

Pregnant Women Receiving Advanced Combustion Cookstoves in India: Implications

for Intervention. Ann. Glob. Health 81, 375–385. doi:10.1016/j.aogh.2015.08.009

Bartington, S.E., Bakolis, I., Devakumar, D., Kurmi, O.P., Gulliver, J., Chaube, G., Manandhar,

D.S., Saville, N.M., Costello, A., Osrin, D., Hansell, A.L., Ayres, J.G., 2017. Patterns of

domestic exposure to carbon monoxide and particulate matter in households using

biomass fuel in Janakpur, Nepal. Environ. Pollut. 220, 38–45.

doi:10.1016/j.envpol.2016.08.074

Batliwala, S., Reddy, A.K., 2003. Energy for women and women for energy (engendering energy

and empowering women) 11A rudimentary version of this paper was presented at the

Brainstorming Meeting of ENERGIA: Women and Energy Network on June 4-5, 1996, at

the University of Twente, Enschede, the Netherlands (cf.[Batliwala and Reddy, 1996]).

Energy Sustain. Dev. 7, 33–43.

Baumgartner, J., Schauer, J.J., Ezzati, M., Lu, L., Cheng, C., Patz, J., Bautista, L.E., 2011. Patterns

and predictors of personal exposure to indoor air pollution from biomass combustion

among women and children in rural China: Biomass smoke exposure in rural China.

Indoor Air 21, 479–488. doi:10.1111/j.1600-0668.2011.00730.x

Baumgartner, J., Zhang, Y., Schauer, J.J., Huang, W., Wang, Y., Ezzati, M., 2014. Highway

proximity and black carbon from cookstoves as a risk factor for higher blood pressure

in rural China. Proc. Natl. Acad. Sci. 111, 13229–13234. doi:10.1073/pnas.1317176111

Begum, B.A., Biswas, S.K., Hopke, P.K., 2007. Source apportionment of air particulate matter by

chemical mass balance (CMB) and comparison with positive matrix factorization (PMF)

model. Aerosol Air Qual. Res. 7, 446–468.

Bhattacharya, S.C., Albina, D.O., Salam, P.A., 2002. Emission factors of wood and charcoal-_red

cookstoves. Biomass Bioenergy 23, 453 – 469.

Biswas, S.K., Tarafdar, S.A., Islam, A., Khaliquzzaman, M., Tervahattu, H., Kupiainen, K., 2003.

Impact of Unleaded Gasoline Introduction on the Concentration of Lead in the Air of

Dhaka, Bangladesh. J. Air Waste Manag. Assoc. 53, 1355–1362.

doi:10.1080/10473289.2003.10466299

Bond, T.C., 2004. A technology-based global inventory of black and organic carbon emissions

from combustion. J. Geophys. Res. 109. doi:10.1029/2003JD003697

Bond, T.C., Anderson, T.L., Campbell, D., 1999. Calibration and Intercomparison of Filter-Based

Measurements of Visible Light Absorption by Aerosols. Aerosol Sci. Technol. 30, 582–

600. doi:10.1080/027868299304435

Bond, T.C., Bergstrom, R.W., 2006. Light Absorption by Carbonaceous Particles: An

Investigative Review. Aerosol Sci. Technol. 40, 27–67.

doi:10.1080/02786820500421521

Bond, T.C., Bhardwaj, E., Dong, R., Jogani, R., Jung, S., Roden, C., Streets, D.G., Trautmann,

N.M., 2007. Historical emissions of black and organic carbon aerosol from energy-

Page 112: INTO THE SMOKE: A RESEARCH ON HOUSEHOLD AIR ...oa.upm.es/49583/1/CANDELA_DE_LA_SOTA_SANDEZ.pdfCada año, la contaminación del aire interior provoca la muerte prematura de 4,3 millones

96

related combustion, 1850-2000: HISTORICAL BC/OC EMISSIONS. Glob. Biogeochem.

Cycles 21, n/a-n/a. doi:10.1029/2006GB002840

Bond, T.C., Doherty, S.J., Fahey, D.W., Forster, P.M., Berntsen, T., DeAngelo, B.J., Flanner, M.G.,

Ghan, S., Kärcher, B., Koch, D., Kinne, S., Kondo, Y., Quinn, P.K., Sarofim, M.C., Schultz,

M.G., Schulz, M., Venkataraman, C., Zhang, H., Zhang, S., Bellouin, N., Guttikunda, S.K.,

Hopke, P.K., Jacobson, M.Z., Kaiser, J.W., Klimont, Z., Lohmann, U., Schwarz, J.P.,

Shindell, D., Storelvmo, T., Warren, S.G., Zender, C.S., 2013. Bounding the role of black

carbon in the climate system: A scientific assessment: BLACK CARBON IN THE CLIMATE

SYSTEM. J. Geophys. Res. Atmospheres 118, 5380–5552. doi:10.1002/jgrd.50171

Bond, T.C., Zarzycki, C., Flanner, M.G., Koch, D.M., 2011. Quantifying immediate radiative

forcing by black carbon and organic matter with the Specific Forcing Pulse.

Atmospheric Chem. Phys. 11, 1505–1525. doi:10.5194/acp-11-1505-2011

Bruce, N., McCracken, J., Albalak, R., Schei, M.A., Smith, K.R., Lopez, V., West, C., 2004. Impact

of improved stoves, house construction and child location on levels of indoor air

pollution exposure in young Guatemalan children. J. Expo. Anal. Environ. Epidemiol.

14, S26–S33. doi:10.1038/sj.jea.7500355

Bruce, N., Smith, K.R., et al., 2014. WHO Indoor Air Quality Guidelines: Household Fuel

Combustion Review 4: Health effects of household air pollution (HAP) exposure. WHO,

Geneva.

Buonanno, G., Marini, S., Morawska, L., Fuoco, F.C., 2012. Individual dose and exposure of

Italian children to ultrafine particles. Sci. Total Environ. 438, 271–277.

doi:10.1016/j.scitotenv.2012.08.074

Cachier, H., Liousse, C., Pertuisot, M.H., Gaudichet, A., Echalar, F., Lacaux, J.P., 1996. African

fire particulate emissions and atmospheric influence. Biomass Burn. Glob. Change 1,

428–440.

Cao, G., Zhang, X., Zheng, F., 2006. Inventory of black carbon and organic carbon emissions

from China. Atmos. Environ. 40, 6516–6527. doi:10.1016/j.atmosenv.2006.05.070

Castell, N., Kobernus, M., Liu, H.-Y., Schneider, P., Lahoz, W., Berre, A.J., Noll, J., 2015. Mobile

technologies and services for environmental monitoring: The Citi-Sense-MOB

approach. Urban Clim. 14, Part 3, 370–382.

doi:http://dx.doi.org/10.1016/j.uclim.2014.08.002

Cavalli, F., Viana, M., Yttri, K.E., Genberg, J., Putaud, P., 2010. Toward a standardised thermal-

optical protocol for measuring atmospheric organic and elemental carbon: the EUSAAR

protocol. Atmospheric Meas. Tech. 3, 79–89. doi:10.5194/amt-3-79-2010

CCAC, 2014. Time To Act to reduce Short-Lived Climate Pollutants.pdf. UNEP, Paris, France.

Chen, Y., Roden, C.A., Bond, T.C., 2012. Characterizing Biofuel Combustion with Patterns of

Real-Time Emission Data (PaRTED). Environ. Sci. Technol. 46, 6110–6117.

doi:10.1021/es3003348

Chen, Y., Sheng, G., Bi, X., Feng, Y., Mai, B., Fu, J., 2005. Emission Factors for Carbonaceous

Particles and Polycyclic Aromatic Hydrocarbons from Residential Coal Combustion in

China. Environ. Sci. Technol. 39, 1861–1867. doi:10.1021/es0493650

Chengappa, C., Edwards, R., Bajpai, R., Shields, K.N., Smith, K.R., 2007. Impact of improved

cookstoves on indoor air quality in the Bundelkhand region in India. Energy Sustain.

Dev. 11, 33–44.

Page 113: INTO THE SMOKE: A RESEARCH ON HOUSEHOLD AIR ...oa.upm.es/49583/1/CANDELA_DE_LA_SOTA_SANDEZ.pdfCada año, la contaminación del aire interior provoca la muerte prematura de 4,3 millones

97

Chow, J.C., Watson, J.G., Chen, L.-W.A., Arnott, W.P., Moosmüller, H., 2004. Equivalence of

elemental carbon by thermal/optical reflectance and transmittance with different

temperature protocols. Environ. Sci. Technol. 38, 4414–4422. doi:10.1021/es034936u

Chowdhury, Z., 2012. Quantification of Indoor Air Pollution from Using Cookstoves and

Estimation of Its Health Effects on Adult Women in Northwest Bangladesh. Aerosol Air

Qual. Res. 12, 463–475. doi:10.4209/aaqr.2011.10.0161

Chowdhury, Z., Campanella, L., Gray, C., Al Masud, A., Marter-Kenyon, J., Pennise, D., Charron,

D., Zuzhang, X., 2013. Measurement and modeling of indoor air pollution in rural

households with multiple stove interventions in Yunnan, China. Atmos. Environ. 67,

161–169. doi:10.1016/j.atmosenv.2012.10.041

Chowdhury, Z., Edwards, R.D., Johnson, M., Naumoff Shields, K., Allen, T., Canuz, E., Smith,

K.R., 2007. An inexpensive light-scattering particle monitor: field validation. J. Environ.

Monit. 9, 1099. doi:10.1039/b709329m

Chung, C.E., Ramanahatan, V., Decremer, D., 2012. Observationally-constrained Estimates of

carbonaceous aerosol radiative forcing.

Chung, S.H., 2002. Global distribution and climate forcing of carbonaceous aerosols. J.

Geophys. Res. 107. doi:10.1029/2001JD001397

Clark, M.L., Peel, J.L., Balakrishnan, K., Breysse, P.N., Chillrud, S.N., Naeher, L.P., Rodes, C.E.,

Vette, A.F., Balbus, J.M., 2013. Health and Household Air Pollution from Solid Fuel Use:

The Need for Improved Exposure Assessment. Environ. Health Perspect. 121.

doi:10.1289/ehp.1206429

Commodore, A.A., Hartinger, S.M., Lanata, C.F., Mäusezahl, D., Gil, A.I., Hall, D.B., Aguilar-

Villalobos, M., Naeher, L.P., 2013. A pilot study characterizing real time exposures to

particulate matter and carbon monoxide from cookstove related woodsmoke in rural

Peru. Atmos. Environ. 79, 380–384. doi:10.1016/j.atmosenv.2013.06.047

Dafrallah, T., 2009. Energy Security in West Africa The Case of Senegal-Final report.

Dasgupta, S., Huq, M., Khaliquzzaman, M., Pandey, K., Wheeler, D., 2006. Indoor air quality for

poor families: new evidence from Bangladesh. Indoor Air 16, 426–444.

doi:10.1111/j.1600-0668.2006.00436.x

Davy, P.M., Tremper, A.H., Nicolosi, E.M.G., Quincey, P., Fuller, G.W., 2017. Estimating

particulate black carbon concentrations using two offline light absorption methods

applied to four types of filter media. Atmos. Environ. 152, 24–33.

doi:10.1016/j.atmosenv.2016.12.010

de la Sota, C., Kane, M., Mazorra, J., Lumbreras, J., Youm, I., Viana, M., 2017. Intercomparison

of methods to estimate black carbon emissions from cookstoves. Sci. Total Environ.

595, 886–893. doi:10.1016/j.scitotenv.2017.03.247

Desai, M.A., Mehta, S., Smith, K.R., World Health Organization, Protection of the Human

Environment, 2004. Indoor smoke from solid fuels: assessing the environmental

burden of disease at national and local levels. Protection of the Human Environment,

World Health Organization, Geneva [Switzerland.

Dherani, M., Pope, D., Mascarenhas, M., Smith, K., Weber, M., Bruce, N., 2008. Indoor air

pollution from unprocessed solid fuel use and pneumonia risk in children aged under

five years: a systematic review and meta-analysis. Bull. World Health Organ. 86, 390–

398. doi:10.2471/BLT.07.044529

Page 114: INTO THE SMOKE: A RESEARCH ON HOUSEHOLD AIR ...oa.upm.es/49583/1/CANDELA_DE_LA_SOTA_SANDEZ.pdfCada año, la contaminación del aire interior provoca la muerte prematura de 4,3 millones

98

Downward, G.S., Hu, W., Rothman, N., Reiss, B., Wu, G., Wei, F., Xu, J., Seow, W.J., Brunekreef,

B., Chapman, R.S., Qing, L., Vermeulen, R., 2015. Outdoor, indoor, and personal black

carbon exposure from cookstoves burning solid fuels. Indoor Air.

doi:10.1111/ina.12255

Dutta, K., Shields, K.N., Edwards, R., Smith, K.R., 2007. Impact of improved biomass cookstoves

on indoor air quality near Pune, India. Energy Sustain. Dev. 11, 19–32.

Edwards, J.D., Ogren, J.A., Weiss, R.E., Charlson, R.J., 1983. Particulate air pollutants: A

comparison of British “Smoke” with optical absorption coefficient and elemental

carbon concentration. Atmos. Environ. 17. doi:10.1016/0004-6981(83)90233-0

Edwards, R., Hubbard, A., Khalakdina, A., Pennise, D., Smith, K.R., 2007. Design considerations

for field studies of changes in indoor air pollution due to improved stoves. Energy

Sustain. Dev. 11, 71–81.

EPA, 2012. Report to Congress on Black Carbon (No. Publication No. EPA-450/R-12-001).

United States Environmental Protection Agency. Department of the Interior,

Environment, and Related Agencies Appropriations Act.

EPA, PCIA, GACC, 2014. The Water Boiling Test. Version 4.2.3. Cookstove Emissions and

Efficiency in a Controlled Laboratory Setting. U.S. Environmental Protection Agency,

Partnership for Clean Indoor Air, Global Alliance for Clean Cookstoves.

Ezzati, M., Mbinda, B.M., Kammen, D.M., 2000. Comparison of Emissions and Residential

Exposure from Traditional and Improved Cookstoves in Kenya. Environ. Sci. Technol.

34, 578–583. doi:10.1021/es9905795

Ezzati, M., World Health Organization, 2004. Comparative quantification of health risks: global

and regional burden of disease attributable to selected major risk factors. World

Health Organization, Geneva.

Fan, C.-W., Zhang, J., 2001. Characterization of emissions from portable household combustion

devices: particle size distributions, emission rates and factors, and potential exposures.

Atmos. Environ. 35, 1281–1290. doi:10.1016/S1352-2310(00)00399-X

Feng, Y., Ramanathan, V., Kotamarthi, V.R., 2013. Brown carbon: a significant atmospheric

absorber of solar radiation? Atmospheric Chem. Phys. 13, 8607–8621.

doi:10.5194/acp-13-8607-2013

Fierz, M., Houle, C., Steigmeier, P., Burtscher, H., 2011. Design, Calibration, and Field

Performance of a Miniature Diffusion Size Classifier. Aerosol Sci. Technol. 45, 1–10.

doi:10.1080/02786826.2010.516283

Fitzgerald, C., Aguilar-Villalobos, M., Eppler, A.R., Dorner, S.C., Rathbun, S.L., Naeher, L.P.,

2012. Testing the effectiveness of two improved cookstove interventions in the

Santiago de Chuco Province of Peru. Sci. Total Environ. 420, 54–64.

doi:10.1016/j.scitotenv.2011.10.059

Foell, W., Pachauri, S., Spreng, D., Zerriffi, H., 2011. Household cooking fuels and technologies

in developing economies. Energy Policy 39, 7487–7496.

doi:10.1016/j.enpol.2011.08.016

GACC, 2014. Results Report: Sharing Progress on the Path to Adoption of Cleaner and More

Efficient Cooking Solutions. Global Alliance for Clean Cookstoves, Washington, DC,

USA.

GACC, 2012. Global alliance for Cleancookstoves. Technology and fuels. Standards. [WWW

Document]. URL http://cleancookstoves.org/technology-and-fuels/standards/

Page 115: INTO THE SMOKE: A RESEARCH ON HOUSEHOLD AIR ...oa.upm.es/49583/1/CANDELA_DE_LA_SOTA_SANDEZ.pdfCada año, la contaminación del aire interior provoca la muerte prematura de 4,3 millones

99

Gao, X., Yu, Q., Gu, Q., Chen, Y., Ding, K., Zhu, J., Chen, L., 2009. Indoor air pollution from solid

biomass fuels combustion in rural agricultural area of Tibet, China. Indoor Air 19, 198–

205. doi:10.1111/j.1600-0668.2008.00579.x

Gorritty, M., Torrico, T., Montenegro, Y., 2011. Determinación de la tasa de emisión de CO en

cocinas mejoradas a leña con chimenea mediante el modelo de caja con ventilación

constante. Acta Nova 5, 96.

Grabow, K., Still, D., Bentson, S., 2013. Test Kitchen studies of indoor air pollution from

biomass cookstoves. Energy Sustain. Dev. 17, 458–462. doi:10.1016/j.esd.2013.05.003

Grieshop, A.P., Marshall, J.D., Kandlikar, M., 2011. Health and climate benefits of cookstove

replacement options. Energy Policy 39, 7530–7542. doi:10.1016/j.enpol.2011.03.024

Guofeng, S., Siye, W., Wen, W., Yanyan, Z., Yujia, M., Bin, W., Rong, W., Wei, L., Huizhong, S.,

Ye, H., Yifeng, Y., Wei, W., Xilong, W., Xuejun, W., Shu, T., 2012. Emission Factors, Size

Distributions, and Emission Inventories of Carbonaceous Particulate Matter from

Residential Wood Combustion in Rural China. Environ. Sci. Technol. 46, 4207–4214.

doi:10.1021/es203957u

Habib, G., Venkataraman, C., Shrivastava, M., Banerjee, R., Stehr, J.W., Dickerson, R.R., 2004.

New methodology for estimating biofuel consumption for cooking: Atmospheric

emissions of black carbon and sulfur dioxide from India: BIOFUEL COMBUSTION AND

ATMOSPHERIC EMISSIONS. Glob. Biogeochem. Cycles 18, n/a-n/a.

doi:10.1029/2003GB002157

Hamrick, K., 2015. Emerging From The Darkness: New Process Aims To Tackle Black Carbon.

Ecosyst. Marketpl. For. Trend Initiat.

Hinds, W.C., 1999. Aerosol Technology: Properties, Behavior, and Measurement of Airborne

Particles, JOHN WILEY & SONS, INC. ed. New York.

Hosgood, H.D., Lan, Q., Vermeulen, R., Wei, H., Reiss, B., Coble, J., Wei, F., Jun, X., Wu, G.,

Rothman, N., 2012. Combustion-derived nanoparticle exposure and household solid

fuel use in Xuanwei and Fuyuan, China. Int. J. Environ. Health Res. 22, 571–581.

doi:10.1080/09603123.2012.684147

Hu, W., Downward, G.S., Reiss, B., Xu, J., Bassig, B.A., Hosgood, H.D., Zhang, L., Seow, W.J., Wu,

G., Chapman, R.S., Tian, L., Wei, F., Vermeulen, R., Lan, Q., 2014. Personal and Indoor

PM 2.5 Exposure from Burning Solid Fuels in Vented and Unvented Stoves in a Rural

Region of China with a High Incidence of Lung Cancer. Environ. Sci. Technol. 48, 8456–

8464. doi:10.1021/es502201s

Huboyo, H.S., Budihardjo, A., Hardyanti, N., 2009. Black carbon concentration in kitchens using

fire-wood and kerosene fuels. J Appl Sci Env. Sanit 4, 55–62.

Ichoku, C., Ellison, L.T., Willmot, K.E., Matsui, T., Dezfuli, A.K., Gatebe, C.K., Wang, J., Wilcox,

E.M., Lee, J., Adegoke, J., Okonkwo, C., Bolten, J., Policelli, F.S., Habib, S., 2016.

Biomass burning, land-cover change, and the hydrological cycle in Northern sub-

Saharan Africa. Environ. Res. Lett. 11, 095005. doi:10.1088/1748-9326/11/9/095005

IEA, 2016. World Energy Outlook Special Report 2016: Energy and Air Pollution. International

Energy Agency, Paris, France.

IEA, 2014. Africa Energy Outlook. A focus on energy propects in Sub-Saharan Africa (World

Energy Outlook Special Report). Paris, France.

IEA, 2006. Chapter 15. Energy for Cooking in Developing Countries, in: World Energy Outlook

2006. WEO,IEA, pp. 419–446.

Page 116: INTO THE SMOKE: A RESEARCH ON HOUSEHOLD AIR ...oa.upm.es/49583/1/CANDELA_DE_LA_SOTA_SANDEZ.pdfCada año, la contaminación del aire interior provoca la muerte prematura de 4,3 millones

100

Iyengar, K., 2015. Sustainable Architectural Design: An Overview. Routledge.

Jacobson, M.Z., 2000. Strong radiative heating due to the mixing state of black carbon in

atmospheric aerosols. Nature 409, 695–697. doi:10.1038/35055518

Janssen, N., Gerlofs-Nijland, M., Lanki, T., Salonen, R., Cassee, F., Hoek, G., Fischer, P.,

Brunekreef, B., Krzyzanowski, M. (Eds.), 2012. Health effects of black carbon. World

Health Organization, Regional Office for Europe, Copenhagen.

Janssen, N.A., Hoek, G., Simic-Lawson, M., Fischer, P., Van Bree, L., Ten Brink, H., Keuken, M.,

Atkinson, R.W., Anderson, H.R., Brunekreef, B., others, 2011. Black carbon as an

additional indicator of the adverse health effects of airborne particles compared with

PM10 and PM2. 5. Environ. Health Perspect. 119, 1691.

Jeong, C.-H., Hopke, P.K., Kim, E., Lee, D.-W., 2004. The comparison between thermal-optical

transmittance elemental carbon and Aethalometer black carbon measured at multiple

monitoring sites. Atmos. Environ. 38, 5193–5204. doi:10.1016/j.atmosenv.2004.02.065

Jetter, J., 2015. Black Carbon and Short- Lived Climate Pollutants Q&A. Global Alliance for

Cleancookstoves.

Jetter, J., Zhao, Y., Smith, K.R., Khan, B., Yelverton, T., DeCarlo, P., Hays, M.D., 2012. Pollutant

Emissions and Energy Efficiency under Controlled Conditions for Household Biomass

Cookstoves and Implications for Metrics Useful in Setting International Test Standards.

Environ. Sci. Technol. 46, 10827–10834. doi:10.1021/es301693f

Jetter, J.J., Kariher, P., 2009. Solid-fuel household cook stoves: Characterization of

performance and emissions. Biomass Bioenergy 33, 294–305.

doi:10.1016/j.biombioe.2008.05.014

Johnson, M., Edwards, R., Alatorre Frenk, C., Masera, O., 2008a. In-field greenhouse gas

emissions from cookstoves in rural Mexican households. Atmos. Environ. 42, 1206–

1222. doi:10.1016/j.atmosenv.2007.10.034

Johnson, M., Edwards, R., Alatorre Frenk, C., Masera, O., 2008b. In-field greenhouse gas

emissions from cookstoves in rural Mexican households. Atmos. Environ. 42, 1206–

1222. doi:10.1016/j.atmosenv.2007.10.034

Just, B., Rogak, S., Kandlikar, M., 2013. Characterization of Ultrafine Particulate Matter from

Traditional and Improved Biomass Cookstoves. Environ. Sci. Technol. 47, 3506–3512.

doi:10.1021/es304351p

Kar, A., Rehman, I.H., Burney, J., Puppala, S.P., Suresh, R., Singh, L., Singh, V.K., Ahmed, T.,

Ramanathan, N., Ramanathan, V., 2012. Real-Time Assessment of Black Carbon

Pollution in Indian Households Due to Traditional and Improved Biomass Cookstoves.

Environ. Sci. Technol. 46, 2993–3000. doi:10.1021/es203388g

Kaufman, A., Williams, R., Barzyk, T., Hoang, A., Sheridan, P., 2014. The Air Sensor Citizen

Science Toolbox: A Collaboration in Community Air Quality Monitoring and Mapping,

in: Second AAAI Conference on Human Computation and Crowdsourcing.

Kinney, P.L., 2015. Black Carbon: Perpetrator or indicator?

Ko, Y.-C., Cheng, L.S.-C., Lee, C.-H., Huang, J.-J., Huang, M.-S., Kao, E.-L., Wang, H.-Z., Lin, H.-J.,

2000. Chinese food cooking and lung cancer in women nonsmokers. Am. J. Epidemiol.

151, 140–147.

Koehler, K.A., Peters, T.M., 2015. New Methods for Personal Exposure Monitoring for Airborne

Particles. Curr. Environ. Health Rep. 2, 399–411. doi:10.1007/s40572-015-0070-z

Page 117: INTO THE SMOKE: A RESEARCH ON HOUSEHOLD AIR ...oa.upm.es/49583/1/CANDELA_DE_LA_SOTA_SANDEZ.pdfCada año, la contaminación del aire interior provoca la muerte prematura de 4,3 millones

101

Lacey, F.G., Henze, D.K., Lee, C.J., van Donkelaar, A., Martin, R.V., 2017. Transient climate and

ambient health impacts due to national solid fuel cookstove emissions. Proc. Natl.

Acad. Sci. 114, 1269–1274. doi:10.1073/pnas.1612430114

Lambe, F., Jürisoo, M., Lee, C., Johnson, O., 2015a. Can carbon finance transform household

energy markets? A review of cookstove projects and programs in Kenya. Energy Res.

Soc. Sci. 5, 55–66.

Lambe, F., Jürisoo, M., Wanjiru, H., Senyagwa, J., 2015b. Bringing clean, safe, affordable

cooking energy to households across Africa: an agenda for action. Prep. Stockh.

Environ. Inst. Stockh. Nairobi New Clim. Econ. Available Httpnewclimateeconomy

Reportmiscworkingpapers.

Leavey, A., Londeree, J., Priyadarshini, P., Puppala, J., Schechtman, K.B., Yadama, G., Biswas, P.,

2015. Real-Time Particulate and CO Concentrations from Cookstoves in Rural

Households in Udaipur, India. Environ. Sci. Technol. 49, 7423–7431.

doi:10.1021/acs.est.5b02139

Lee, C.M., Chandler, C., 2013. Assessing the climate impacts of cookstove projects: issues in

emissions accounting. Chall. Sustain. 1, 53.

Li, C., Bosch, C., Kang, S., Andersson, A., Chen, P., Zhang, Q., Cong, Z., Chen, B., Qin, D.,

Gustafsson, Ö., 2016. Sources of black carbon to the Himalayan–Tibetan Plateau

glaciers. Nat. Commun. 7, 12574. doi:10.1038/ncomms12574

Li, X., Wang, S., Duan, L., Hao, J., Nie, Y., 2009. Carbonaceous Aerosol Emissions from

Household Biofuel Combustion in China. Environ. Sci. Technol. 43, 6076–6081.

doi:10.1021/es803330j

Liyama, M., 2013. Charcoal: A driver of dryland forest degradation in Africa? (Fact sheet).

World Agroforestry Centre, Nairobi, Kenya.

L’Orange, C., DeFoort, M., Willson, B., 2012. Influence of testing parameters on biomass stove

performance and development of an improved testing protocol. Energy Sustain. Dev.

16, 3–12. doi:10.1016/j.esd.2011.10.008

MacCarty, N., Ogle, D., Still, D., Bond, T., 2008a. A laboratory comparison of the global

warming impact of five major types of biomass cooking stoves. Energy Sustain. Dev.

12, 56–65.

MacCarty, N., Ogle, D., Still, D., Bond, T., Roden, C., 2008b. A laboratory comparison of the

global warming impact of five major types of biomass cooking stoves. Energy Sustain.

Dev. 12, 56–65. doi:10.1016/S0973-0826(08)60429-9

MacCarty, N., Ogle, D., Still, D., Bond, T., Roden, C., Willson, B., 2007a. Laboratory comparison

of the global-warming potential of six categories of biomass cooking stoves. Creswell

Aprovecho Res. Cent.

MacCarty, N., Ogle, D., Still, D., Bond, T., Roden, C., Willson, B., 2007b. Laboratory comparison

of the global-warming potential of six categories of biomass cooking stoves. Creswell

Aprovecho Res. Cent.

MacCarty, N., Still, D., Ogle, D., 2010. Fuel use and emissions performance of fifty cooking

stoves in the laboratory and related benchmarks of performance. Energy Sustain. Dev.

14, 161–171. doi:10.1016/j.esd.2010.06.002

Mahmood, A., 1993. Suitability of casuarina equisetifolia wood for pulp and paper industry in

Pakistan. Pak. J. Bot. 25, 179–182.

Page 118: INTO THE SMOKE: A RESEARCH ON HOUSEHOLD AIR ...oa.upm.es/49583/1/CANDELA_DE_LA_SOTA_SANDEZ.pdfCada año, la contaminación del aire interior provoca la muerte prematura de 4,3 millones

102

Martin, W.J., Hollingsworth, J.W., Ramanathan, V., 2014. Household Air Pollution from

Cookstoves: Impacts on Health and Climate, in: Pinkerton, K.E., Rom, W.N. (Eds.),

Global Climate Change and Public Health. Springer New York, New York, NY, pp. 237–

255. doi:10.1007/978-1-4614-8417-2_13

Masera, O.R., Saatkamp, B.D., Kammen, D.M., 2000. From linear fuel switching to multiple

cooking strategies: a critique and alternative to the energy ladder model. World Dev.

28, 2083–2103.

Mazorra, J., 2017. Vínculos entre acceso a energía, cambio climático y género en países en

desarrollo: Una aproximación a través de las cocinas mejoradas. Universidad

Politécnica de Madrid, Madrid.

Miller, G., Mobarak, A.M., 2013. Gender differences in preferences, intra-household

externalities, and low demand for improved cookstoves. National Bureau of Economic

Research.

Mitchell, E.J.S., Lea-Langton, A.R., Jones, J.M., Williams, A., Layden, P., Johnson, R., 2016. The

impact of fuel properties on the emissions from the combustion of biomass and other

solid fuels in a fixed bed domestic stove. Fuel Process. Technol. 142, 115–123.

doi:10.1016/j.fuproc.2015.09.031

Naeher, L.P., Leaderer, B.P., Smith, K.R., 2000. Particulate matter and carbon monoxide in

highland Guatemala: indoor and outdoor levels from traditional and improved wood

stoves and gas stoves. Indoor Air 10, 200–205.

Ndiaye, O., Komivi, B.S., Diei-Ouadi, Y., Food and Agriculture Organization of the United

Nations, 2015. Guide for developing and using the FAO-Thiaroye processing technique

(FTT-Thiaroye).

Nelms, S., Coombes, C., Foster, L., Galloway, T., Godley, B., Lindeque, P., Witt, M., 2016.

Marine anthropogenic litter on British beaches: A 10-year nationwide assessment

using citizen science data. Sci. Total Environ. doi:10.1016/j.scitotenv.2016.11.137

Northcross, A., Chowdhury, Z., McCracken, J., Canuz, E., Smith, K.R., 2010. Estimating personal

PM2.5 exposures using CO measurements in Guatemalan households cooking with

wood fuel. J. Environ. Monit. 12, 873. doi:10.1039/b916068j

Novakov, T., Corrigan, C., 1995. Influence of Sample Composition on Aerosol Organic and Black

Carbon Determinations (No. LBL-37153). California.

Ntziachristos, L., Polidori, A., Phuleria, H., Geller, M.D., Sioutas, C., 2007. Application of a

Diffusion Charger for the Measurement of Particle Surface Concentration in Different

Environments. Aerosol Sci. Technol. 41, 571–580. doi:10.1080/02786820701272020

Ochieng, C., Vardoulakis, S., Tonne, C., 2017. Household air pollution following replacement of

traditional open fire with an improved rocket type cookstove. Sci. Total Environ. 580,

440–447. doi:10.1016/j.scitotenv.2016.10.233

Ochieng, C.A., Vardoulakis, S., Tonne, C., 2013. Are rocket mud stoves associated with lower

indoor carbon monoxide and personal exposure in rural Kenya?: Kitchen and personal

CO from traditional and rocket stoves. Indoor Air 23, 14–24. doi:10.1111/j.1600-

0668.2012.00786.x

OECD/IEA, 2014. Africa Energy Outlook.A Focus on energy prospects in Sub-Saharan Africa.

(World Energy Outlook Special Report). International Energy Agency, Paris, France.

Park, E., Lee, K., 2003. Particulate exposure and size distribution from wood burning stoves in

Costa Rica. Indoor Air 13, 253–259.

Page 119: INTO THE SMOKE: A RESEARCH ON HOUSEHOLD AIR ...oa.upm.es/49583/1/CANDELA_DE_LA_SOTA_SANDEZ.pdfCada año, la contaminación del aire interior provoca la muerte prematura de 4,3 millones

103

Patange, O.S., Ramanathan, N., Rehman, I.H., Tripathi, S.N., Misra, A., Kar, A., Graham, E.,

Singh, L., Bahadur, R., Ramanathan, V., 2015. Reductions in Indoor Black Carbon

Concentrations from Improved Biomass Stoves in Rural India. Environ. Sci. Technol. 49,

4749–4756. doi:10.1021/es506208x

Patel, S., Leavey, A., He, S., Fang, J., O’Malley, K., Biswas, P., 2016. Characterization of gaseous

and particulate pollutants from gasification-based improved cookstoves. Energy

Sustain. Dev. 32, 130–139. doi:10.1016/j.esd.2016.02.005

Pennise, D., Brant, S., Agbeve, S.M., Quaye, W., Mengesha, F., Tadele, W., Wofchuck, T., 2009.

Indoor air quality impacts of an improved wood stove in Ghana and an ethanol stove in

Ethiopia. Energy Sustain. Dev. 13, 71–76. doi:10.1016/j.esd.2009.04.003

Petzold, A., Ogren, J.A., Fiebig, M., Laj, P., Li, S.-M., Baltensperger, U., Holzer-Popp, T., Kinne,

S., Pappalardo, G., Sugimoto, N., Wehrli, C., Wiedensohler, A., Zhang, X.-Y., 2013.

Recommendations for reporting “black carbon” measurements. Atmospheric Chem.

Phys. 13, 8365–8379. doi:10.5194/acp-13-8365-2013

Pokhrel, A.K., Bates, M.N., Acharya, J., Valentiner-Branth, P., Chandyo, R.K., Shrestha, P.S.,

Raut, A.K., Smith, K.R., 2015. PM2.5 in household kitchens of Bhaktapur, Nepal, using

four different cooking fuels. Atmos. Environ. 113, 159–168.

doi:10.1016/j.atmosenv.2015.04.060

Pope, C.A., Dockery, D.W., 2006. Health Effects of Fine Particulate Air Pollution: Lines that

Connect. J. Air Waste Manag. Assoc. 56, 709–742.

doi:10.1080/10473289.2006.10464485

Presser, C., Conny, J.M., Nazarian, A., 2014. Filter Material Effects on Particle Absorption

Optical Properties. Aerosol Sci. Technol. 48, 515–529.

doi:10.1080/02786826.2014.890999

Putti, V.R., Tsan, M., Mehta, S., Kammila, S., 2015. The state of the global clean and improved

cooking sector (Technical report No. 007/15). Energy Sector Management Assistance

Program | Global Alliance for Clean Cookstoves | The World Bank, Washington DC.

Puzzolo, E., 2013. Factors influencing the largescale uptake by households of cleaner and more

efficient household energy technologies.

Ramanahatan, V., 2007. Role of Black Carbon in Global and Regional Climate Changes.

Washington DC.

Ramanathan, N., Khan, B., Leong, I., Lukac, M., 2011a. Comparison Between Elemental Carbon

Measured Using Thermal nOptical Analysis and Black Carbon Measurements Using A

Novel Cellphone-Based System. Presented at the American Geophysical Union, Fall

Meeting, San Francisco, CA.

Ramanathan, N., Lukac, M., Ahmed, T., Kar, A., Praveen, P.S., Honles, T., Leong, I., Rehman,

I.H., Schauer, J.J., Ramanathan, V., 2011b. A cellphone based system for large-scale

monitoring of black carbon. Atmos. Environ. 45, 4481–4487.

doi:10.1016/j.atmosenv.2011.05.030

Ramanathan, V., Carmichael, G., 2008. Global and regional climate changes due to black

carbon. Nat. Geosci 1, 221–227. doi:10.1038/ngeo156

Rau, J.A., 1989. Composition and Size Distribution of Residential Wood Smoke Particles.

Aerosol Sci. Technol. 10, 181–192. doi:10.1080/02786828908959233

Page 120: INTO THE SMOKE: A RESEARCH ON HOUSEHOLD AIR ...oa.upm.es/49583/1/CANDELA_DE_LA_SOTA_SANDEZ.pdfCada año, la contaminación del aire interior provoca la muerte prematura de 4,3 millones

104

Reche, C., Viana, M., Brines, M., Pérez, N., Beddows, D., Alastuey, A., Querol, X., 2015.

Determinants of aerosol lung-deposited surface area variation in an urban

environment. Sci. Total Environ. 517, 38–47. doi:10.1016/j.scitotenv.2015.02.049

Rehman, I.H., Ahmed, T., Praveen, P.S., Kar, A., Ramanathan, V., 2011. Black carbon emissions

from biomass and fossil fuels in rural India. Atmospheric Chem. Phys. 11, 7289–7299.

doi:10.5194/acp-11-7289-2011

Riojas-Rodriguez, H., Schilmann, A., Marron-Mares, A.T., Masera, O., Li, Z., Romanoff, L.,

Sjödin, A., Rojas-Bracho, L., Needham, L.L., Romieu, I., 2011. Impact of the Improved

Patsari Biomass Stove on Urinary Polycyclic Aromatic Hydrocarbon Biomarkers and

Carbon Monoxide Exposures in Rural Mexican Women. Environ. Health Perspect. 119,

1301–1307. doi:10.1289/ehp.1002927

Rivas, I., Mazaheri, M., Viana, M., Moreno, T., Clifford, S., He, C., Bischof, O.F., Martins, V.,

Reche, C., Alastuey, A., Alvarez-Pedrerol, M., Sunyer, J., Morawska, L., Querol, X., 2017.

Identification of technical problems affecting performance of DustTrak DRX aerosol

monitors. Sci. Total Environ. 584–585, 849–855. doi:10.1016/j.scitotenv.2017.01.129

Rivas, I., Viana, M., Moreno, T., Bouso, L., Pandolfi, M., Alvarez-Pedrerol, M., Forns, J.,

Alastuey, A., Sunyer, J., Querol, X., 2015. Outdoor infiltration and indoor contribution

of UFP and BC, OC, secondary inorganic ions and metals in PM2.5 in schools. Atmos.

Environ. 106, 129–138. doi:10.1016/j.atmosenv.2015.01.055

Roden, C.A., Bond, T.C., Conway, S., Osorto Pinel, A.B., MacCarty, N., Still, D., 2009. Laboratory

and field investigations of particulate and carbon monoxide emissions from traditional

and improved cookstoves. Atmos. Environ. 43, 1170–1181.

doi:10.1016/j.atmosenv.2008.05.041

Roden, C.A., Bond, T.C., Conway, S., Pinel, A.B.O., 2006. Emission Factors and Real-Time Optical

Properties of Particles Emitted from Traditional Wood Burning Cookstoves. Environ.

Sci. Technol. 40, 6750–6757. doi:10.1021/es052080i

Roth, C., 2011. Micro-gasification: cooking with gas from dry biomass. An introduction to

concepts and applications of wood-gas burning technologies for cooking. GIZ.

Ruiz-Mercado, I., Masera, O., 2015. Patterns of Stove Use in the Context of Fuel–Device

Stacking: Rationale and Implications. EcoHealth 12, 42–56. doi:10.1007/s10393-015-

1009-4

Ruiz-Mercado, I., Masera, O., Zamora, H., Smith, K.R., 2011. Adoption and sustained use of

improved cookstoves. Energy Policy 39, 7557–7566. doi:10.1016/j.enpol.2011.03.028

Ruth, M., Maggio, J., Whelan, K., DeYoung, M., May, J., Peterson, A., Paterson, K., 2014.

Kitchen 2.0: Design guidance for healthier cooking environments. Int. J. Serv. Learn.

Eng. Humanit. Eng. Soc. Entrep. 151–169.

Ryan, S.E., 2014. Rethinking gender and identity in energy studies. Energy Res. Soc. Sci. 1, 96–

105. doi:10.1016/j.erss.2014.02.008

Rysankova, D., Putti, V.R., Hyseni, B., Kammila, S., Kappen, J.F., 2014. Clean and Improved

Cooking in Sub-Saharan Africa (No. 98664), African Clean Cookin Solutions Initiative.

The World Bank, Washington, DC, USA.

Safo-Adu, G., Ofosu, F.G., Carboo, D., Armah, Y.S., 2014. Heavy metals and black carbon

assessment of PM10 particulates along Accra–Tema highway in Ghana. Int. J. Sci.

Technol. 3.

Page 121: INTO THE SMOKE: A RESEARCH ON HOUSEHOLD AIR ...oa.upm.es/49583/1/CANDELA_DE_LA_SOTA_SANDEZ.pdfCada año, la contaminación del aire interior provoca la muerte prematura de 4,3 millones

105

Saho, B.F.M., Reiss, K., 2013. West Africa Clean Cooking Alliance (ECOWAS Clean Cooking

Energy Initiative).

Sahu, M., Peipert, J., Singhal, V., Yadama, G.N., Biswas, P., 2011. Evaluation of Mass and

Surface Area Concentration of Particle Emissions and Development of Emissions

Indices for Cookstoves in Rural India. Environ. Sci. Technol. 45, 2428–2434.

doi:10.1021/es1029415

Salako, G.O., 2012. Exploring the Variation between EC and BC in a Variety of Locations.

Aerosol Air Qual. Res. doi:10.4209/aaqr.2011.09.0150

Saleh, R., Robinson, E.S., Tkacik, D.S., Ahern, A.T., Liu, S., Aiken, A.C., Sullivan, R.C., Presto, A.A.,

Dubey, M.K., Yokelson, R.J., Donahue, N.M., Robinson, A.L., 2014. Brownness of

organics in aerosols from biomass burning linked to their black carbon content. Nat.

Geosci. 7, 647–650. doi:10.1038/ngeo2220

Samba, S.A.N., 2001. Effect of Cordyla pinnata litter on crop yield: An experimental

agroforestry approach. Ann. For. Sci. Fr.

Sambandam, S., Balakrishnan, K., Ghosh, S., Sadasivam, A., Madhav, S., Ramasamy, R.,

Samanta, M., Mukhopadhyay, K., Rehman, H., Ramanathan, V., 2015. Can Currently

Available Advanced Combustion Biomass Cook-Stoves Provide Health Relevant

Exposure Reductions? Results from Initial Assessment of Select Commercial Models in

India. EcoHealth 12, 25–41. doi:10.1007/s10393-014-0976-1

Sander, K., Hyseni, B., Haider, W., 2011a. Wood-Based Biomass Energy Development for Sub-

Saharan Africa: Issues and Approaches. (Working paper No. 74545). The World Bank,

Washington, DC, USA.

Sander, K., Hyseni, B., Haider, W., 2011b. Wood-Based Biomass Energy Development for Sub-

Saharan Africa. The World Bank, Washington, DC, USA.

Shailaja, R., 2000. Women, energy and sustainable development. Energy Sustain. Dev. 4, 45–

64.

Shen, G., Tao, S., Wei, S., Chen, Y., Zhang, Y., Shen, H., Huang, Y., Zhu, D., Yuan, C., Wang, H.,

Wang, Y., Pei, L., Liao, Y., Duan, Y., Wang, B., Wang, R., Lv, Y., Li, W., Wang, X., Zheng,

X., 2013. Field Measurement of Emission Factors of PM, EC, OC, Parent, Nitro-, and

Oxy- Polycyclic Aromatic Hydrocarbons for Residential Briquette, Coal Cake, and Wood

in Rural Shanxi, China. Environ. Sci. Technol. 47, 2998–3005. doi:10.1021/es304599g

Shen, G., Yang, Y., Wang, W., Tao, S., Zhu, C., Min, Y., Xue, M., Ding, J., Wang, B., Wang, R.,

Shen, H., Li, W., Wang, X., Russell, A.G., 2010. Emission Factors of Particulate Matter

and Elemental Carbon for Crop Residues and Coals Burned in Typical Household Stoves

in China. Environ. Sci. Technol. 44, 7157–7162. doi:10.1021/es101313y

Shoemaker, J.K., Schrag, D.P., Molina, M.J., Ramanahatan, V., 2013. What Role for Short-Lived

Climate Pollutants in Mitigation Policy? Science 342, 132–132.

doi:10.1126/science.1240162

Siddiqui, A.R., Lee, K., Bennett, D., Yang, X., Brown, K.H., Bhutta, Z.A., Gold, E.B., 2009. Indoor

carbon monoxide and PM 2.5 concentrations by cooking fuels in Pakistan. Indoor Air 19,

75–82. doi:10.1111/j.1600-0668.2008.00563.x

Simon, G.L., Bailis, R., Baumgartner, J., Hyman, J., Laurent, A., 2014. Current debates and

future research needs in the clean cookstove sector. Energy Sustain. Dev. 20, 49–57.

doi:10.1016/j.esd.2014.02.006

Page 122: INTO THE SMOKE: A RESEARCH ON HOUSEHOLD AIR ...oa.upm.es/49583/1/CANDELA_DE_LA_SOTA_SANDEZ.pdfCada año, la contaminación del aire interior provoca la muerte prematura de 4,3 millones

106

Smith, K.R., Bruce, N., Balakrishnan, K., Adair-Rohani, H., Balmes, J., Chafe, Z., Dherani, M.,

Hosgood, H.D., Mehta, S., Pope, D., Rehfuess, E., 2014. Millions Dead: How Do We

Know and What Does It Mean? Methods Used in the Comparative Risk Assessment of

Household Air Pollution. Annu. Rev. Public Health 35, 185–206. doi:10.1146/annurev-

publhealth-032013-182356

Smith, K.R., Mccracken, J.P., Thompson, L., Edwards, R., Shields, K.N., Canuz, E., Bruce, N.,

2010. Personal child and mother carbon monoxide exposures and kitchen levels:

Methods and results from a randomized trial of woodfired chimney cookstoves in

Guatemala (RESPIRE). J. Expo. Sci. Environ. Epidemiol. 20, 406–416.

doi:10.1038/jes.2009.30

Smith, K.R., Uma, R., Kishore, V.V.N., Lata, K., Joshi, V., Zhang, J.J., Rasmussen, R.A., Khalil,

M.A.K., 2000. GREENHOUSE GASES FROM SMALL-SCALE COMBUSTION DEVICES IN

DEVELOPING COUNTRIES (No. EPA-600/R-00-052), Research and Development. United

States Environmental Protect ion Agency, Washington, DC, USA.

Soneja, S.I., Tielsch, J.M., Curriero, F.C., Zaitchik, B., Khatry, S.K., Yan, B., Chillrud, S.N., Breysse,

P.N., 2015. Determining Particulate Matter and Black Carbon Exfiltration Estimates for

Traditional Cookstove Use in Rural Nepalese Village Households. Environ. Sci. Technol.

49, 5555–5562. doi:10.1021/es505565d

Sota, C. de la, Lumbreras, J., Mazorra, J., Narros, A., Fernández, L., Borge, R., 2014.

Effectiveness of Improved Cookstoves to Reduce Indoor Air Pollution in Developing

Countries. The Case of the Cassamance Natural Subregion, Western Africa. J. Geosci.

Environ. Prot. 02, 1–5. doi:10.4236/gep.2014.21001

The Gold Standard, 2015. Quantification of climate related emission reductions of Black

Carbon and Co-­emitted Species due to the replacement of less efficient cookstoves

with improved efficiency cookstoves.

Tillman, D.A., 1987. Wood as an Energy Resource. Academic Press.

Tiwari, M., Sahu, S.K., Bhangare, R.C., Yousaf, A., Pandit, G.G., 2014. Particle size distributions

of ultrafine combustion aerosols generated from household fuels. Atmospheric Pollut.

Res. 5, 145–150. doi:10.5094/APR.2014.018

Todea, A., Beckmann, S., Kaminski, K., Monz, C., Dahmann, D., Neumann, V., Dozol, H., 2016.

Inter-comparison or personal monitors for nanoparticle exposure at workplaces and in

the environment. Sci. Total Environ.

Torres-Duque, C., Maldonado, D., Perez-Padilla, R., Ezzati, M., Viegi, G., on behalf of the Forum

of International Respiratory Societies (FIRS) Task Force on Health Effects of Biomass

Exposure, 2008. Biomass Fuels and Respiratory Diseases: A Review of the Evidence.

Proc. Am. Thorac. Soc. 5, 577–590. doi:10.1513/pats.200707-100RP

UNEP, WMO, 2011. Integrated Assessment of Black Carbon and Tropospheric Ozone.

Venkataraman, C., Habib, G., Eiguren-Fernandez, A., Miguel, A.H., Friedlander, S.K., 2005.

Residential Biofuels in South Asia: Carbonaceous Aerosol Emissions and Climate

Impacts. Science 307, 1454–1456. doi:10.1126/science.1104359

Venkataraman, C., Rao, G.U.M., 2001. Emission Factors of Carbon Monoxide and Size-Resolved

Aerosols from Biofuel Combustion. Environ. Sci. Technol. 35, 2100–2107.

doi:10.1021/es001603d

Viana, M., Rivas, I., Reche, C., Fonseca, A.S., Pérez, N., Querol, X., Alastuey, A., Álvarez-

Pedrerol, M., Sunyer, J., 2015. Field comparison of portable and stationary instruments

Page 123: INTO THE SMOKE: A RESEARCH ON HOUSEHOLD AIR ...oa.upm.es/49583/1/CANDELA_DE_LA_SOTA_SANDEZ.pdfCada año, la contaminación del aire interior provoca la muerte prematura de 4,3 millones

107

for outdoor urban air exposure assessments. Atmos. Environ. 123, 220–228.

doi:10.1016/j.atmosenv.2015.10.076

Wang, Yungang, Sohn, M.D., Wang, Yilun, Lask, K.M., Kirchstetter, T.W., Gadgil, A.J., 2014. How

many replicate tests are needed to test cookstove performance and emissions? ?

Three is not always adequate. Energy Sustain. Dev. 20, 21–29.

doi:10.1016/j.esd.2014.02.002

Wathore, R., Mortimer, K., Grieshop, A.P., 2017. In-Use Emissions and Estimated Impacts of

Traditional, Natural- and Forced-Draft Cookstoves in Rural Malawi. Environ. Sci.

Technol. 51, 1929–1938. doi:10.1021/acs.est.6b05557

Watson, J.G., Chow, J.C., Chen, L.-W.A., 2005. Summary of Organic and Elemental

Carbon/Black Carbon Analysis Methods and Intercomparisons. Aerosol Air Qual. Res. 5,

65–102.

WB, 2013. Global Tracking Framework. Sustainable Energy for all. The World Bank,

Washington, DC, USA.

Weingartner, E., Saathoff, H., Schnaiter, M., Streit, N., Bitnar, B., Baltensperger, U., 2003.

Absorption of light by soot particles: determination of the absorption coefficient by

means of aethalometers. J. Aerosol Sci. 34, 1445–1463. doi:10.1016/S0021-

8502(03)00359-8

WHO, 2016. Burning opportunity: clean household energy for health, sustainable

development, and wellbeing of women and children.

WHO, 2015. Improving and harmonizing methods and tools that incorporate measurement of

fuel stacking and impacts on women and children. : World Health Organization.

Geneva, Switzerland.

WHO, 2014a. Burden of disease from Household Air Pollution for 2012.

WHO (Ed.), 2014b. WHO guidelines for indoor air quality: household fuel combustion. World

Health Organization, Geneva, Switzerland.

WHO, 2009. Country profile of Environmental Burden of Disease. Senegal. World Health

Organization, Geneva, Switzerland.

WHO, 2007. Indoor Air Pollution: National Burden of Disease Estimates. World Health

Organization, Geneva, Switzerland.

WHO, 2006. WHO Air quality guidelines for particulate matter, ozone, nitrogen dioxide and

sulfur dioxide. Geneva, Switzerland.

WHO, EPA, 2016. Using stove emissions Data to Estimate Air Quality: An overview of WHO

current modelling and future plans.

WHO, Scovronik, N., 2015. Reducing global health risks through mitigation of short-lived

climate pollutants. Scoping report for policymakers. WHO, Switzerland.

Wickramasinghe, A., 2003. Gender and health issues in the biomass energy cycle: impediments

to sustainable development. Energy Sustain. Dev. 7, 51–61.

Wodon, Q., Blackden, C.M. (Eds.), 2006. Gender, Time Use, and Poverty in Sub-Saharan Africa,

World Bank Working Papers. The World Bank. doi:10.1596/978-0-8213-6561-8

Woolson, R.F., 2008. Wilcoxon Signed-Rank Test. Wiley Encycl. Clin. Trials 1–3.

Yan, B., Kennedy, D., Miller, R.L., Cowin, J.P., Jung, K., Perzanowski, M., Balleta, M., Perera,

F.P., Kinney, P.L., Chillrud, S.N., 2011. Validating a nondestructive optical method for

apportioning colored particulate matter into black carbon and additional components.

Atmos. Environ. 45, 7478–7486. doi:10.1016/j.atmosenv.2011.01.044

Page 124: INTO THE SMOKE: A RESEARCH ON HOUSEHOLD AIR ...oa.upm.es/49583/1/CANDELA_DE_LA_SOTA_SANDEZ.pdfCada año, la contaminación del aire interior provoca la muerte prematura de 4,3 millones

108

Zanatta, M., Gysel, M., Bukowiecki, N., Müller, T., Weingartner, E., Areskoug, H., Fiebig, M.,

Yttri, K.E., Mihalopoulos, N., Kouvarakis, G., Beddows, D., Harrison, R.M., Cavalli, F.,

Putaud, J.P., Spindler, G., Wiedensohler, A., Alastuey, A., Pandolfi, M., Sellegri, K.,

Swietlicki, E., Jaffrezo, J.L., Baltensperger, U., Laj, P., 2016. A European aerosol

phenomenology-5: Climatology of black carbon optical properties at 9 regional

background sites across Europe. Atmos. Environ. 145, 346–364.

doi:10.1016/j.atmosenv.2016.09.035

Zhang, H., Wang, S., Hao, J., Wan, L., Jiang, J., Zhang, M., Mestl, H.E.S., Alnes, L.W.H., Aunan, K.,

Mellouki, A.W., 2012. Chemical and size characterization of particles emitted from the

burning of coal and wood in rural households in Guizhou, China. Atmos. Environ. 51,

94–99. doi:10.1016/j.atmosenv.2012.01.042

Zhang, J., Kotani, K., 2012. The determinants of household energy demand in rural Beijing: Can

environmentally friendly technologies be effective? Energy Econ. 34, 381–388.

doi:10.1016/j.eneco.2011.12.011

Zhang, J., Smith, K.R., Ma, Y., Ye, S., Jiang, F., Qi, W., Liu, P., Khalil, M.A.K., Rasmussen, R.A.,

Thorneloe, S.A., 2000. Greenhouse gases and other airborne pollutants from

household stoves in China: a database for emission factors. Atmos. Environ. 34, 4537–

4549.

Zhou, Z., Dionisio, K.L., Verissimo, T.G., Kerr, A.S., Coull, B., Howie, S., Arku, R.E., Koutrakis, P.,

Spengler, J.D., Fornace, K., Hughes, A.F., Vallarino, J., Agyei-Mensah, S., Ezzati, M.,

2014. Chemical Characterization and Source Apportionment of Household Fine

Particulate Matter in Rural, Peri-urban, and Urban West Africa. Environ. Sci. Technol.

48, 1343–1351. doi:10.1021/es404185m

Zuk, M., Rojas, L., Blanco, S., Serrano, P., Cruz, J., Angeles, F., Tzintzun, G., Armendariz, C.,

Edwards, R.D., Johnson, M., others, 2007. The impact of improved wood-burning

stoves on fine particulate matter concentrations in rural Mexican homes. J. Expo. Sci.

Environ. Epidemiol. 17, 224–232.

Page 125: INTO THE SMOKE: A RESEARCH ON HOUSEHOLD AIR ...oa.upm.es/49583/1/CANDELA_DE_LA_SOTA_SANDEZ.pdfCada año, la contaminación del aire interior provoca la muerte prematura de 4,3 millones