Você está na página 1de 398

Universidade de Aveiro Universidade do Porto

Ano 2023

Bernardino José MODELAÇÃO DOS IMPACTES AMBIENTAIS E DE


Bernardo SAÚDE HUMANA NA ÁREA ENVOLVENTE À
LIXEIRA DE HULENE-B, MAPUTO, MOÇAMBIQUE

MODELLING OF ENVIRONMENTAL AND HUMAN


HEALTH IMPACTS ON THE SURROUNDINGS OF
HULENE-B DUMP, MAPUTO, MOZAMBIQUE
II
III

Universidade de Aveiro Universidade do Porto


2023 Faculdade de Ciências

Bernardino José MODELLING OF ENVIRONMENTAL AND HUMAN


Bernardo HEALTH IMPACTS ON THE SURROUNDINGS OF
HULENE-B DUMP, MAPUTO, MOZAMBIQUE

MODELAÇÃO DOS IMPACTES AMBIENTAIS E DE


SAÚDE HUMANA NA ÁREA ENVOLVENTE À
LIXEIRA DE HULENE-B, MAPUTO, MOÇAMBIQUE

Tese apresentada à Universidade de Aveiro e Universidade do Porto para


cumprimento dos requisitos necessários à obtenção do grau de Doutor em
Geociências da Universidade de Aveiro / Universidade do Porto, realizada sob
a orientação científica do Professor Doutor Fernando Rocha Professor, do
Departamento de Geociências da Universidade de Aveiro, e da Doutora Carla
Candeias, Investigadora no Departamento de Geociências da Universidade de
Aveiro.

Apoio financeiro pelo Instituto Camões, IP;


Universidade Pedagógica de Maputo e
Fundo Nacional de Investigação,
Moçambique.
IV
V

À minha mãe (Angélica Maiela Macamo), guerreira e um amor incondicional.

Ao meu pai, Bernardo Chiquelequete (em memória).

Aos meus irmãos e irmãs, especialmente ao Miguel Bernardo um forte abraço


e Anibal Bernardo (em memória), pela inspiração.

À todos que me querem bem!


VI
VII

o júri
Presidente Prof. Doutor João Carlos de Oliveira Matias
Professor Catedrático, Universidade de Aveiro

Vogais Prof. Doutor Fernando Joaquim Fernandes Tavares Rocha


Professor Catedrático, Universidade de Aveiro

Prof. Doutor Rui Miguel Marques Moura


Professor Auxiliar da Universidade do Porto

Prof. Doutor Maria Isabel Garrido Prudêncio


Investigadora Coordenadora do Instituto Superior Técnico

Prof. Doutor Gustavo Sobrinho Dgedge


Professor Associado da Universidade Pedagógica de Maputo

Doutora Slavka Carvalho Andrejkovicová


Investigadora Principal Universidade de Aveiro

Prof. Doutor Hassina Mouri


Professor Catedrático, Universidade de Joanesburgo

Prof. Doutor Daúd Liace Jamal


Professor Associado da Universidade Eduardo Mondlane
VIII
IX

Agradecimentos Gostaria de agradecer à todos que tornaram possível, direta e indiretamente, a


concretização deste sonho.
Aos meus orientadores pelo apoio incondicional. Em especial ao Professor
Doutor Fernando Rocha pela amizade, respeito e cordialidade, o meu muito
obrigado. A Doutora Carla Candeias muito obrigado pela paciência ao longo de
todo processo, desde as análises laboratoriais até análise de dados.
Aos docentes do programa doutoral em Geociências, muito obrigado pelos
ensinamentos, que serviram de base na escolha do tema do projeto doutoral.
Um agradecimento muito especial vai para a Professora Doutora Deolinda
Flores, que sempre esteve disponível e com boa disposição para clarificar
todas inquietações desde a logística e no decurso do programa doutoral. Um
abraço especial ao Professor Doutor Rui Moura que gentilmente esteve aberto
para ensinar e aprimorar as abordagens em Geofísica no contexto do meu
projeto de Tese.
Aos meus colegas e amigos, especialmente João Unguana pela energia e
positividade transmitida ao longo da longa marcha da prospeção geofísica,
muito obrigado.
Ao Instituto Camões, IP pela bolsa de estudos concedida em 2018 e pela
simpatia e respeito de todos colaboradores do Instituto que prontamente
estavam disponíveis para auxiliar em todos os aspetos técnicos. Ao Instituto de
Bolsa de Estudo de Moçambique, muito obrigado pela comparticipação no
pagamento das passagens aéreas.
A Universidade Pedagógica de Maputo, muito obrigado pela autorização para
a continuação de estudos e pelo apoio incondicional prestado no decurso do
programa doutoral.
Ao Fundo Nacional de Investigação (FNI) de Moçambique, muito obrigado por
ter financiado a primeira campanha de amostragem.
À Direção da Faculdade de Ciências da Terra e Ambiente, por toda
cordialidade e apoio incondicional prestado no decurso da pesquisa. A
Faculdade de Ciências Naturais e Matemática, por ter permitido a secagem
das amostras no laboratório de Microbiologia.
Aos meus colegas e amigos da Faculdade de Ciências da Terra e Ambiente,
especialmente a Professora Doutora Elisa Eda Nhambire e Alice Freia, muito
obrigado pelo todo apoio. Ao meu colega e amigo José Neves, muito obrigado
pela toda força e apoio nesta luta doutoral.
Aos meus colegas da Universidade de Porto, Faculdade de Ciências, muito
obrigado por todos momentos partilhado durante o ano lectivo 2018/2019.
Aos funcionários do Laboratório GeoBioTec - Universidade de Aveiro, em
especial às Engenheiras Cristina Cerqueira e Denise Terroso pelas análises de
FRX e DRX.
A toda minha família, mãe (Angélica Maiela Macamo), irmãos, irmãs, sobrinhos
e sobrinhas pelo carinho e amor incondicional.
Aos amigos da longa marcha doutoral e dos momentos, em especial ao Sérgio
Cossa, Rafael Baieta e tantos outros, muito obrigado.
X

Acknowledgement I would like to thank everyone who made it possible, directly, and indirectly,
for this dream to come true.
To my supervisors for their unconditional support. In special to Professor
Fernando Rocha for his friendship, respect, and cordiality, thank you very
much.
To Doctor Carla Candeias, thank you very much for your patience throughout
the process, from laboratory analysis to data analysis.
To the teachers of the doctoral program in Geosciences, thank you very
much for the teachings, which served as a basis in the choice of the subject
of the doctoral project. A very special thanks goes to Professor Deolinda
Flores, who was always available and willing to clarify all concerns since the
logistics and during the doctoral program. A special thanks to Professor Rui
Moura who was kindly open to teach and improve the approaches in
Geophysics in the context of my thesis project.
To my colleagues and friends, especially João Unguana for the energy and
positivity transmitted along the long march of geophysical prospection, thank
you very much.
To Instituto Camões, IP for the scholarship granted in 2018 and for the
sympathy and respect of all collaborators of the Institute who were readily
available to assist in all technical aspects. To the Instituto de Bolsa de
Estudo de Moçambique, thank you very much for the contribution in the
payment of the air tickets.
To Pedagogical University of Maputo, thank you very much for the
authorization to continue studies and for the unconditional support provided
during the doctoral program.
To the National Research Fund (FNI) of Mozambique, thank you very much
for funding the first sampling campaign.
To the Direction of the Faculty of Earth and Environmental Sciences, for all
cordiality and unconditional support provided during the research. To the
Faculty of Natural Sciences and Mathematics, for allowing the drying of the
samples in the Microbiology laboratory.
To my colleagues and friends at the Faculty of Earth and Environmental
Sciences, especially Professor Elisa Eda Nhambire, Octávio de Jesus and
Alice Freia, thank you very much for all the support. To my colleague and
friend José Neves, thank you very much for all the strength and support in
this doctoral struggle.
To my colleagues at the University of Porto, Faculty of Sciences, thank you
very much for all the moments shared during the academic year 2018/2019.
To the employees of GeoBioTec Laboratory - University of Aveiro, especially
to the engineers Cristina Cerqueira and Denise Terroso for the XRF and
XRD analyses.
To all my family, mother (Angélica Maiela Macamo), brothers, sisters,
nephews and nieces for the affection and unconditional love.
To my friends from the long doctoral march and the moments, especially Sérgio
Cossa, Rafael Baieta and many others, thank you very much.
XI

Geofísica, geoquímica, lixeira de Hulene-B, contaminação ambiental, avaliação


palavras-chave
de riscos de saúde humana

resumo A lixeira de Hulene-B está localizada na área suburbana à norte da cidade de


Maputo, capital de Moçambique. É considerada a maior lixeira ao ar livre do
país. Na sua extremidade ocidental, caracteriza-se pela dispersão de lixiviados
no solo e pela deposição de cinzas da incineração de resíduos, sendo que
algumas faixas desta área são utilizadas para o cultivo de pequenas hortas,
para além de habitações. Devido à baixa profundidade do lençol freático, parte
da população recorre à água dos poços para consumo diário.
Esta investigação, desenvolvida no contexto de ambiente e saúde, visa
explicar os impactes ambientais da lixeira (solos, águas subterrâneas, água
dos superficiais, solos da rizosfera, plantas comestíveis e poeiras das
estradas) e os riscos para a saúde humana. Foram definidos locais de
referência, considerados menos impactadas pela lixeira, para a resistividade
elétrica e a análise geoquímica. Assim, foram integrados métodos geofísicos
(resistividade elétrica) para explicar os mecanismos de contaminação das
águas subterrâneas por plumas de lixiviado e métodos geoquímicos para
avaliar os níveis de contaminação nos solos, águas subterrâneas, águas
superficiais, solos da rizosfera, plantas comestíveis e poeiras de estradas. A
distribuição espacial das alterações das propriedades do solo e dos níveis de
contaminação foi baseada em SIG. Os modelos de resistividade elétrica
mostraram a predominância de áreas anómalas consideradas contaminadas
por plumas de lixiviado que se deslocam através de estruturas litológicas para
as águas subterrâneas. As plumas mostraram ser extensas no Verão e com
um movimento predominantemente E-W. A distribuição espacial das plumas
no ambiente subterrâneo mostrou uma diminuição à medida que nos
afastamos da lixeira. A análise das propriedades do solo integrada aos
modelos de avaliação do risco ambiental mostrou uma forte alteração dos
solos na envolvente da lixeira e um elevado risco de contaminação do solo,
das águas subterrâneas e das águas superficiais. Os resultados da análise das
águas subterrâneas e de superfície revelaram concentrações elevadas de
PO43-, Hg,Pb e Zn. No entanto, Pb representa o risco saúde mais significativo.
A degradação ecológica nas proximidades da lixeira mostrou ser mais elevada
para os elementos Zn, Cu, Cr, Cr, Cr, Zr , Pb, Ni, Mn e maior risco para saúde
por Cr, Cu, Mn, Ni, Pb, Zn e Zr e o Pb representou risco elevado
principalmente para crianças. Os solos da rizosfera mostraram uma maior
contaminação por Pb, Cu, Zn, Zr e nas plantas comestíveis Zn, Ni, Cr, Cu,
Mn, Pb, Fe, Ti, Zr, sugerindo maior transferência solo para as plantas de Pb,
Cu, Zn e Zr e por possível deposição de cinza de Ni, Cr, Mn, Fe, Ti e Zr. O
coeficiente de risco de risco de saúde dos solos da rizosfera e das plantas
comestíveis era muito elevado, sugerindo possíveis efeitos adversos para a
saúde, principalmente para crianças. Nas poeiras de estrada Cr, Cu, Fe, Mn,
Ni, Pb, Ti, Zn, e Zr representaram um elevado nível de degradação ecológica.
O quociente de risco saúde identificou Cu, Fe, Ni, e Pb como os elementos de
maior risco para saúde e Ni e Pb representando risco cancerígeno. Os
resultados sugerem a necessidade de monitorização contínua da lixeira e a
definição de medidas de mitigação do impacto ambiental tendo em conta os
principais mecanismos de contaminação (lixiviados, cinzas de incineração e
partículas) da área circundante, bem como a redução dos níveis de exposição
da população local e das crianças em particular dada ao elevado risco de
saúde que a lixeira representa.
XII

keywords
Geophysic, geochemistry, Hulene-B dump, environmental contamination,
human health risk assessment

abstract Hulene-B dump is located in the northern suburban area of Maputo city, the
capital of Mozambique. It is considered the largest open-air dump in the
country. At its western end it is characterized by the dispersion of leachate on
the ground and the deposition of ashes from waste incineration and some strips
of this area are used for the cultivation of small vegetable gardens in addition to
housing. Due to the low depth of the water table, part of the population resorts
to well water for daily consumption. This research, developed in the context of
environment and health, aims at explaining the environmental impacts of the
dump (soils, groundwater, surface water, rhizosphere soils, edible plants, and
road dust) and the risks for human health. Reference sites, considered to be
less impacted by the dump, were defined for electrical resistivity and
geochemical analysis. Thus, geophysical methods (electrical resistivity) were
integrated to explain the mechanisms of groundwater contamination by
leachate plumes and geochemical methods to assess contamination levels in
soils, groundwater, surface water, rhizosphere soils, edible plants, and road
dust. The spatial distribution of changes in soil properties and contamination
levels was GIS-based. The electrical resistivity models showed the
predominance of anomalous areas considered contaminated by leachate
plumes moving through lithological structures into groundwater. The plumes
were shown to be extensive in summer and with a predominantly E-W
movement. The spatial distribution of the plumes in the subsurface environment
showed a decrease as one moves away from the dump. Analysis of soil
properties integrated into the environmental risk assessment models showed a
strong alteration of soils in the surroundings of the landfill and a high risk of soil,
groundwater, and surface water contamination. The results of groundwater
and surface water analysis showed high concentrations of PO 43-, Hg, Pb and
Zn. However, Pb represents the most significant health risk. Ecological
degradation of the area surrounding the Hulene-B dump has been shown to be
higher for Zn, Cu, Cr, Cr, Zr , Pb , Ni , Mn and higher health risk for Cr, Cu, Mn,
Ni, Pb, Zn and Zr and Pb represented a high risk mainly for children.
Rhizosphere soils showed higher contamination for Pb, Cu, Zn, Zr and in edible
plants by Zn, Ni, Cr, Cu, Mn, Pb, Fe, Ti, Zr suggesting transfer factor effect of
soil contaminants for Pb, Cu, Zn and Zr and by possible deposition of
contaminated ash for Ni, Cr, Mn, Fe, Ti and Zr. The health risk coefficient of
rhizosphere soils and edible plants was very high, suggesting possible adverse
health effects, especially for children. In road dust Cr, Cu, Fe, Mn, Ni, Pb, Ti,
Zn, and Zr represented a high level of ecological degradation. The health risk
quotient identified Cu, Fe, Ni, and Pb as the elements with the highest health
risk and Ni and Pb represented a carcinogenic risk.The results suggest the
need for continuous monitoring of the dump and the definition of environmental
impact mitigation measures taking into account the main contamination
mechanisms (leachates, incineration ashes and particles) of the surrounding
area, as well as the reduction of exposure levels of the local population and
children in particular
XIII

Contents
Part I. Approach ............................................................................................................................................................................ 1
1. Introduction ............................................................................................................................................................................ 3
1.1 Thesis presentation .......................................................................................................................................................... 7
2. Study Area ........................................................................................................................................................................... 11
2.1 Location of Mozambique ................................................................................................................................................ 11
2.2 LocatioHulene-B waste dump ........................................................................................................................................ 12
2.3 Climate of the study area ............................................................................................................................................... 13
2.4 Geomorphology .............................................................................................................................................................. 14
2.4.1 Morphology of Mozambique ................................................................................................................................ 14
2.4.2 Morphology of Maputo City ................................................................................................................................. 15
2.4.3 Morphology of the study area.............................................................................................................................. 16
2.5 Geological context .......................................................................................................................................................... 16
2.5.1 Regional geology ................................................................................................................................................ 16
2.5.2 African Karoo system .......................................................................................................................................... 19
2.5.3 Mozambique Geology ......................................................................................................................................... 20
2.5.4 Geology of Maputo City ...................................................................................................................................... 22
2.5.5 Geology of the study area ................................................................................................................................... 24
2.6 Soils ............................................................................................................................................................................... 26
2.7 Hydrogeology ................................................................................................................................................................. 27
2.8 Flora and fauna .............................................................................................................................................................. 28
2.9 Brief history of the Hulene-B Waste dump ..................................................................................................................... 30
2.10 Socio-environmental context ........................................................................................................................................ 33
2.11 Environmental impact of the surroundings of the dump and governmental mitigation initiatives .................................. 34
2.12 Social and environmental context of the background local area .................................................................................. 36
3. Materials and Methods ......................................................................................................................................................... 37
3.1 Geophysical studies ....................................................................................................................................................... 37
3.1.1 Methods adopted ................................................................................................................................................ 41
3.1.2 Equipment ........................................................................................................................................................... 41
3.1.3 Planning and prospecting campaigns ................................................................................................................. 42
3.1.4 Data acquisition .................................................................................................................................................. 44
3.1.5 Acquisition and recording of readings ................................................................................................................. 44
3.1.6 Executed profiles ................................................................................................................................................ 47
3.1.7 Data conversion .................................................................................................................................................. 50
3.1.8 Model noise and correction methods .................................................................................................................. 51
3.1.9 Inversion of the final model ................................................................................................................................. 52
3.1.10 Groundwater Vulnerability Assessment (DRASTIC modified) ........................................................................... 53
3.1.11 Validation .......................................................................................................................................................... 55
3.2 Geochemical studies (soil, rhizosphere, stream water, groundwater, edible plants, road dust) ..................................... 56
3.2.1 Soil sampling....................................................................................................................................................... 56
3.2.2 Rhizosphere soils and edible plants .................................................................................................................... 58
3.2.3 Water sampling (wells and stream water) ........................................................................................................... 59
3.2.4 Road dust Sampling............................................................................................................................................ 60
3.3 Sample preparation and analysis ................................................................................................................................... 61
3.3.1 Soils, rhizosphere soils and road dust ................................................................................................................ 61
3.3.2 pH ....................................................................................................................................................................... 62
3.3.3 Electrical conductivity (EC) ................................................................................................................................. 62
3.3.4 Organic matter (OM) ........................................................................................................................................... 63
XIV

3.3.5 Moisture .............................................................................................................................................................. 64


3.3.6 Granulometry ...................................................................................................................................................... 65
3.4 Chemistry (FRX) and Mineralogy analyses (XRD) ......................................................................................................... 65
3.4.1 Soils, rhizosphere soils, road dust and edible plants .......................................................................................... 65
3.4.2 Water (wells water and stream water) analyses.................................................................................................. 65
3.4.3 Data Analyze....................................................................................................................................................... 66
3.4.4 Descriptive analysis ............................................................................................................................................ 67
3.4.5 Correlations analysis (r) ...................................................................................................................................... 67
3.4.6 Principal Component Analyses (PCA) ................................................................................................................ 67
3.4.7 Analysis of Variance (1-way ANOVA) ................................................................................................................. 68
3.4.8 Spatial distribution............................................................................................................................................... 68
3.5 Ecological risk assessment (soils, road dust) ................................................................................................................. 69
3.5.1 Potential ecological risk index (PERI) ................................................................................................................. 69
3.5.2 Soil Nemerow index (Pn) .................................................................................................................................... 70
3.5.3 Pollution Load index (PLI) ................................................................................................................................... 70
3.5.4 Pollution Index water........................................................................................................................................... 71
3.6 Landfill pollution Index .................................................................................................................................................... 71
3.6.1 Soils and groundwater ........................................................................................................................................ 71
3.6.2 Surface water risk ............................................................................................................................................... 73
3.7 Human health risk assessment ...................................................................................................................................... 75
3.8 Soils (surroundings area of Hulene -B Dump) ................................................................................................................ 75
3.8.1 Water health risk assessment (WRHA) ............................................................................................................... 77
3.8.2 Soil/plant transfer factor (TF) .............................................................................................................................. 78
3.8.3 Plant health risk assessment .............................................................................................................................. 78

Part II. Results and Discussion ................................................................................................................................................. 81


4. Paper 1 (Appendix 2) ........................................................................................................................................................... 83
4.1 Geophysical Studies....................................................................................................................................................... 84
4.1.1 Profile 1: 2020 (a) and 2021 (b) .......................................................................................................................... 84
4.1.2 Profile 2 2020 (a) and 2021(b) ............................................................................................................................ 85
4.1.3 Profile 3 2020 (a) and 2021 (b) ........................................................................................................................... 87
4.1.4 Profile 4 in 2020 (a) and 2021 (b) ....................................................................................................................... 89
4.2 Spatial Distribution of Possible Leachate Plumes (2020–2021) ..................................................................................... 90
5. Paper 2 (Appendix 3) ........................................................................................................................................................... 93
5.1 Results and discussion................................................................................................................................................... 94
5.1.1 Resistivity models and potential contamination risk ............................................................................................ 94
5.2 Modified DRASTIC index ............................................................................................................................................... 98
5.2.1 Depth to water table (D) ...................................................................................................................................... 98
5.2.2 Net Recharge (R) ................................................................................................................................................ 98
5.2.3 Aquifer media (A) ................................................................................................................................................ 98
5.2.4 Distance of the anomalous surface layer and groundwater (S)........................................................................... 99
5.2.5 Topography (T) ................................................................................................................................................... 99
5.2.6 Vadose (I) ........................................................................................................................................................... 99
5.2.7 Hydraulic conductivity (C) ................................................................................................................................. 100
5.3 Descriptive statistics of electrical resistivity and DRASTIC index ................................................................................. 101
5.4 Data Integration ............................................................................................................................................................ 102
6. Paper 3 (Appendix 4) ......................................................................................................................................................... 107
6.1 Results and discussion................................................................................................................................................. 108
6.1.1 Principal Component Analysis (PCA) .................................................................................................................. 114
6.2 Landfill risk index assessment (Ip) ............................................................................................................................... 116
XV

7. Paper 4 (Appendix 5) ......................................................................................................................................................... 119


7.1 Results and Discussion ................................................................................................................................................ 120
7.2 Principal Component Analysis (PCA) ........................................................................................................................... 126
7.3 Risk of surface water contamination (Pbci) .................................................................................................................. 129
8. Paper 5 (Appendix 6) ......................................................................................................................................................... 131
8.1 Results and Discussion ................................................................................................................................................ 132
8.1.1 Mineralogical and Chemical Characterization ................................................................................................... 132
8.2 Environmental Risk Assessment .................................................................................................................................. 138
8.3 Human Health Risk Assessment .................................................................................................................................. 141
9. Paper 6 (Appendix 7) ......................................................................................................................................................... 145
9.1 Results and discussion................................................................................................................................................. 146
9.1.1 Risk assessment of rhizosphere soils .................................................................................................................. 146
9.2 Water risk assessment ................................................................................................................................................. 150
9.3 Edible plants................................................................................................................................................................. 152
9.4 Data integration ............................................................................................................................................................ 153
10. Paper 7 (Appendix 8) ......................................................................................................................................................... 159
10.1 Results and discussion............................................................................................................................................... 160

Part III. Conclusions ................................................................................................................................................................. 169


11. Conclusions........................................................................................................................................................................ 171
11.1 Future Studies ............................................................................................................................................................ 174
12. References ......................................................................................................................................................................... 177
13. Appendixes ........................................................................................................................................................................ 204
13.1 Appendix 1. Characterization of some Potentially Toxic Elements ............................................................................. 207
13.2 Paper in their published format and submitted ........................................................................................................... 215
13.2.1 Appendix 2. Paper 1 ....................................................................................................................................... 217
13.2.2 Appendix 3. Paper 2 ....................................................................................................................................... 233
13.2.3 Appendix 4. Paper 3 ....................................................................................................................................... 255
13.2.4 Appendix 5. Paper 4 ....................................................................................................................................... 271
13.2.5 Appendix 6. Paper 5 ....................................................................................................................................... 309
13.2.6 Appendix 7. Paper 6 ....................................................................................................................................... 327
13.2.7 Appendix 8. Paper 7 ....................................................................................................................................... 349
XVI

List of figures
Figure 1.1 Relationship between the variables studied ................................................................................................................. 7
Figure 2.1 Location of Mozambique, adapt. https://www.nationsonline.org/oneworld/map/mozambique_map. ........................... 11
Figure 2.2 Location of Hulene-B waste dump. Adapt. Google Earth Pro. .................................................................................... 12
Figure 2.3 Precipitation and average monthly temperatures from 1987-2017 .............................................................................. 13
Figure 2.4 wind direction along the year ...................................................................................................................................... 14
Figure 2.5 Geomorphology of Mozambique ................................................................................................................................. 15
Figure 2.6 Topography of the study area - Projected from USGS/Maputo MDE .......................................................................... 16
Figure 2.7 African Craton ............................................................................................................................................................. 18
Figure 2.8 Archean cratonic nucleic ............................................................................................................................................. 18
Figure 2.9 Distribution of the Karoo basins in central and southern Africa ................................................................................... 20
Figure 2.10 Geological units of Mozambique ................................................................................................................................. 23
Figure 2.11 Geology of Maputo City............................................................................................................................................... 24
Figure 2.12 Geology of the study area: TPv Ponta Vermelha Formation and QMa Malhazine Formation ..................................... 25
Figure 2.13 Cross-section of the geological formations in the study area ...................................................................................... 26
Figure 2.14 Distribution of soils in the study area, with (dAj) the white/yellow sandy soils dune phase, and
(G) the reddish sandy silty soils ........................................................................................................................................ 27
Figure 2.15 (a) Surface water retained in the dune depression used for agriculture, and (b) Surface water in
the depression, dwellings area. ................................................................................................................................... 28
Figure 2.16 (a) Ground vegetation (b) Phragmites Australis .......................................................................................................... 29
Figure 2.17 (a) Vegetable crop, (b) Pumpkin crop and (c) Rice crop. ............................................................................................ 30
Figure 2.18 Hulene-B waste dump extension................................................................................................................................. 31
Figure 2.19 (a) Types of waste; (b) Height of waste and (c) Uncontrolled leachate flows at the western boundary of the dump ... 32
Figure 2.20 (a) Houses buried by the dump collapse, (b) Roads blocked with waste mass. .......................................................... 32
Figure 2.21 Environmental Context: (a) Maputo International Airport; (b) Hulene-B intradune depression;
(c) Hulene-B waste dump; (d) Julius Nyerere Avenue. .............................................................................................. 35
Figure 2.22 (a) wastewater from the dump, and (b) air pollution. ................................................................................................... 35
Figure 2.23 (a) Compacted waste inside the dump, (b) waste drainage channels and (c) leachate collection tanks ..................... 35
Figure 2.24 (a) e (b) New deposits to the south-west of the waste dump ...................................................................................... 36
Figure 2.25 Local background area (a) Soils of the TPv Formation, (b) Soils of the QMa Formation. ........................................... 37
Figure 3.1 Methodological summary. ........................................................................................................................................... 41
Figure 3.2 (a and b) ABEM SAS 4000 Resistivity meter used for data acquisition....................................................................... 42
Figure 3.3 Profiles executed in 2020 and 2021 (a) intradune depression, (b) Hulene -B dump (c) Julius Nyerere Avenue ......... 43
Figure 3.4 (a) electrode placement process, (b) cable submerged in the surface leachates of the reception basin and
(c) cables between the pikes of the reception basin................................................................................................... 44
Figure 3.5 (a) arrangement of three cables in the first measurement and (b) cable connection .................................................. 45
Figure 3.6 (a) and (b) Arrangement of the cable reels in the second reading station ................................................................... 46
Figure 3.7 (a) and (b): Arrangement of the cable drums in the third reading station .................................................................... 46
Figure 3.8 (a) Section south of profile-1 2020 (b) Section south of profile-1 2021. ...................................................................... 47
Figure 3.9 (a) Section West of profile-2 2021 (b) Section West of profile-2 2021......................................................................... 48
Figure 3.10 (a) South of profile 3 (b) North of profile. ..................................................................................................................... 48
Figure 3.11 (a) South-West of profile-4 (b) North-West of profile-4. ............................................................................................... 49
Figure 3.12 (a) Profile-5 layout (b) Profile-5 environmental context. .............................................................................................. 49
Figure 3.13 (a) Profile 6 layout (b) Profile 6 environmental context................................................................................................ 50
Figure 3.14 Arrangement of the blocks that loosely link to the distribution of data points in the pseudo selection. ........................ 51
Figure 3.15 Arrangement of the blocks that loosely link to the distribution of data points in the pseudosecction .......................... 51
Figure 3.16 Example of noise corrections: (a) profile with overlapping points (b) overlapping point correction. ............................ 52
Figure 3.17 Optimization of the profiles by the smoothing method: (a) Calculated apparent resistivity
(b) Calculated resistivity (c) Final smoothed model. ................................................................................................... 53
Figure 3.18 (a) Water collection - South well (b) Water collection - North well; (c) Chemical analysis process
(d) phosphate analysis. .............................................................................................................................................. 56
Figure 3.19 Soil sampling area in the surrounding of the dump (yellow) , and background samples (green). ............................... 57
Figure 3.20 soils sampling and drying process: (a) Soil collection 2020, (b) Soil collection 2021
(c) Soil drying in UPM laboratories. ............................................................................................................................ 57
Figure 3.21 Sampling points for soil, rhizosphere, plants, and water (R)-Rhizosphere and (P) Plants (edible plants) ................... 58
Figure 3.22 (a) vegetable variety (b) pumpkin leaves (c) cabbage................................................................................................. 59
Figure 3.23 (a) south well, (b) north well, (c) stream water ............................................................................................................ 59
Figure 3.24 Sampling locations (yellow dots) and wind directions. Red dashed - Hulene-B dump ................................................ 60
Figure 3.25 Road dust collection (a) sample 3, (b) sample 6, (c) sample 5, (d) sample 9. ............................................................. 61
Figure 3.26 Methodological summary of soil, rhizosphere, and dust analysis ................................................................................ 62
Figure 3.27 pH and EC measurement............................................................................................................................................ 63
Figure 3.28 Identifying colors using the Munsell Soil Chart ............................................................................................................ 64
Figure 3.29 Equipment used in the chemical analysis of water (a) Reagent kit (b) Chemical analysis process. ............................ 66
Figure 4.1 Electrical resistivity model of profile 1, 2020, (a) and 2021 (b). ................................................................................... 84
Figure 4.2 Geophysical surveys: (a) profile 1 in 2020 the arrow indicates the surface leachate concentration ditch
parallel to the profile 1; (b) drainage ditch with uninsulated surface leachate 2021; (c) southern section
of profile 1 in 2021....................................................................................................................................................... 85
Figure 4.3 Electrical resistivity model of profile 2, 2020 (a) and 2021 (b). .................................................................................... 86
Figure 4.4 Electrical resistivity model of profile 3 in 2020 (a) and 2021 (b). ................................................................................. 88
Figure 4.5 Electrical resistivity model of profile 4 in 2020 (a) and 2021 (b). ................................................................................. 89
XVII

Figure 4.6 Possible flow directions of contamination plumes 2020–2021, modified from Bernardo et al. (2022a) ....................... 91
Figure 5.1 Calibrated color scale showing the resistivity range of materials and their characteristics.......................................... 94
Figure 5.2 Profiles of the variation of the electrical resistivity in the study area............................................................................ 97
Figure 5.3 Electrical resistivity values (minimum, maximum, mean, standard deviation Ω.m) and modified DRASTIC
Index. Maximum value of profile 5 (477.1 Ω.m) was set to (200) to maintain the scale within limits of
perceived index. ......................................................................................................................................................... 102
Figure 5.4 Spatial projection of the modified DRASTIC in surroundings of the Hulene -B dump: dashed in white
represents low vulnerability and in black high vulnerability. ....................................................................................... 104
Figure 6.1 Soil texture classification of studied soils (blue) and background (yellow) samples ................................................... 108
Figure 6.2 Granulometric spatial distribution of sand, silt and clay fractions of the studied soils (in %; n = 71) .......................... 109
Figure 6.3 Soil samples pH in sand and silt fractions. ................................................................................................................. 111
Figure 6.4 Soil samples EC in sand and silt fractions. ............................................................................................................... 112
Figure 6.5 Soil samples OM content, color, and moisture content. ............................................................................................. 113
Figure 6.6 PCA variables distributed by component. .................................................................................................................. 115
Figure 6.7 (a) Leachate drainage channels without isolation, (b)dispersed leachate, (c) uncontrolled leachate flow.................. 117
Figure 7.1 Soil texture classification of studied soils (blue) and background (yellow) sample..................................................... 120
Figure 7.2 Soil samples pH in (a) sand and (b) silt fractions on 2021 collected samples ............................................................ 122
Figure 7.3 (a) Soil samples EC in sand and (b) silt fractions ....................................................................................................... 124
Figure 7.4 (a) Soil samples OM content, (b) color, and (c) moisture content .............................................................................. 125
Figure 7.5 Principal component analysis variables distributed by compon ................................................................................. 127
Figure 8.1 Mineral phases identified on the studied soils and local background samples. Quartz was not considered. ............. 133
Figure 8.2 Hierarchical cluster analysis of the selected PTEs. .................................................................................................... 138
Figure 8.3 Potential ecological risk index (PERI), pollution load index (PLI), and soil Nemerow index calculated with local
background soils and with world soils proposed by Reimann and Caritat (1998). ..................................................... 139
Figure 8.4 Hazard index for systemic toxicity for children and adults, as well as adjusted for both children and adults. ............. 143
Figure 9.1 Rhizosphere mineral content ..................................................................................................................................... 148
Figure 9.2 Rhizosphere soil hazard quotient (HQ) by ingestion and dermal contact for children (in %) ...................................... 150
Figure 9.3 Transfer factor (TF) from soil to plants ...................................................................................................................... .154
Figure 9.4 Targeted hazard quotient (THQ) in edible plants ....................................................................................................... 156
Figure 9.5 Hazard index in edible plants ..................................................................................................................................... 156
Figure 9.6 Systemic toxicity of wells water samples.................................................................................................................... 157
Figure 10.1 Dust samples identified mineral phases of the < 63 and < 2000 µm fractions relative distribution
(m-m – magnetite maghemite). *Quartz is not projected ............................................................................................ 162
Figure 10.2 Dust samples Nemerow Index using soils background (PNbkg), and soils surrounding Hulene-B dump
(PNHB) mean concentration ..................................................................................................................................... 165
Figure 10.3 Relative distribution of elemental contribution for Nemerow Index using soils background (PNbkg), and soils
surrounding Hulene-B dump (PNHB) mean concentration (Bernardo et al., 2022) .................................................... 166
Figure 10.4 Dust samples, fraction < 63 µm, hazard quotient by ingestion (HQing) in children and adjusted for children
and adults by potentially toxic elements and their sum .............................................................................................. 167
Figure 10.5 Dust samples Risk by Pb, Ni and combination of both ............................................................................................... 168
XVIII

List of tables
Table 2.1 Tectonic structural-magmatic domains of Africa ......................................................................................................... 19
Table 3.1 Applicability of some geophysical techniques to different environmental problems .................................................... 39
Table 3.2 DRASTIC parameters ................................................................................................................................................. 55
Table 3.3 Monomial potential ecological risk factor (EF) and Potential ecological risk index (PERI) classification levels ........... 70
Table 3.4 Landfill risk index assessment .................................................................................................................................... 73
Table 3.5 Criteria for assigning weights and ranking .................................................................................................................. 74
Table 3.6 Variables used to assess human health risks ............................................................................................................. 76
Table 5.1 Parameter values considered in the DRASTIC Index ................................................................................................. 99
Table 5.2 Mean, maximum, minimum, standard deviation of resistivity values (Ω.m) and DRASTIC index ............................... 100
Table 5.3 VData validation of electrical resistivity and modified DRASTIC ................................................................................ 104
Table 6.1 Granulometric statistical information (in %) ................................................................................................................ 108
Table 6.2 Spearman correlation ................................................................................................................................................. 115
Table 6.3 Risk index (Ip) of the studied area.............................................................................................................................. 116
Table 7.1 Granulometric statistical information (in %) ................................................................................................................ 120
Table 7.2 Soil physical parameters of background (Bkg) 2020 and 2021 samples .................................................................... 120
Table 7.3 Spearman correlation ................................................................................................................................................. 127
Table 7.4 Contamination risk of surface water around Hulene-B waste dump ........................................................................... 131
Table 8.1 Descriptive statistics of studied soil and local background samples (in mg/kg) .......................................................... 134
Table 8.2 Spearman correlation between the selected PTEs of the soil sample ....................................................................... 136
Table 9.1 Rhizosphere Proprieties ............................................................................................................................................. 145
Table 9.2 Chemical composition of rhizosphere (R1 to R4), background (Bkg), and world soils ............................................... 147
Table 9.3 Pollution index considering background (PIbkg) and world soils (PIW) and pollution load index for background
soils (PLIbkg) and world soils (PLIW) of the studied soil samples ............................................................................. 148
Table 9.4 Water samples pH, Pb, Hg and Zn concentration and reference values (in µg/l) ....................................................... 150
Table 9.5 Pollution index (Pi) of groundwater and surface water ............................................................................................... 151
Table 9.6 Chemical analysis of potentially toxic elements in edible plants ................................................................................ 152
Table 9.7 Daily elemental intake (DIM), and hazard risk index (HRI)......................................................................................... 154
Table 10.1 Dust samples < 63 and 2000 µm fractions pH, electrical conductivity (EC; μS/cm) and color ................................... 159
Table 10.2 Dust samples granulometric distribution (in %) .......................................................................................................... 160
Table 10.3 Dust samples potentially toxic elements chemical composition of the < 63 and 2000 µm fractions (in mg/kg) .......... 163
Part I approach
2
3

1 Introduction
In recent decades waste production increased in quantity and diversity worldwide as a
result of population, and economic fast growth (Morita et al., 2020; Sharma et al., 2021).
Global annual production of municipal solid waste (MSW) in 2025 is expected to reach
~2.2 billion metric tons (Mor et al., 2023). This fact implies many challenges in defining
and managing the final disposal sites for municipal solid waste (Lashen et al., 2022;
Hoang et al., 2021). It was estimated that ~33 % municipal solid waste produced
worldwide is disposed in inappropriate places (Khan et al., 2022). In developing
countries, solid waste disposal becomes more challenging due to the lack of adequate
disposal infrastructures, which often leads to waste deposition in unprepared landfills
(Iddrisu et al., 2021; Wijekoon et al., 2022). Therefore, the areas where the waste is
deposited and their surroundings are considered with high risk of environmental
contamination (Morita et al., 2020).

Morita et al. (2021) estimated that open dumpsites will account for 10 % of global
greenhouse gas emissions, by 2025. Municipal solid waste produces leachates, a highly
contaminated liquid containing high amounts of inorganic ions, organic compounds, and
potentially toxic elements, such as heavy metals and ammonia (Koda et al. 2017; El
Mouine et al., 2021). Leachates are commonly mobilized to the surrounding environment
(e.g., soils, groundwater and surface/irrigation waters, edible plants), causing
contamination and endangering the health of the surrounding population (Aydi et al.
2020; El Naqa, 2004; Fatoba et al. 2021; Igboama et al. 2022; Jayawardhana et al. 2016;
Mester et al. 2022). Leachates dispersion vary with climatic seasons (Zaki et al.; 2022).
Nilam et al. (2016) and Wijekoon et al. (2022), suggested that recent dumps produce
more leachates, especially during hot and rainy season, due to the increased
decomposition of solid waste with higher temperatures, causing dispersion and
contamination in the surrounding environment. Leachate metal ions content vary
according to the type of waste and the environmental conditions (Nilam et al., 2016), in
general, being significantly higher in scarce water conditions (Githaiga et al. 2021; Siddiqi
et al., 2022; Tang et al. 2021).

Solid waste heterogeneity contributes to potentially toxic elements (PTEs) concentration,


being considered as a major threat to soil and groundwater quality in urban areas (El
Naqa, 2004; Lashen et al., 2022). Heavy metals are among the most hazardous soil
pollutants, being highly reactive; when organisms are unable to eliminate them
chemically, being retained in the ecosystem, thus representing a potential risk (Breus et
4

al., 2022; Singh et al., 2022). The dynamics of environmental contamination around
landfills is strongly influenced by soil characteristics (Ihedioha et al., 2017). Previous
studies considered soil as the first environmental component to be impacted by
contamination caused by leachate (Xiao et al., 2021). Soils act as a contaminant filter
and can control elemental transport, to the hydrosphere, atmosphere, and biota (Hasan
et al. 2021; Zaini et al. 2022). Studies reported soil contamination near dumpsites by As,
Hg, and Al in Italy, (Meloni et al., 2021); Cu, Pb, Co, Mn, and Hg in in China (Li et al.,
2022); As, Cd, Pb, and Cr in Malaysia (Hussein et al., 2021); Pb, Cr, Ni, Cd and As in
Nigeria (Audu et al., 2021).

Ground and surface water are also vulnerable to contamination by leachate from landfills
(El Mouine et al. 2021; Ozbay et al. 2021; Soumahoro et al. 2021). Groundwater
vulnerability depends not only on its flow system properties but also on contaminant
sources proximity, and contaminant characteristics, among other factors. Parvin et al.
(2021) recently showed high levels of groundwater contamination in areas near dumps
in Bangladesh. The same was demonstrated Brahmi et al. (2021) when assessed
groundwater and soil pollution by leachate in Tebessa municipality, and by Ololade et al.
(2019) by describing groundwater contamination by leachate mechanism on a landfill
site in Bloemfontein, Nigeria.

Another mechanism of environmental contamination by landfills is the open air burning


of wastes (Soubra et al., 2021). Studies demonstrated that waste (electronic, medical,
industrial, commercial, and domestic) burn in landfills has been a source of dispersion of
contaminated ashes into the surrounding ecosystems (Itai et al. 2014; Ferronato et al.
2019; Li et al. 2022). Narayana, (2009) and Nanda et al. (2021) showed the influence of
waste incineration on air quality in the surrounding area of waste dump in India, and
China. Studies by Tao et al. (2020) and Ruengruehan et al. (2021) revealed the influence
of waste burning on mercury contamination of groundwater.

In developing countries, where there is poor monitoring of groundwater for human


consumption, the health risks are significant (Boateng et al. 2019; Ahmad et al. 2021;
Wang et al. 2021; Essien et al. 2022). Edible plants cultivated in areas around dumpsites
are also impacted by contaminants (Guadie et al., 2021). Sharma et al. (2021), and
Parvin et al. (2021) demonstrated the influence of landfills on the contamination of
agricultural fields with various crops in the vicinity of landfills in different parts of the world.

The different mechanisms of environmental contamination by dumpsites have led


several authors to consider dumpsites a complex heterogeneous source of
5

contamination in urban ecosystems (Nta et al. 2020; Koliyabandara et al. 2020). Soil,
groundwater, irrigation water, edible plants and particulate matter contaminated by
dumps constitute a potential health human risk (Audu et al. 2021; Chaudhary et al. 2021;
Balasooriya et al., 2021). Studies associated the contamination of the surrounding areas
of dumps with the degradation of human health of the surrounding population (Githaiga
et al. 2021; Soumahoro et al. 2021). A study near a dumpsite in Kenya suggested that
high soil PTEs levels (Pb, Hg Cd, Cu, and Cr) pose a risk to children and adolescents
living and studying near the dumpsite, that presented a high incidence of respiratory
diseases and blood Pb levels, with blood cell aberrations (Kimani, 2012). Morita et al.
(2020), and Wang et al. (2022) systematized data from waste dumps in developing
countries and concluded that contamination levels caused by PTEs posed potential
health risk. Essien et al. (2022) showed that soil and groundwater high levels of heavy-
metal contamination around Udoyo dumpsite, in Nigeria, constituted a risk to human
health. Plants grown and consumed in the surroundings of dumps, were found to pose
a high health risk related to contamination by As, Cd, Cu, Hg, Pb, Mg, and Zn (Parvin et
al., 2021; Thongyuan et al., 2021). Consumption of these contaminated plants can
induce different diseases, given the PTEs toxicity of (Sharma et al., 2021).

In Mozambique, Tvedten et al. (2018), Matsinhe et al. (2020), and Langa et al. (2021),
refer that the production of urban solid waste has been increasing as a result of fast
demographic growth and increased consumption. This situation was combined with
deficiencies in urban solid waste management system (Ferrão, 2006; Langa et al., 2021).
A reduction in the useful space available for waste disposal and the low frequency of
waste collection is needed, due to an accumulation of these materials with all the
consequences that this entails, such as, a direct threat to public health from pathogens,
indirect damage to health, pollution of the air, soil and groundwater (Serra, 2012; Langa
et al. 2021). Vicente et al. (2006), and FAO (2018) points out that urban soil
contamination in Mozambique is mostly a result of poor disposal of solid waste and
untreated wastewater, which carry pollutants to the ecosystems and food chain with
serious consequences to human health.

Previous studies on groundwater in Mozambique, particularly in Maputo city, considered


it vulnerable to contamination due to the hydrogeological context characterized by
shallow level and high urban growth without adequate planning and sanitation systems
(Momade et al. 1996; Nogueira et al. 2019; Cendón et al. 2020). Vicente, (2011), and
Koliyabandara et al. (2020) reported risk of contamination of surface and groundwater
by improper disposal of solid waste. In Maputo city (capital of Mozambique) it is
6

estimated that on average between 1250 and 1600 tonnes of different waste are
produced every day (Serra, 2012; CMCM, 2020). Mertanen et al. (2013) reported that
collected waste contains 68 % organic matter, 12 % paper and cardboard, and 10 %
plastics. Hospital, electronic and commercial waste is also dumped (Matsinhe et al.,
2020).

Hulene-B waste dump located in the northern direction of Maputo city is the largest waste
dump in the country, with daily deposit of about ~1000 tons (CMCM, 2018). The
surrounding area is densely populated, with ~75,500 residents (Hulene-B ~49,000, and
Laulane ~26,000 inhabitants). Studies by Serra (2012), Mertanen et al. (2013), Buque et
al. (2015), Tvedten et al. (2018), Matsinhe et al. (2020), and Langa et al. (2021) reported
possible environmental contamination associated with Hulene-B dump. However,
quantitative studies are lacking. Integrated use of geophysical and geochemical methods
is recommended to assess environmental and human health impacts associated with
waste dumps, (Brahmi et al., 2021). Geophysical prospection (electrical resistivity) is
widely used to study the influence of dumpsites on underground environment (Koda et
al. 2017; Arifin et al. 2021; El Mouine et al. 2021; Brahmi et al. 2021). Electrical resistivity
has been applied due to the non-invasive nature of data acquisition and processing (Lau
et al., 2019). Bichet et al. (2016), Kayode et al. (2019), Udosen (2021) demonstrated
that electrical resistivity is effective in locating buried hazardous waste and identifying
contamination plumes resulting from leachate flow. Geochemistry is used to study
chemical elements content in different media samples (Ávila et al., 2017; Khattak et al.,
2021). Since the 1970s several studies have been carried out suggesting that urbanized
areas or areas with strong anthrogenic influence, chemical elements content reflect an
enrichment of heavy metals associated with human activities (Ferreira et al. 2021;
Hussein et al. 2021).

One common challenge in soil contamination studies is to adopt a comparative


background reference in order to identify outliers in both soil resistivity and geochemical
behavior by means of a local defined threshold (Reimann et al., 2005). Aboubakar et al.
(2021), and Ferreira et al. (2021) defined background as a relative measure to be used
in distinguishing between natural concentrations of a given element and concentrations
influenced by anthropic activities.

The present research, was focused on the study of the impact of Hulene-B waste dump
on the surrounding environment and human health, integrating geophysical and
geochemical methods, where soils, groundwater and surface water/irrigation, edible
plants were studied (Fig.1.1). Three non-impacted areas were selected as reference,
7

one for electrical resistivity (north of the dump) and two for soils (north and east of the
dump).

Figure 1.1. graphical abstract of the study

1.1 Thesis presentation


This thesis presents is composed of three sections:

Section 1: Chapter (1) Introduction and research perspectives; Chapter (2)


Characterization of the study area, that includes location of the study area,
geomorphology, geology, hydrogeology, soils, flora and fauna, history of the dump,
socio-environmental context, governmental initiatives of mitigation measures, socio-
environmental context of the studied areas; Chapter (3) Materials and Methods
describing planning, execution, preparation, laboratorial analysis and data processing,
both of geophysical electrical resistivity, and physical, mineralogical, and geochemical
studies of soils, road dust, superficial and groundwater, rhizosphere soils, and edible
plants.

Section 2: Results and Discussion, that includes scientific papers published and
submitted in peer-reviewed journals of indexed international journals. This section
8

consists of six sub-chapters: Chapter (4) Paper 1 - Characterization of the Dynamics of


Leachate Contamination Plumes in the Surroundings of the Hulene-B Waste Dump in
Maputo, Mozambique (published); Chapter (5) Paper 2 - Integration of Electrical
Resistivity and Modified DRASTIC Model to Assess Groundwater Vulnerability in the
Surrounding Area of Hulene-B Waste Dump, Maputo, Mozambique (published); Chapter
(6) Paper 3 - Soil properties and environmental risk assessment of soils in the
surrounding area of Hulene-B waste dump, Maputo (Mozambique) (published); Chapter
(7) Paper 4 - Soils in the surroundings of Hulene-B dump, Maputo (Mozambique) –
properties temporal variation. (Submitted); Chapter (8) Paper 5 Soil Risk Assessment in
the Surrounding Area of Hulene-B Waste Dump, Maputo (Mozambique) (published);
Chapter (9) Paper 6 - The contribution of Hulene-B waste dump (Maputo, Mozambique)
for the contamination of stream and ground waters, edible plants, and rhizosphere soils
(in progress); Chapter (10) Paper 7 - Maputo urban area (Mozambique) road dust
characterization and risk assessment (Submitted).

Section 3: Chapter (11) Conclusions, with concluding notes of the work and future
research to be developed. Chapter (12) with references, and Chapter (13) appendixes.

The published and in-publication articles selected to compose the thesis are as follows:

1. Bernardo, B.; Candeias, C.; Rocha, F. (2022). Characterization of the Dynamics of


Leachate Contamination Plumes in the Surroundings of the Hulene-B Waste Dump
in Maputo, Mozambique. Environments, 9, 19.
https://doi.org/10.3390/environments9020019
2. Bernardo, B.; Candeias, C.; Rocha, F. (2022). Integration of Electrical Resistivity
and Modified DRASTIC Model to Assess Groundwater Vulnerability in the
Surrounding Area of Hulene-B Waste Dump, Maputo, Mozambique. Water , 14,
1746. https://doi.org/10.3390/w14111746
3. Bernardo, B., Candeias, C., Rocha, F. (2022). Soil properties and environmental risk
assessment of soils in the surrounding area of Hulene-B waste dump, Maputo
(Mozambique). Environ Earth Sci 81, 542 https://doi.org/10.1007/s12665-022-
10672-7
4. Bernardo, B.; Candeias, C.; Rocha, F. (2022). Soils in the surroundings of Hulene-
B dump, Maputo (Mozambique) – Temporal variation of properties and risk
assessment. submitted.
5. Bernardo, B.; Candeias, C.; Rocha, F. (2022). Soil Risk Assessment in the
Surrounding Area of Hulene-B Waste Dump, Maputo
9

(Mozambique). Geosciences, 12, 290.


https://doi.org/10.3390/geosciences12080290
6. Bernardo, B.; Candeias, C.; Rocha, F. (2023). The Contribution of the Hulene-B
Waste Dump (Maputo, Mozambique) to the Contamination of Rhizosphere Soils,
Edible Plants, Stream Waters, and Groundwaters. Environments, 10, 45.
https://doi.org/10.3390/environments10030045
7. Bernardo, B.; Candeias, C.; Rocha, F. (2023). Maputo urban area (Mozambique)
road dust characterization and risk assessment. submitted.
10
11

2 Study Area
2.1 Location of Mozambique
The Republic of Mozambique is located in the Southern Hemisphere, between parallels
10º 27’ South and 26˚52' South latitude. The country has boundaries with Tanzania to
the north, Malawi, Zambia, Zimbabwe, South Africa, and Swaziland to the west and
South Africa to the south (Fig.2. 1). The continental surface of Mozambique is of 786 380
km2, corresponding to ~2.6% of the African continent surface (Muchangos, 1999).

The continental shelf of Mozambique, whose limit is fixed at 200 miles from the coastline,
has an extension of 120 000 km2, or 0.24% of the approximately 30 Mkm2 of the Indian
Ocean surface. The country maximum altitude above sea level is 2 436 m at Mount
Binga, in the Chimanimani chain in Manica, while the greatest continental depth is
recorded at Lake Niassa, at 706 m below sea level (Muchangos, 1999).

Figure 2.1 Location of Mozambique, adapt.


https://www.nationsonline.org/oneworld/map/mozambique_map.
12

2.2 Location Hulene-B waste dump


The capital of Mozambique, Maputo City, with an average altitude of 47 m, is located
west of Maputo Bay, to where the Tembe, Umbeluzi, Matola and Infulene rivers flow. Its
limits correspond to latitudes 25º 49' 09" S (northern extremity) and 26º 05' 23" S
(southern extremity) and longitudes 33º 00' 00" E (eastern extremity - island of Inhaca)
and 32º 26' 15" E (western extremity) (Fig. 2.2). According to the current administrative
division, the Municipality of Maputo is divided into 7 districts (Autarchic Administrative
Units) and 61 neighborhoods with an estimated population of 1 101 170 inhabitants (INE,
2020). To the north, Maputo city borders the district of Marracuene tto the north and west
the municipality of Matola. In particular, the area covered by the study, Hulene-B waste
dump, is located east to Maputo International Airport, and Hulene-B and Laulane
neighborhoods in the center-west of Kamavota municipal district in Maputo municipality
(Fig.2.2), with about 48 717 and 27 996 inhabitants, respectively. According to the spatial
differentiation of the urban space occupation structure, the study area is classified as
suburban (INE, 2020). Most of the population is young and involved in the collection and
resale of solid waste (Tvedten et al., 2018).

Figure 2.2. Location of Maputo City and Hulene-B waste dump Adapt. Google Earth Pro.
13

2.3 Climate of the study area


Maputo City is inserted in the tropical anticyclones zone which, forms the African
Continent depression of thermal origin (Muchangos, 1999). The climate is predominantly
subtropical with an average annual precipitation of ~789 mm, with two climatic seasons:
(a) hot (average 25 ºC) and rainy period, from December to March, with > 60% of the
annual precipitation, with a peak in January (~125 mm), and (b) dry and cold period, from
April to September, with lower temperatures in June and July (average 21 ºC), and
scarce precipitation, whose minimum values were recorded in August (~12 mm) (Fig.2.3)
(CIAT, 2017; INAM, 2020). The prevailing winds are SE (Fig. 2.4) (Muchangos, 1999).

Maputo City
200 80
Precipitation mm)

Temperature (ºC)
150 60

100 40

50 20

0 0
Jan Feb Mar Apr May Jun Jul Aug Sep Out Nov Dec

mm °C

Figure 2.3. Precipitation and average monthly temperatures from 1987-2017 (adapt. INAM,
2020).
14

Figure 2.4 wind direction along the year (adap.


https://www.meteoblue.com/pt/tempo/historyclimate/climatemodelled/maputo_mo%C3
%A7ambique_1040652)

2.4 Geomorphology
2.4.1 Morphology of Mozambique

The relief of Mozambique is arranged in the form of an amphitheater where one can
distinguish a mountain zone in the west, which descends in flattened steps to the coastal
plain in the east (fig. 2.3) (Muchangos, 1999). About 44% of the Mozambican territory is
composed of plains, with maximum altitudes of 200 m. The largest extension of
Mozambique flat surfaces, with altitudes between 1000 and 2000 m, is in the northern
half of Mozambique, constituting the Mozambican Plateau (Fig. 2.4). Two plateau zones
can be distinguished in Mozambique, (a) the middle plateau, with altitudes between 200
and 500 m, , represented north of the 17th parallel S, in the provinces of Cabo Delgado,
Nampula and in the interior of Inhambane; and (b) altiplanitic, with altitudes above 500
m, which include landforms with altitudes above 1000 m, are not very extensive (5%)
and do not constitute continuous bands, similarly to the plateau. They occur mainly north
of the 21st parallel S, in the provinces of Niassa, Zambézia, Tete and Manica. The
southern region of Mozambique, where the present study area is included, is of plains
type, interspersed by coastal and inland dunes (Fig. 2.5).
15

Figure 2.5 Geomorphology of Mozambique (adap.


https://www.mapsland.com/maps/africa/mozambique/detailed-physical-map-of-
mozambique.jpg)

2.4.2 Morphology of Maputo City

Maputo city is part of the country coastal area, inserted in two zones, the first one
consisting of recent dunes and alluvial fans, and another one consisting of older and
fixed inland dunes. In particular this zone, reveal some flattening, where sandy-clay soils
have developed, to which Barradas (1962) attributed Mananga soils classification (INIA,
1993). Momade et al. (1996) recognized the existence of four morpho-structural plateaus
namely: (i) the coastal accumulation zone corresponding to the beach and tidal deposits;
(ii) the coastal zone, sloping towards the sea with a maximum altitude of 8 m, consisting
of dunes and alluvium; (iii) the 40-50 m platform, slightly sloping towards the west, where
predominantly degraded inland fixed dunes and sand sheets occur, and (iv) the Maputo
hill, with 50 to 60 m, which constitutes the residual relief resistant to erosion probably
related to the neo-tectonism phenomenon
16

2.4.3 Morphology of the study area

From a morphological point of view, the study area belongs to the dune system of the
southern Save sedimentary basin, where the dunes have visible expression and are of
the longitudinal type, with crests systematically oriented towards the north-northwest,
and intradune depressions oriented in the same direction, that reach some km in length
filled with white very fine sand (Momade et al., 1996; Vicente et al., 2006). The dunes
reach a height of up 50 m, being parallel to the red tip formation, suggesting that were
affected by a similar wind system, and that the reddish intercalations of the Malhazine
formation could result from the removal of the red tip formation (Momade et al., 1996).

The Hulene-B waste dump is located in the intradune depression and on the dune slope
to the east (Bernardo et al., 2022). The eastern limit corresponds to the high elevations
with altitude ranging from 52 to 54 m, while the western limit on the low elevations,
ranging 34 to 32 m (Fig. 2.6).

Figure 2.6. Topography of the study area - Projected from USGS/Maputo MDE

2.5 Geological context


2.5.1 Regional geology

The African continent is composed of a mosaic of archaic mobile cratons and belts,
amalgamated by elongated fold belts of Proterozoic-Cambrian age (Schlüter, 2008).
Cratons are stable continental areas which have not been deformed for a long period of
time, since the Precambrian or early Palaeozoic, and in which post-Palaeozoic
17

sediments, if present, are not deformed. Similarly, seismic or volcanic disturbances, if


any, are negligible. These zones have emerged and remained submerged since the
Precambrian, more than six hundred million years ago. Thus, cratons are vestiges of
Pangea. They are characterized by an intensely eroded continental crust and are always
located within the same plate.

The main African cratons, the following: (i) West African Craton; (ii) Central African
Craton, accompanied to the east by the, (iii) East African Craton, and (iv) South African
Craton (fig.2.7 and 8). The cratons are separated by mobilized zones consisting of
younger Precambrian folded systems that were often affected by orogenic activity in the
late Precambrian (Kampunzu et al., 1991). Some of these mobilized zones are chains
that have been folded, such as the Mozambique belt or Katanga or Damara, but more
often they are older zones that have been reworked and rejuvenated - as witnessed by
radiometric dating from the mobilized zone extending from Ahaggar (Hoggar) to
Dahomey (Schlüter, 2006).

Dirks et al. (2003) point out that most African cratons are covered by undeformed
sediments and associated extrusive rocks of Neoproterozoic, Late Carboniferous to
Early Jurassic and Cretaceous - Quaternary age (Table 2.1). The younger fold belts,
deformed during the Hercynian and Alpine orogenies, show only local extension.
Southern Africa (where Mozambique is included) is greatly influenced by the Karoo
system.
18

Figure 2.7. African Craton (source: Key, 1992 and Schlüter, 2008).

Figure 2.8. Archean cratonic nucleic (source: Key,1992 and Schlüter, 2006).
19

Table 2.1 Tectonic structural-magmatic domains of Africa (adap. Dirks et al., 2003 and Schlüter,
2008)

Event/System Era/Period (Ma) Event/System Event/System


Era/Period Age (Ma) Era/Period (Ma)
Principal phase Neogene Present 23 – 0
East African Rift Initial phase Cretaceous 140 – 23
System Palaeogene
Karoo Upper Carboniferous - 318 – 180
Gondwanides terrain Lower Jurassic
Post Pan-African < 542
platforms

Pan-African Orogeny Neoproterozoic 750 – 490


Initial Pan-African Neoproterozoic 900 – 700
Pan-African
basins
Kibarian/Irumides/Gr Mesoproterozoic 1450 – 900
envillian
Ubendian/Usagarian Late fase Paleoproterozoic ~ 1860
Ubendian/Usagarian Initial phase 2100 – 2025
Cratons (granite and greenstone terrain) and Archaic 3800 – 2500
mobile belts

2.5.2 2.5.2 African Karoo system

The Karoo basins of Africa are positioned in the central and southern regions of the
continent (Fig. 2.9) and preserve the record of a special period in Earth's history when
the supercontinent Pangea reached its greatest extent during the late Palaeozoic and
early Mesozoic. The term "Karoo" was extrapolated from the main Karoo Basin in South
Africa to describe the sedimentary infill of all other basins of similar age that occurred
throughout Gondwana. The onset of sedimentation of this Karoo depositional sequence
is generally placed in the Late Carboniferous, around 300 Ma ago, followed by a major
tectonic reversal event along the southernmost margin of the supercontinent Pangea.
Sedimentation of the Karoo in Gondwana continued until the fragmentation of the
supercontinent in the middle Jurassic (about 183 Ma) when the accumulation of
sediments was replaced by the installation of a large igneous province. The upper part
of the Karoo sequence was eroded during the post-Gondwana period, so the youngest
preserved Karoo deposits generally range from the Triassic to Middle Jurassic
(Catuneanu et al., 2005)
20

Figure 2.9. Distribution of the Karoo basins in central and southern Africa (source: Schlüter,
2006)

2.5.3 Mozambique Geology

Mozambique can be subdivided into two geological-structural regions:(a) Precambrian


terrains with a surface area of about 534 000 km2 (b) a Phanerozoic region with about
237 000 km2 (Afonso, 1978). In Mozambique, the geological differences are very large
between the north and the south of the country, being the north fundamentally
Proterozoic and the south Phanerozoic (Fig. 2.10) (Vasconcelos, 2014). The general
geological setting of Mozambique comprises: Precambrian Formations of the Basement
Complex, Karoo Sedimentary Formations, Post-Karoo intrusive and extrusive rocks, and
Meso-Cenozoic Sedimentary Formations (Afonso, 1978; Baruya et al., 2013).

The Base Complex is located in the western part of the central region of the country and
in almost the entire region north of the Zambezi River. It comprises formations of the
Zimbabwe Craton and formations of the Mozambican belt. The older formations are
mainly composed of green schists with interstratifications of quartzites, limestones,
greywacke, and conglomerates. The Mozambican belt is formed predominantly by rocks
21

of high degree of metamorphism such as gneisses, granite-gneiss-migmatite, alternating


with metasediments, charnockite and basic and intermediate rocks of the Tete Gabbro
Complex. The predominant structural orientation of the rocks of the Mozambique belt is
N-S (Baruya et al., 2013; Vasconcelos, 2014).

The Karoo sedimentary basins (Middle Zambezi and Maniamba Basin). The two basins
are surrounded by base complex formations and form flat depressions with rare
individual hills and terraces, relicts of ancient erosion cycles. The Middle Zambezi basin
is composed of several sloping blocks, in which the sedimentary sequence can reach
over 3,000m thick. The Lower Karoo starts with bedded conglomerates and tills followed
by pelites, argillites and siltstones (Dwycka and Ecca Series) and coal beds (Afonso,
1998).

The Upper Karoo sedimentary sequence is formed of fine-textured sandstones, coarse


sandstones, fossiliferous sandstones and marls (Beaufort Series). The total sedimentary
sequence is frequently intersected by dolerite veins. From base to top, the Maniamba
Sedimentary Basin, comprises conglomeratic sandstones, pelites, carbonaceous shales
and siltstones, shales with silicified logs, shales with calcareous concretions and friable
clayey sandstones. Quaternary fine sands are common in the Middle Zambezi basin and
along the eastern and southern edges of the Maniamba basin. They comprise colluvium,
alluvial terraces, eluvium and possibly eolian sands. Post-Karoo intrusive and extrusive
rocks comprise especially basalts, rhyolites, and alkaline lavas. This volcanic sequence
was emitted from faults and fissures. The Libombos chain located in the southwestern
part of the country and extending to the Zambezi River is the most important (Afonso,
1978).

The Meso-Cenozoic sedimentary formations (Rovuma Sedimentary Basin and Southern


Mozambique Basin) are characterized by predominantly continental series of Arcosic
sandstones in the western parts of the basins and predominantly marine and transitional
series in the coastal areas. Almost 70% of the sedimentary basins are covered by loose
material. It is generally a 5 to 10 m thick meteoric alteration cover, sandy-clayey in nature,
resting on sedimentary rocks (Afonso, 1998).

The lithostratigraphic units of the Mozambican territory can be conveniently divided


between a crystalline basement of Archaic-Cambrian age and a rock cover of
Phanerozoic age. The crystalline basement comprises a heterogeneous set of
metamorphosed supracrustal paragneiss, granulites and migmatites, orthogneisses and
igneous rocks. From a geodynamic point of view, it is generally accepted that the
22

crystalline basement of Mozambique is composed of three different terrains that collided


and amalgamated during the Pan-African Orogenic Cycle (COPA). Prior to the Pan-
African amalgamation, each terrane was characterized by an individual and specific
geodynamic development. Provisionally, these terrains are referred to as the East
Gondwana Terrane, the West Gondwana Terrane and the South Gondwana Terrane
(Westerhof et al., 2008).

Figure 2.10. Geological units of Mozambique (Source: Baruya et al., 2013)

2.5.4 Geology of Maputo City

Maputo City is embedded in the southern Mozambique Meso-Cenozoic sedimentary


basin, formed by subsidence along N-S oriented deep sinks associated with the East
African rift system and along the axis of the Mozambique Channel (Vicente, 2011). The
substrate of this basin is made up of basalts and rhyolites belonging to the Karroo
deposits, of Permian to Jurassic age, filled by sedimentary strata of Tertiary and
Quaternary age, with an eastward slope and horizontal to sub-horizontal strata. The
23

following units are distributed in this basin: Inharrime Formation (Late Tinian - Oligocene)
- composed of fine sandstone, siltstones and clays impregnated with organic matter, and
intercalations of calcareous sandstones; Santiago Formation (Tsa-Miocene-Pliocene) -
composed of clayey sandstone, calcareous sandstone with ostreacullata in the upper
part.
The Maputo city area, however, is covered by the following units: (i) The Ponta Vermelha
Formation (TPv-Pliocene-Lower Pleistocene) consists in the upper part of ferruginous
red sandstone and red silty sands, which pass gracefully below to yellow and whitish
sands. (ii) Matola Formation (lower Pleistocene-QM) which consists of alternating coarse
to medium, yellow-orange and sandy-clayey sands. It is about 15 m thick; (iii) Machava
Formation (QMc- Middle Pleistocene) composed of alternations of yellowish sandy clay,
with carbonate and saliferous intercalations, of ferruginous concretions, and basal
conglomerates with artefacts; (iv) Malhazina Formation (QMa- Upper Pleistocene) - This
unit is part of the so-called interior dunes consisting of coarse to fine, poorly consolidated
sands, with reddish to white colors (Momade et al.,1996; Vicente, 2011)). This formation
is contemporaneous with the Congolote Formation(v) (QCo); (vi) Costa de Sol Formation
(QCs-Holocene) consisting of poorly consolidated sandstone, with sub-fossiliferous
marine fossils; (vii) Xefina Formation (QXf - Holocene) consisting of coastal dunes and
fossil illmenitic dune core sands. In addition to this formation in the Holocene there are
intradunal deposits (Qi), alluvial deposits (Qa) and beach sand deposits (Qm), (Fig.2.11)
(Momade et al., 1996).
24

Figure 2.11. Geology of Maputo City (adap. by Momade et al, 1996)

2.5.5 Geology of the study area

The two most important formations in this research, i.e. where the sampling area lies,
are: Ponta Vermelha Formation and Malhazine (Fig. 2.12). The Ponta Vermelha
Formation (TPv) dated between the upper Pliocene to lower Pleistocene. The unit
extends from Maputo to Marracuene, with NE-SW orientation over a 3 to 4 km band. At
the surface this unit is red colored, and poorly consolidated or loose sands may appear.
Momade et al. (1996) admits the occurrence of a conglomerate at the base of the unit.
The textural analysis evidence well calibrated material. The shape of the curve, negative
asymmetric and leptokurtic, indicates that the sediments were deposited in fluvial or eolic
environment. The chemical composition of the rocks of this formation reveals high
contents of silicon oxides [SiO2], followed by aluminum oxides [Al2O3] and iron oxides
[Fe2O3]. The fact that the unit sits directly on a marine substrate suggests that it is
possibly an aeolian mantle on a marine abrasion platform (Momade et al.,1996; Vicente
et al., 2011). The Malhazine Formation (QMa), dated to the upper Pleistocene, this unit
is part of the so-called interior dunes. It is composed of coarse to fine, poorly consolidated
sands, with whitish to reddish colors, fixed by the vegetation and as a result of successive
consolidation processes.
25

In textural terms, the fine sandy fraction (92 to 98%), very well graded, is dominant. The
shape and symmetry of the curve suggest eolian deposition. Mineralogically, quartz is
the dominant mineral (98%), with small percentages of ilmenite, leucoxene, monazite,
sillimanite, zircon and rutile appearing. Chemical analyses show high concentrations of
silicon (Si), iron (Fe) and aluminum (Al) oxides (Momade et al., 1996).

Drilling (L16 and L125) done on two geological formations around the dump revealed
that the QMa formation underlies the TPv formation, and the Santiago Formation (TSa-
Miocene-Pliocene) composed of clayey, calcareous sandstone with ostreacullata in the
upper part (Momade et al.,1996) (Fig. 2.13).

According to Barradas (1962) and Beernaert (1987), the areas covered by these two
formations are part of the dune field deposition linked to regressive phenomena,
associated with the glacial maximum of the Upper Wurm, i.e., Upper Pleistocene.

Figure 2.12. Geology of the study area: TPv Ponta Vermelha Formation and QMa
Malhazine Formation (adapt. Momade et al., 1996)
26

Figure 2.13. Cross-section of the geological formations in the study area (adapt. Momade et al.,
1996).

2.6 Soils

INIA (1993) classifies the soils of the study area as dune phase whitish/yellow sandy
soils (dAj), corresponding to the Malhazine Formation (QMa) and reddish sandy silty
soils (G) or red sandstone derived soils that gradually change to whitish yellow sands
corresponding to the Ponta Vermelha formation (TPv) (Fig. 2.14). A recently published
study on the geological and pedological resources of Maputo city, classified soils of the
two lithological units, where Hulene-B dump is located, into yellowish eolian sands with
reddish phases (Malhazine Formation) and red silty eolian sands (Ponta Vermelha
Formation) (Momade et al., 1996). Observations on site showed a strong change in the
color of the soils in the study area, often with a color that tends to be dark in areas of
accumulation of organic matter and with little influence from the leaching process. The
alteration of color and odors caused by surface contaminants that are mobilized towards
the lower areas, is noteworthy.

Studies on soil quality in urban areas (Silva et al., 2021; Shahab et al., 2022; Bisht et
al., 2022) concluded urban centers are exposed to heterogeneous contamination
sources, such as heavy metals, oils, aliphatic and aromatic hydrocarbons, chlorides,
sulphates and nitrates, due to the inadequate disposal of solid waste and combustion
gases. Wear and tear of braking systems and metal structure of motor vehicles, on major
roads and railways, are a source of lead [Pb], zinc [Zn], copper [Cu], chromium [Cr] and
27

nickel [Ni], and other substances such as nitrogen [N], sulphur [S], phosphorous [P] and
chlorine [Cl] contamination, which may affect both sides of the roads.

Preliminary studies on the levels of soils contamination around Hulene-B dump, shown
that the area receives all types of waste produced in Maputo with leachate flow resulting
from waste decomposition enriching soils (Vicente, 2011; Langa et al., 2021; Bernardo
et al., 2022).

Figure 2.14. Distribution of soils in the study area, with (dAj) the white/yellow sandy
soils dune phase, and (G) the reddish sandy silty soils (adapt. INIA, 1990)

2.7 Hydrogeology

Hulene-B hydrogeological system is part of the Tertiary - Quaternary aquifer system


(Cendón et al., 2020). The aquifer substrate is formed by a layer of clayey marl to grey
clay (Muchimbane, 2010; Nogueira et al., 2019). In the surroundings of Hulene-B dump
there is the presence of a semi-impermeable layer (clayey sands), between the fine to
coarse sand and sandstones, causing water circulation between these two sectors
(Muchimbane, 2010). There are locations where the coarse sands lie directly on top of
the clay layer, developing semi-confined conditions (Nogueira et al., 2019). The water
28

level of the shallow wells varies between 1.5 and 9.3 m deep, with an average of 3.8 m
(Cendón et al., 2020). The hydraulic conductivity has been estimated to be 1-5 m/d
(Muchimbane, 2010). Momade et al. (1996) estimated that Maputo region average
aquifer recharge is 165-185 mm/year and may decrease to 140-150 mm/year. Surface
and groundwater aquifers are recharged by precipitation, seepage from rivers, lakes, and
groundwater (Vicente et al., 2011 and Cendón et al., 2020).

Locally, at the western boundary of the dump, an intradune current is evidenced, which
is characterized by the accumulation of clastic surface waters, which are constantly
replenished by leachate from the E-W surface flow (Fig. 2.15).

Figure 2.15. (a) Surface water retained in the dune depression used for agriculture, and
(b) Surface water in the depression, dwellings area.

2.8 Flora and fauna

Maputo City is integrated in the Tongaland-Pondoland Regional Mosaic, which


comprises mostly lowlands (Diniz et al., 2013). Balcedos occur in coastal dunes and sub-
coastal regions with savannah formations of sandy soils develop with many woody
species with edible fruits, extratropical species and introduced species. In the estuary of
rivers, mangroves (rhizophora) develop. On the banks, sandy river islets and intra-dune
depressions occur communities of reed (Phragmites australis) and cattail (Typhala tifolia)
and small trees (Bandeira et al., 2014).
29

In the surroundings of the Hulene-B dump are distributed along the western boundary
the undergrowth vegetation, reed (Phragmites australis) and cattail (Typha latifolia)
(Fig.II.14), while on the eastern boundary there is cultural flora consisting of yellow
acacias (Vachellia farnesiana), cashew tree (Anacardium occidentale), Trichilia (Trichilia
emetica), mangifera (Mangifera indica), Sclerocarya (Sclerocarya caffra).

Anthropic pressure is felt in the city of Maputo, which causes the destruction of natural
vegetation and reduction of some species resulting from the historical process of use by
local people in food, construction, fuel, materials for artifacts and traditional medicine
(Diniz et al., 2013). The fauna is quite reduced, however in the surroundings of the
Hulene-B dump, birds, street dogs and rats that feed on the waste now deposited stand
out. At the western boundary, due to the seasonal accumulation of rainwater in the
intradune depression, there are records of seasonal occurrence of catfish, wild birds,
stray dogs and domesticated animals that feed on the local vegetation (goats).

(a) (b)

Figure 2.16. (a) Ground vegetation (b) Phragmites Australis

Since 2021 (second year of the present research) some areas on the western border of
the dumpsite have been converted into agricultural fields. Here, rice and vegetables are
cultivated for the subsistence of several families (Fig.2.17). However, vegetables occupy
the smallest cultivated areas. The most cultivated species are Amaranthus (Amaranthus
spinosus) in the western end of the riparian zone, sweet potato (Ipomoea batatas) in the
eastern end of the riparian zone, pumpkin (Cucurbita pepo) in the south-eastern end and
south-western cabbage (Brassica oleracea L). Rice is grown in the intradune depression
(fig. 2.17c), which serves as a livelihood for many families.
30

Figure 2.17. (a) Vegetable crop, (b) Pumpkin crop and (c) Rice crop.

2.9 Brief history of the Hulene-B Waste dump

Hulene-B dump dates back to 1973, when the community of the Hulene neighborhood
decided to reuse an abandoned white sand open- quarry for waste disposal (Ferrão,
2006). The dump was institutionalized after a short period of time and officially became
the disposal site for municipal solid waste in Maputo City (Fig. 2.18). The waste disposal
grew rapidly eastwards, outgrowing the former quarry area and settling on
heterogeneous sandy topsoil resulting from the mixture of soils from Ponta Vermelha
and Malhazine. Today, most of the dump is located in the Malhazine formation. The
Hulene-B dump is considered to be the largest in the country, with an extension of
approximately 17 hectares (Fig. 2.18).
31

Figure 2.18. Hulene-B waste dump extension

The dump is open 24 hours a day and is frequented by about 700 waste pickers
(Matsinhe et al. 2020). Since August 2006, all waste arriving at the dump has been
weighed by weighbridge. To mitigate the negative effects on the people living near the
dump as well as on the traffic, especially during the rainy season, the perimeter of the
dump has been fenced on the south and east borders with a wall. The unloading of the
waste is done without much control and with little compaction (Fig 2.19a). The height of
the waste varies between 6 and 15 m (Fig 2.19b). The leachate resulting from
decomposition of the waste is not properly controlled (Fig 2.19c).

The ground is not waterproofed and there is no cover for the waste deposited, which
allows the waste pickers access to the solid waste deposited (Palalane et al., 2008).

Serra (2012) estimates that the city of Maputo produces about 1250 tons of waste per
day, much of which ends up at the Hulene-B dump. In turn, the projections of the
Municipal Council of Maputo City indicate that in 2017 the average daily disposal of solid
waste at the Hulene-B dump stood at 1000 tons (Buque et al. 2015).

The types of waste deposited in the Hulene-B dump vary according to their origin
(CMCM, 2020). In formal city zone, the collected rubbish contains 68% organic matter,
32

13.2% paper, 9% plastics, 7.5% glass, 2.8%metals, 1.5% rubber by weight. In the
informal areas of the city, 44.2% fine fraction including debris, 37% organic matter, 6.6%
plastics, 4% paper, 2.8% glass (CMCM, 2020). Given the opening of the dumpsite
throughout the day coupled with weak enforcement, there has been deposition by
different actors of different wastes (Industrial, biomedical, commercial, etc.) (Palalane et
al., 2008).

Figure 2.19. (a) Types of waste; (b) Height of waste and (c) Uncontrolled leachate
flows at the western boundary of the dump

In the early morning of 19 February 2018, part of the waste from the western boundary
of the Hulene-B dump slid down burying about 32 houses and causing death of 17
residents (DW, 2018) (fig. 2. 20). This episode marked the beginning of the process of
compulsory removal of all inhabitants from perimeters considered at risk (~100 m) to
other areas considered safe.

Figure 2.20. (a) Houses buried by the dump collapse, (b) Roads blocked with waste
mass (adapt. DW, 2019).
33

2.10 Socio-environmental context

The Hulene-B dump is located in the suburban area of Maputo city, characterized by the
predominance of young people, who often have no guaranteed source of income (INE,
2020). Thus, the Hulene-B dumpsite is a source of income for the livelihoods of several
families who collect and resell solid waste (CMCM, 2020). It is estimated that the
dumpsite hosts 300 to 400 waste pickers per day, with women participating the most
(Matsinhe, 2020). The level of education of the waste pickers does not reach high school,
and most of them are considered illiterate. At the dump, it is possible to find orphaned
children who usually collect waste or search for food (Matsinhe, 2020). Men collect
broken appliances, pieces of Zn, Fe, Al and Cu sheets, women and children collect
plastics, cardboard, and food leftovers among others. The time of waste collection is
diverse and the practice of this activity extensive (Matsinhe, 2020).

However, there are waste pickers who have been collecting waste for more than 20
years. (Mertanen et al. 2013). Disease risks associated with the high exposure of waste
pickers are also reported, given the absence of effective protection mechanisms for
collectors (Matsinhe, 2020). In environmental terms, the Hulene-B dump is considered
one of the most significant sources of environmental contamination in Maputo city. Given
that, contamination occurs in multiple ways such as: intensive burning of waste;
migration of leachates that can contaminate soils as well as migrate to groundwater and
incorporate into the food chain. However, the heterogeneous urban character of the
environment where the landfill is located, suggests considering other sources that are
relevant to consider (Fig.2.19). Julius Nyerere Avenue, characterized by a degraded
asphalt in the axis of the dump, may be a source of Mn, Cu and Pb to be considered.
Further west of the dumpsite, the Maputo International Airport, is a factor to be
considered for geo-environmental change as it may be a source of Pb (Fig. 2.19 a and
d). These factors may be significantly altering the local environment, where strong
changes in soils, air and vegetation cover can be noticed on the western boundary, which
coincides with the lower elevations of the area, where part of the leachate from the
dumpsite accumulates superficially (Fig.2.21). The environmental reconnaissance
allowed us to define the sampling areas, as well as to identify the appropriate methods
for their analysis.
34

Figure 2.21. Environmental Context: (a) Maputo International Airport; (b) Hulene-B
intradune depression; (c) Hulene-B waste dump; (d) Julius Nyerere Avenue.

2.11 Environmental impact of the surroundings of the dump and


governmental mitigation initiatives

Studies published by Vicente et al. (2006), Serra (2012), Buque et al. (2015), CMCM,
(2020), Matsinhe et al. (2020), Langa et al. (2021); Bernardo et al. (2022a); Bernardo et
al. (2022b) denounced the socio-environmental impacts caused by the Hulene-B dump
in the surrounding areas. Vicente et al. (2006) demonstrated that the permeability of the
soil supporting the waste allows the migration of leachates produced by the
decomposition of the waste and consequently the enrichment of soils and groundwater
by potentially toxic elements [e.g., Cd, Cr, Pb, Cu and Zn], contained in the leachate
flows. In turn, Serra (2012), Mertanen et al. (2013), Langa et al. (2021) report that the
Hulene-B Dump environmentally poses a danger to public health due to the proliferation
of rats, flies, mosquitoes infesting the area and its surroundings; the spread of
wastewater, gases, and nauseating smell from the piles of rubbish deposited daily (Fig.
2.22). The dump contributes to the alteration of the environment resulting from the
unregulated incineration of waste from all parts of the capital including hospitals (Serra,
20212; Sarmento et al., 2015). The burning of waste occurs at all times of the day, which
contributes to the release of ash residues into the surrounding area.
35

The other impact and risk presented by Livaningo (2018), refers to the danger of the
proximity of the Hulene-B dump to the Maputo international airport, since the dump
attracts many birds that fly over the airport area, endangering the air safety of airplanes.

Figure 2.22. (a) wastewater from the dump, and (b) air pollution.

Given the growing concern for environmental quality, in 2020 the municipality of Maputo
began to take a series of structural measures to reduce the environmental impact of the
waste dump, which consisted of compacting waste by bulldozers, building leachate
drainage channels and channeling part of the leachate into storage tanks (Fig.2.23).

Figure 2.23. (a) Compacted waste inside the dump, (b) waste drainage channels and
(c) leachate collection tanks. Source (Bernardo, 2020 - 2021).

Recent studies published by CMCM, (2020) have demonstrated that the dumpsite is
unsustainable and recommend its closure to minimize the associated environmental
problems and reduce the associated public health risks. It was also demonstrated that
the families living around the dump are subjected to precarious subsistence conditions
and various risks due to direct contact with the waste, besides living with the high level
of violence in the area (Matsinhe, 2020). However, the possible closure of the dumpsite
36

could affect the sources of subsistence of many families who live and depend on the
collection and commercialization of waste, some with more than 20 years in this activity.
During the fieldwork, new solid waste disposal areas were found to the west of the main
dump. These areas arose spontaneously, and the deposition of waste is done in the
open along the intradune depression , which implies more contact with surface water and
potential risk of leachate migration in depth, affecting subsoil and groundwater (Mertanen
et al., 2013; Parvin et al., 2021) (fig.2.24)

Figure 2.24. (a) e (b) New deposits to the south-west of the waste dump

2.12 Social and environmental context of the background local area

Two areas were defined as local background areas, considered as not impacted by the
dump. The two areas are situated far from the dump but in the same geological context,
the first one is situated east of the dump and is characterized by being a protection area
and the local soil is used for subsistence agriculture. Geologically it is founded in the
Ponta Vermelha Formation (TPv) Fig. 2.25a. The second is located north of the
dumpsite, the soils are used for subsistence agriculture and is a protected area, which
was previously designed to be a cemetery for the Muslim community but due to evidence
of the existence of a water table very close to the surface the project was abandoned
and has since been a protected area. Geologically the area is situated in the Malhazine
Formation (QMa) Fig. 2.25b
37

Figure 2.25. Local background area (a) Soils of the TPv Formation, (b) Soils of the
QMa Formation.

3 Materials and Methods


In this research, to assess the influence of Hulene-B dump on the environment and
human health, several media samples were collected, and methods were combined:
geophysical methods (electrical resistivity), chemical and mineralogical study of soils,
rhizosphere soils, groundwater, stream water edible plants, road dust and human health
risks.

3.1 Geophysical studies

Over the last 50 years, geophysical survey methods have been used successfully in
environmental studies (Jayawickreme et al., 2014; Adamo et al., 2020), due to its non-
invasive nature in the acquisition and processing of data, as well as the various ways of
presenting the results that guarantee the full understanding of the final products,
especially by non-geophysical professionals (Reynolds, 1997; Koda et al., 2017;
Sentenac et al., 2018).

The most common environmental problems in soil and groundwater that geophysical
methods can be relevant to study are locating buried hazardous waste, identifying
38

different sources of contamination, leachate plumes from dumps and planning safe sites
for the disposal of industrial and domestic waste (Bichet et al., 2016; Hoai et al., 2021).

In the study of complex environmental problems, such as subsoil and groundwater


contamination, the three-dimensional nature of contamination plumes makes traditional
analytical methods, which include the collection and analysis of soils, sediments, waters
and observation wells, very limiting (Al Manmi et al., 2019). However, geophysical
methods have shown better results (Giang et al., 2018)

The weak point of geophysical methods is their uncertainty and therefore they have the
common designation of soft data. The results obtained should always be validated and
compared with data from direct exploration (hard data) such as chemical analyses
(Corradini et al., 2020).

Among the various geophysical methods that present good results in environmental
investigations are the electrical methods (Kaufman et al., 2010; Telford et al., 2012).

Electrical methods are based on the intrinsic electrical properties of geological materials
Conductivity (George, 2021), magnetic permeability and dielectric permittivity are
characteristics that can be measured using this methodology (Perrone et al.,2014).

The methods that make up this group are electrical resistivity, induced polarization,
spontaneous potential and electromagnetic (Perrone et al., 2014).

(a) In the electrical resistivity method, an electric current is introduced into the ground
by direct contact. The potential difference generated in the soil is measured using
electrodes (Telford et al., 2012). The relationship between the two magnitudes
provides an apparent electrical resistivity of the materials (Loke et al., 1996). This
method is widely used in the identification of heavy metal contamination areas,
contamination plumes (Abdulrahman et al., 2016; Liao et al., 2018; Udosen,
2021; Wu et al., 2021). However, its logistics consist of an arduous field
procedure (displacement of electrodes and cables) (Aizebeokhai, 2010).

(b) Induced polarization is an electrical phenomenon stimulated by electrical current


observed as a delayed response to voltage, in natural materials (Keller, 1978).
Currently, the induced polarization method is also used in the identification of
contamination plumes from inorganic waste (Liao et al., 2018; Kessouri et al.,
2019).

(c) The spontaneous potential method is based on certain natural or spontaneous


potentials produced underground due to electrochemical or mechanical activity,
39

where groundwater is the most important agent in the spontaneous potential


generation mechanism (Sato et al., 1960). Chen et al. (2018) the potentials may
be associated with the presence of metallic bodies, contacts between rocks of
different electrical properties (mainly conductivity), bioelectric activity of organic
materials, corrosion, thermal and pressure gradients in the subsurface fluids.

(d) Electromagnetic (EM) methods are based on the propagation of electromagnetic


fields (continuous or transient) underground (Baawain et al., 2018). The method
makes it possible to determine contamination plumes and the direction of their
flow, in former landfills or landfill sites (Tezkan, 1999).

Table 3.1 summarizes the applications of geophysical methods in environmental studies


is presented below, where electrical resistivity is one of the most suitable methods for
studying the flow and temporal variation of leachate plumes in and around the dumps

Table 3.1. Applicability of some geophysical techniques to different environmental


problems (adapted from Reynolds, 1997; Oliveira, 2009)

Prospecting Geological Hydraulically waste Contamination


technique structures active dumps plumes
and structures
barriers
Gravimetry + 0 - 0
Magnetic + - 0 -
Spontaneous 0 + 0 -
Potential
IP Resistivity + + + +
Electromagnetics - + + 0
Ground penetrating 0 + 0 -
radar
Radiometry 0 0 0 -
Seismic refraction + 0 - 0
+ applicable; 0 limited applicability; - not applicable.

In this research, given the objective of our investigation and the availability of the
equipment, the electrical resistivity method was employed.

The use of the electrical resistivity method originated in the 1920s with the works of the
Schlumberger brothers, being further developed in the second half of the 20th century
(Reynolds, 1997).

Electrical resistivity studies are based on electric current injected into the ground through
a pair of electrodes (A and B - current electrodes), and the resulting potential difference
between another pair of electrodes (M and N - potential electrodes) (Koda et al. 2017).
40

Ground resistivity is calculated by distances between electrodes, applied current and


measured potential difference, based on Law of Ohm (Lau et al., 2019).

Soils electrical resistivity is a characteristic closely linked to the type, nature, and state
of alteration of geological formations (Kayode et al., 2019). In areas with potential
groundwater contamination has been used for, determining the depth of the groundwater
table (Akhtar et al., 2021), determining the distribution of contamination areas and the
direction of migration of pollutants, assessing the thickness of wastes deposited in a
landfill, and identifying possible leachate plumes (Koda et al., 2017; Adamo et al., 2020;
Bernardo et al., 2022). Soil apparent resistivity (ρa) can be determined, based on the
known difference between electric field potential (∆V), the current (I), and the distance
between electrodes (Koda et al., 2017), given by the equation Eq.1:
∆V
ρa = 𝑘 𝐼
(1)

where ρa is apparent resistivity, I the intensity of current applied to the soil by electrodes
A and B (mA), ∆V the differential potential between electrodes M and N (mV), and k the
geometrical coefficient of electrode positioning (m). The geometrical factor k is
dependent on the distribution geometry of the electrodes, as follows Eq.2:


𝑘= 1 1 1 1
(2)
( ___ – __ – ___ + __ )
AM BM AN BN

where AM, BM, AN and BN represent geometrical distance between electrodes A and
M, B and M, A and N, and B and N, respectively.

Abdulrahman et al. (2016), Liao et al. (2018), Udosen (2021), Wu et al. (2021), Touzani
et al. ( 2021), Igboama et al. (2022), Bernardo et al. (2022b) point out that in the study
of hazardous or non-hazardous waste disposal sites, the method can be employed for
the following purposes:

i. Locate and delimit the contamination plumes;

ii. Establish the direction and flow of the spread of the contamination plumes if
successive campaigns are carried out;

iii. Locate buried waste sites;

iv. Investigate groundwater quality.


41

3.1.1 Methods adopted

Schematic model of the methodological phases adopted in this research in the study
area is presented below (Fig.3.1).

Figure 3.1. Methodological summary.

3.1.2 Equipment

The resistivimeter is the main equipment for carrying out a study of subsoil resistivity.
This device is responsible for injecting electric current into the subsoil through two current
electrodes (A and B) and measuring the potential difference between the potential
electrodes (M and N) (Fig. 3. 2a).

Multi-electrode resistivimeters are currently used, which make the automatic exchange
between the electrodes responsible for the measurements. It also allows the temporary
storage of the readings for later transfer to the processing software. There are
resistivimeters with several channels (multi-channel) that enable simultaneous
measurements of the potential difference for the same injection (Oliveira, 2009).
42

The ABEM SAS 4000 resistivity meter was used for this purpose (Fig. 3. 2b). This
equipment has four channels that enable the measurement and storage of four data
simultaneously, giving the equipment a faster measurement speed. The following
equipment is included in this resistivimeter:

- An electrode selector, which is the unit responsible for selecting the electrodes
to be used in a given measurement according to the protocol added to the
resistivimeter;

- Electrodes - these are conductors shaped like a cross and pointed at one end.
The shape allows its fixation by means of a human effort;

- Cable roll - corresponds to 100-meter-long cables with 21 outlets separated by 5


meters. The exits are points used for connecting the cable to the electrodes;

- Connectors - are white cylindrical shaped connectors which are used as a


connecting link between the cable rolls;

- Short connection cables (cable jumpers) - are cables that establish contact
between the electrodes and the cable roll;

- Battery - serves as the 12V voltage source required for the operation of the
equipment (ABEM Terrameter SAS 4000 / SAS 1000, 2011).

Figure 3.2. (a and b) ABEM SAS 4000 Resistivity meter used for data acquisition.

3.1.3 Planning and prospecting campaigns

For the study of the contamination of the dumps' surroundings by leachate plumes, geo-
environmental knowledge of the area is necessary(Olla et al., 2015; Netto et al., 2021).
43

Such as geological material, relief, groundwater flow direction, establishment of


reference areas, carrying out at least two campaigns (Aliewi et al., 2021)

The performance of at least two campaigns is vital to understand the pollutant plume
flows in the different seasons of the year (Ubechu et al., 2021; Touzani et al., 2021;
Helene et al., 2021)

In this way, 10 profiles of electrical resistivity prospection were carried out, between
2020 and 2021 (Fig. 3.3). Being 4 overlapping profiles between 2020 and 2021 and 2
more in 2021. The addition of two profiles in 2021 was relevant for the global
comprehension of the Hulene-B area, concerning the possible flow directions of the
leachate plumes.

Figure 3.3. Profiles executed in 2020 and 2021 (a) intradune depression, (b) Hulene -B
dump (c) Julius Nyerere Avenue (adapt. Google earth Pro)
44

3.1.4 Data acquisition

Once the electrical survey criteria had been defined, stainless steel electrodes were
placed and connected to the cables (Fig. 3.4). In the usual protocol in the ABEM SAS
4000 resistivimeter, the electrodes are 5 m apart from each other (ABEM, 2018)

Thus, a total of 21 electrodes were placed for each 100 m cable, both connected. In
some cases, given the particularities of the site (natural basin of leachate reception), the
cable passed through an aqueous medium, which allowed the submersion of the area of
current transmission between the cable and the other electrodes (Fig. 3.4).

Figure 3.4 (a) electrode placement process, (b) cable submerged in the surface
leachates of the reception basin and (c) cables between the pikes of the reception
basin.

3.1.5 Acquisition and recording of readings

The data acquisition process took place in the month of January 2020 and May 2021,
having successfully executed the planned lines (10 lines of 400 m), one per day.

An ABEM SAS 4000 resistivity meter and its accessories were used to acquire data on
all the lines, including four (4) rolls of 100m cable with 21 outlets each. The layout
produced by this sequence of cables (100m and 21 outlets) corresponds to the standard
of the reading program housed in the LUND Imaging System resistivimeter.
45

The readings in this program employ a current frequency of 50Hz, the data acquisition
protocol was GRAD4L8 and GRAD4S8 and readings were taken from 1081 to 946
points. The adopted arrangement was multigrid that accommodates the Wenner -
Schlumberger. In studies of contamination plumes in areas with heterogeneous
structures, the Wenner - Schlumberger arrangements are most recommended given
their complementarity in surface structures and horizontal as well as vertical variations
in resistivity (Loke, 2010; Wang et al., 2021).

The electrode spacing for data acquisition was 5 m. After the readings, the data was
transferred to the resistivimeter, which then takes the readings, 3 at minimum and 6 at
maximum, in order to obtain an average minimum error between readings. Given the
length of the lines adopted in data acquisition, we opted for a half spacing model for the
distance of the electrodes (2.5m). This process allows acquiring a very detailed final
model of the resistivity distribution (Loke, 2010).

The first and third measurements were made on the basis of three (3) cables. In this
scope, the first measuring station was located between the first two (2) designated cables
(2) and (3), where the first-cable (1) was ignored. For all cable roll joints, the overlap
between the last exit of one cable and the beginning of the subsequent cable was made
(Fig. 3.5 a and b).

Figure 3.5. (a) arrangement of three cables in the first measurement and (b) cable
connection (adapt. ABEM, 2018)

At the second station all four cables were connected. The resistivimeter remained at the
midpoint of the cable and readings were taken corresponding to all four cables (Fig. 3.6
a and b).
46

Figure 3.6. (a) and (b) Arrangement of the cable reels in the second reading station
(adapt. ABEM, 2018)

In the third measurement which corresponds to the last reading in the 400 m profile, the
last cable (4) was ignored. Three cables were used and designated cables 1, 2 and 3
(Fig. 3.7 a and b).

This procedure was employed in view of its versatility in acquiring good results in
Wenner and Schlumberger measurements where the last cable is excluded.

Figure 3.7. (a) and (b): Arrangement of the cable drums in the third reading station
(adapt. ABEM, 2018)

The use of these measurement sequences makes it possible to achieve a high resolution
of the surface near the ends of the measured section and this is important for good
resolution at depth. Before proceeding with the readings at the measurement stations,
electrode tests were performed, which allowed identifying electrodes poorly connected
to the cables, electrodes with conductive problems and proceed with their proper
47

correction. This process reduces the Root Mean Square (RMS) error that results from
noises in data acquisition that can be systematic or random (Loke, 2004).

3.1.6 Executed profiles

Profile 1 - was executed in a S-N direction (Fig.3.8), it corresponds to the western limit
of the dump and is characterized by low reliefs, marking a transition zone from the lowest
level of the dump to the natural basin for receiving surface leachates, where most of the
surface leachates are naturally directed and accumulated.

Figure 3.8. (a) Section south of profile-1 2020 (b) Section south of profile-1 2021.

Profile 2 - corresponds to the northern limit of the dump, it was executed in the E-W
direction (Fig.3.9), we consider it essential to understand the dynamics of the
contaminants from the source of contamination to the reception basin.
48

Figure 3.9. (a) Section West of profile-2 2021 (b) Section West of profile-2 2021.

Profile 3 - was performed in the S-N direction (Fig.3.10), from the edge of the catchment
basin to the area temporarily flooded by black waters, therefore, we consider it as
fundamental to understand the dynamics of the contaminants from the catchment basin
to the possible areas of influence of the catchment basin. This line, also served as a
reference given its distant location from the dump.

Figure 3.10. (a) South of profile 3 (b) North of profile.

Profile 4 - was run NE-SW in 2020 and SW-NE in 2021 (Figure.3.11) on the western
edge of the dump, we consider as fundamental to understand the different levels of
contamination of the dump towards the southern part of the catchment.
49

Figure 3.11. (a) South-West of profile-4 (b) North-West of profile-4.

Profile 5 - corresponds to the southern boundary of the dump, extending in E-W (Fig.
3.12) direction, it was carried out in order to understand the possible migration of
leachate towards the south of the dump.

Figure 3.12. (a) Profile-5 layout (b) Profile-5 environmental context.

Profile 6 - extending in a S-N direction on the eastern edge of the dump (Fig. 3.13),
corresponds to the highest elevations of the local relief and was built to understand the
possible flow of the contamination plumes in the eastern direction of the dump.
50

Figure 3.13. (a) Profile 6 layout (b) Profile 6 environmental context.

3.1.7 Data conversion

The data inversion was done based on the RES2DINV3.59.106 program, which is many
used in similar studies, such as Abdulrahman et al. (2016); Lau et al. (2019); Kayode et
al. (2019); Ugbor et al. (2021); Ashraf et al. (2022).

RES2DINV is a program that automatically determines a two-dimensional (2-D)


resistivity model for the subsurface from data obtained from electrical imaging surveys
(Loke, 2010).

This program can be used for searches using the Wenner, pole-pole, dipole-dipole, pole-
dipole, Wenner-Schlumberger, gradient and dipole-dipole arrays. In addition to these
common arrays, the program also supports unconventional arrays with an almost
unlimited number of possible electrode configurations (Loke et al.,1996). D model used
by the inversion program employs a large number of rectangular blocks (Fig.3.15 and
16). The distribution and size of the blocks are generated automatically by the program,
using the distribution of data points as an approximate guide (Loke et al.,1996)
51

Figure 3.14. (a) Arrangement of the blocks that loosely link to the distribution of data
points in the pseudo selection (adapt. Loke, 2010).

Figure 3.15. (b) Arrangement of the blocks that loosely link to the distribution of data
points in the pseudosecction (adapt. Loke, 2010).

3.1.8 Model noise and correction methods

Systematic noise can occur due to some fault during the data acquisition process,
including cable breaks, poor contact between the electrodes and ground so that current
is not injected into the ground, forgetting to connect an electrode to the cable, etc. This
type of noise is easy to detect in a large data set, since it is usually present in a small
number of readings, and the incorrect values stand out well from the rest (Fig.3.16).
Random noise includes effects such as telluric currents that affect all readings, giving
rise to data with values higher or lower than their noise-free counterparts (Loke, 2004).
52

Figure 3.16. Example of noise corrections: (a) profile with overlapping points (b)
overlapping point correction (adapt. ABEM, 2018).

3.1.9 Inversion of the final model

The inversion of the final model by the program uses matrix Gauss-Newton least squares
smoothing, the vertical and horizontal smoothing contrast is the same. A direct modelling
subroutine is used to calculate the apparent resistivity values (Olayinka et al.,2000).

The smoothing contrast increases by 10% per layer. As there is an increase in thickness
the resolution is reduced with depth (Figure.3.17). The inversion is based on the
analytical calculation of the sensitivity matrix (Jacobian matrix) for a homogeneous half-
space; and the matrix is recalculated, using the finite element (FE) method, at each
iteration step.

The optimization method basically tries to reduce the difference between the original
points, of the pseudosecction, and the new points, calculated by the transformation
53

process in the synthetic pseudosecction. The measure of this difference is given by the
root mean square error (RMS5).

Figure 3.17. Optimization of the profiles by the smoothing method: (a) Calculated
apparent resistivity (b) Calculated resistivity (c) Final smoothed model.

3.1.10 Groundwater Vulnerability Assessment (DRASTIC modified)

DRASTIC model has been commonly used in areas where geographical, and
hydrogeological information is available and has been successfully applied in different
regions (Asfaw et al., 2020; Anshumala et al., 2021; Shah et al., 2021). The model has
been applied to determine the groundwater vulnerability index of watersheds (Aller,
1987). The word DRASTIC is an abbreviation of initial letter of different parameters such
as ‘D’ to depth to water; ‘R’ to net recharge, ‘A’ to aquifer media, ‘S’ to soil media, ‘T’ to
topography, ‘I’ the impact of the vadose zone media, and ‘C’ to the hydraulic conductivity
of the aquifer intrinsic vulnerability of groundwater is evaluated by DRASTIC index
formula which is given below Eq. 3:

DRASTIC Index = DrDw + RrRw + ArAw + SrSw + TrTw +IrIw + CrCw (3)

where "r" is the rating value, and "w" is the weight assigned to each parameter. Each
factor is assigned a relative weight ranging 1 to 5 (Table 1). Each DRASTIC factor is
divided into ranges that affect the contaminant potential. The range for each factor lies
54

from 1 to 10. DRASTIC model depends on seven boundaries or layers, which are used
as input boundaries for modeling. Thus, the interpretation of the index follows three
categories, (i) indices < 135 denote low vulnerability; (ii) indices between 135 and 150
represent medium vulnerability; and (iii) indices > 150 suggest high vulnerability of
groundwater-related environmental impacts (Aller, 1987; Ghosh et al., 2021).

In this study was applied the modified DRASTIC model combined with electrical
resistivity data. This allows a specific assessment of the vulnerability on the dump
surrounding area, based on the distance between anomalous surface layer (low
resistivity; leachate influenced) and groundwater. Soil variable was replaced, since soils
around the landfill were classified as sandy, which is characteristic of the whole
surroundings (Momade et al., 1996). Thus, ‘S’ corresponds to the distance between
superficial anomalous layer and groundwater. Values and weight of the variable were
kept the same as for soil, given processes similarity, controlled by these factors
(migration and attenuation of leachate) (Table 3.2). Layers spatial distribution was not
made, given detailed description for each factor in depth and superficial slight change,
as well as study area size (Hosseini et al., 2018; Kozłowski et al., 2019).
55

Table 3.2 DRASTIC parameters.

Factor Interval/Characteristics Value (r) Weight (w)


(D) 0 – 1.5 10
Groundwater depth (m) 1.5 – 4.6 9
4.6 – 9.3 7 5
9.3 – 15 5
15 – 23 3
23 – 30 2
> 30 1
(R) 0 – 50 1
Net recharge rate 50 - 100 3
(mm/year) 100 - 175 6 4
175 – 250 8
> 250 9
(A) Sand 7 3
Aquifer media
0 – 1.5 10
(S) 1.5 – 4.6 9
Distance between the 4.6 – 9.3 7 2
anomalous surface layer and 9.3 – 15 5
groundwater 15 – 23 3
23 – 30 2
> 30 1
(T) 0–2 10
Terrain slope (%) 2–6 9
6 – 12 5 1
12 – 18 3
> 18 1
(I) Sandstones 4–8
Vadose Zone Limestones, sandstones, and 4–8
shales 5
Sands and gravels with 4–8
significant silt and clay content
Sands 8
(C) 1 – 4.1 1 3
Hydraulic conductivity 4.1 – 12.2 2
(m/day) 12.2 – 28.5 4
28.5 – 40.7 6
40.7 – 81.5 8
> 81.5 10

3.1.11 Validation

For resistivity models and groundwater vulnerability validation, groundwater depth, pH


and sulphates (PO43–) were measured in two wells in the surrounding of the dump
(Fig.3.17). chemical analysis of total phosphate was performed with a HANNA - HI96713
Multisonde instrument, with a resolution level of 0.01 mg/l.
56

Figure 3.18. (a) Water collection - South well (b) Water collection - North well; (c)
Chemical analysis process (d) phosphate analysis

3.2 Geochemical studies (soil, rhizosphere, stream water, groundwater,


edible plants, road dust)

3.2.1 Soil sampling

Contamination of soils by PTEs is influenced by several factors, such as soil texture, pH,
organic matter, climatic conditions, topography, direction of leachate surface flows, wind
direction and others (Campos, 2010; Chaudhary et al., 2021). The waste characteristics
(type of waste, age, coverage of waste) are also relevant as they define the capacity for
leachate generation and dispersion of incineration ashes which are the main soil surface
contamination mechanisms (Zamorano et al., 2005; Calvo et al., 2005; Rapti-Caputo et
al., 2006).

A total of 71 superficial soil samples were collected in the surrounding area of Hulene-B
waste dump, in January 2020 and May 2021, in an area of about 40 hectares (Fig. 3.19).
Given the difficulties of access to the sampling points, the sampling grid was changed
and the distance between samples varied from 20 to 45 m (Fig. 3.19). The 2020 samples
were collected in January, a period characterized by heavy rainfall in Maputo, with a
57

monthly average of up to 123 mm, and the 2021 samples were collected in May, a period
characterized by low rainfall, with a monthly average of 25 mm(CIAT, 2017). To
determine the background, 10 samples were collected in areas considered not impacted
by the dump (Fig.3. 19).

Figure 3.19. Soil sampling area in the surrounding of the dump (yellow) , and
background samples (green).

Soil samples were collected at 0-20 cm depth (Fig. 3. 20a). Samples were georeferenced
at the collection sites and preserved in plastic bags until laboratorial treatment (Fig. 3.
20b). On the laboratory, samples were oven dried ≤ 40 ºC (Pedagogical University of
Maputo, Mozambique) (Fig. 3. 20c). Afterwards, samples were transported to the
laboratory of GeoBioTec Research Center (University of Aveiro, Portugal-UAVR), for
analyses.
58

Figure 3.20. soils sampling and drying process: (a) Soil collection 2020, (b) Soil
collection 2021 (c) Soil drying in UPM laboratories.

3.2.2 Rhizosphere soils and edible plants

Rhizosphere soils were collected from horticultural fields west of the dump (Figure 3. 21).
These fields spread along the depression of the intra-dune rapidity and are the source
of livelihood of some families.

Figure 3.21. Sampling points for soil, rhizosphere, plants, and water (R)-Rhizosphere
and (P) Plants (edible plants)
59

Four cultivated edible crops were collected from four fields at the western end of Hulene-
B dump (Fig. 3. 22). The species collected were Amaranthus leaves (Amaranthus
spinosus) at the western end of the riparian zone, sweet potato leaves (Ipomoea batatas)
at the eastern end of the riparian zone, pumpkin leaves (Cucurbita pepo) at the south-
eastern end and south-western cabbage leaves (Brassica oleracea L). The leaves of the
plants were collected, as these are the parts that are consumed locally. The plants are
grown on the periphery of the riparian zone, due to the humidity of the soil and the
proximity of the irrigation water (stream water) (Fig.3.22). After collection the samples
were transported to the laboratory and dried in an oven ≤ 40ºc.

Figure 3.22. (a) vegetable variety (b) pumpkin leaves (c) cabbage.

3.2.3 Water sampling (wells and stream water)

Four water samples were collected on the western limit of Hulene-B waste dump
(Fig.3.23). All samples were collected in May 2021. Two samples were collected in
stream water used for irrigation, one in north-western of stream water (IWN) and south
(IWS), and two samples collected in wells used for human consumption, one in south-
west (WS) of dump and north-west of the dump (WN). Sample WS, with ~ 5 m depth,
had no cover allowing materials deposition and sample WN, with ~ 6 m depth, was
covered, reducing materials deposition. Wells water was collected using the local used
containers. All samples were collected in clean polyethylene bottles during the morning
and transported to the laboratory for analysis. Waters pH was measured during
sampling. pH classification adopted for this study was the one proposed by (WHO, 2017):
neutral if = 7, acidic if < 7, and basic if > 7.
60

Figure 3.23. (a) south well, (b) north well, (c) stream water

3.2.4 Road dust Sampling

Road dust samples were collected in the main roads of Maputo city, characterized by
heavy traffic, informal roadside commercial activities, with a high density daily (Fig. 3.24).
For each sample a plastic shovel and a broom were used. These streets are enclosed
by numerous dwellings (Fig.3.24). Other existing features in the surroundings were
Maputo International Airport, and the largest open-air dumpsite of the city (Hulene-B
waste dump; Bernardino et al., 2022)

Figure 3.24. Sampling locations (yellow dots) and wind directions. Red dashed -
Hulene-B dump (adapt. Google Earth, 2022).
61

Sample 1 was collected southeast of Maputo International Airport, characterized by


intense traffic and street food commercial activities. Sample 2 was collected in Beira
Street, in the bus stop, with diverse commercial small shops and dwellings (Fig.3.25a).
Sample 3 was collected between the eastern boundary of Hulene-B waste dump and
Julius Nyerere Avenue, with heterogeneous features and densely inhabited (Fig.3.25b).
Sample 4 was collected in the southwest of Maputo airport, characterized by heavy traffic
and densely populated. Samples 5, 6 and 7 were collected in bus stops along the
National Road n. º 1, characterized by intense traffic and several commercial activities
(Fig. 25 b and c). Samples 8 and 9 were collected in bus stops, with intense commerce
on the sidewalks (Fig. 25 d). All samples were collected in June 2021. This period was
characterized by scarce precipitation.

Figure 3.25. Road dust collection (a) sample 3, (b) sample 6, (c) sample 5, (d) sample
9

3.3 Sample preparation and analysis


3.3.1 Soils, rhizosphere soils and road dust

Samples were georeferenced and preserved in clean plastic bags until laboratory
treatment. In the laboratory, samples were dried in an oven ≤ 40ºc and sieved to for the
< 2000 and < 63 µm (Pedagogical University of Maputo, Mozambique). Afterwards,
samples were transported to the laboratory of GeoBioTec research center (University of
Aveiro, Portugal), for physical, chemical, and mineralogical analyses (Fig. 3.26).
62

Figure 3.26. Methodological summary of soil, rhizosphere, and dust analysis. * XRD for
soils and rhizosphere was analyzed only at < 2000 µm and for dust < 63 µm.

3.3.2 pH

Soil pH has a major influence on heavy metal solubility (Sparling, 2020). pH refers to the
concentration of the H+ proton in the soil solution (Campos, 2010). pH conditions > 6
favour the dissociation of H+ from OH groups in organic matter and Fe and Al oxides,
increasing the adsorption of heavy metals and subsequent precipitation and reducing
their bioavailability (Pagnanelli et al., 2003). However, in soils of tropical regions, heavy
metal retention can occur under high pH conditions (Campos, 2010), due to the
predominance of oxidic (Al, Fe and Mn) and kaolinitic mineralogy in the clay fraction,
which increases the metal adsorption capacity (Alleoni et al., 2005).

Soil samples were sieved to achieve the < 2000 (sand) and < 63 (silt) µm fractions. pH
was determined in the two fractions with a 1:2.5 soil/water solution using a pH meter. pH,
3.5 - 4.4 extremely acid, 4.5 - 5.0 very strongly acid, 5.1 - 5.5 strongly acid, 5.6 - 6.0
moderately acid, 6.1 - 6.5 slightly acid, 6.6 - 7.3 neutral, 7.4 - 7.8 slightly alkaline, 7.9 -
8.4 moderately alkaline, and 8.5 - 9.0 strongly alkaline. The pH was determined in the
two fractions with a 1:2.5 soil/water solution using a HI9126 high resolution pH meter.

3.3.3 Electrical conductivity (EC)

Soil electrical conductivity (EC) is a measure of soil ability to conduct an electric current,
being influenced by, e.g., moisture, OM, texture, and others (Grisso et al., 2009). In
general, sandy soils present EC ≤ 100 μS/cm (Lund, 2008; USDA, 2014).
63

Hussein et al. (2021) found that soils with high EC values were associated with high soil
contamination, by As, Cd, Pb and Cr in a study of 6 dumpsites in Malaysia. Similar
findings by Wu et al. (2021) on a study of waste dumps in China, Fatoba et al. (2021)
on a waste dump in Nigeria, with contamination by metal-ion enriched leachates.

Electrical conductivity (EC) was measured under the same conditions as pH in the two
fractions, using a high-resolution conductivity meter. (Fig. 3.27).

Figure 3.27. pH and EC measurement

Electrical conductivity classes were defined based on internationally adopted reference


values for sandy soils (Lund, 2008; USDA, 2011): ≤ 100 μS/cm natural medium
conductivity, 101 - 200 μS/cm high conductivity, and > 201 μS/cm very high conductivity.

Rhizosphere EC was analyzed taking into account the reference values for sandy soils
in the contamination study ≤ 100 μS/cm (USDA, 2011), and soil salinity conditions for
horticultural crops ≤ 1100 μS/cm (USDA, 2014).

3.3.4 Organic matter (OM)

OM shows great affinity for heavy metals, thus increasing its absorption capacity by the
soil (Ghobadi et al., 2021). Soil organic matter is described as the carbon-rich material
that includes plant, animal and microbial residues at various stages of decomposition
(USDA, 2001a). Organic matter (OM) content was determined with method proposed by
USDA (2001). Thus, it was weighed ~ 5 g of sample in a crucible, in the analytical balance
with resolution of (0.001g), placed about 24h in a muffle furnace at a temperature of
105ºC degrees for the determination of moisture. Sequentially, the weight was measured
64

after 105 ºC degrees and placed in a muffle furnace at a temperature of 430 ºC degrees
for 20h for the determination of the organic matter.

3.3.5 Moisture

Soil moisture is the water content of the soil (USDA, 1998). Soil moisture is an important
factor in the dissolution of metal ions contained in leachate (Kumar et al., 2013).
Moisture was measured based on the procedures defined by (Reeuwijk, 2002).

The moisture classes were defined according to the average of the local background
(2020), value of the samples considered standard, 0.41 - 0.82 % high and 0.82 - 1.9 %
very high. In 2021 the aim of the study was to make the comparison, the moisture classes
were kept with those of 2020 as class I - 0.18 to 2< 0.41 % low; class ii - 2.10.41 to 4
0.82 % high; and class iii 4.1 to > 0.82 8.7% - very high. Color

One of the parameters affecting soil color is OM content (USDA, 2001). In solid waste
contaminated areas, soil color is influenced by leachates, which are a potential for metal-
organic interactions through organic ligands (Lee et al., 2022). In these areas soil color
has been used to indicate the possible interference of leachate in changing soils physical
parameters (Gonçalves et al., 2019). Darker colors are associated with soils influenced
by leachate (Naveen et al., 2017), and whitish colors are often indicative of leaching
processes (Lee et al., 2022).

Soils color was determined using Munsell Color, (2009) soil chart. Soil color classes were
defined taking in consideration color intensity: blackish, grayish, and brownish. This
classification was used for the soils of the two campaigns (Fig. 3.28)

Figure 3.28. Identifying colors using the Munsell Soil Chart


65

3.3.6 Granulometry

soil granulometry plays a key role in the fixation and migration of heavy metals by
leaching processes (Dakheel et al., 2022). In general, sandy soils are propitious to
leachates migration in depth, promoting subsoil and groundwater contamination (Victor
et al., 2019).

Silt fraction granulometric distribution was determined with a Micromeritics Sedigraph III
Plus grain size analyzer (UAVR). This technique determines the relative mass
distribution of a sample by particle size and is based on two physical principles:
sedimentation theory (Stokes' law) and the absorption of X-radiation (Beer-Lambert law).
The analytical accuracy and precision of the methods were determined using analyses
of reference materials and duplicate samples in each analytical set. Results were within
95% confidence limits of the recommended values for the certified material. The relative
standard deviation was between 5 and 10%.

3.4 Chemistry (FRX) and Mineralogy analyses (XRD)


3.4.1 Soils, rhizosphere soils, road dust and edible plants

Samples chemical composition was accessed by X-ray fluorescence (XRF)


spectrometry, using a PANalytical PW 4400/40 45 Axios with Cr Kα radiation.
Mineralogical phases were determined by powder X-ray diffraction (XRD) using a
Phillips/Panalytical powder diffractometer, model X’Pert-Pro MPD, equipped with an
automatic slit. A Cu-X-ray tube was operated at 50 Kv and 30 mA. Data were collected
from 2 to 70º 2θ with a step size of 1º and a counting interval of 0.02 s. Precision and
accuracy of analyses and procedures were monitored with UAVR internal standards,
certified reference material and quality control blanks. Results were within the 95%
confidence limits. The relative standard deviation was between 5 and 10 %.

3.4.2 Water (wells water and stream water) analyses

Water samples chemical analysis was performed using a Metalyser H1000 Trace-2oto
analyze Pb, Hg, Zn concentrations (Fig. 3. 29). This equipment is recommended in the
analysis of heavy metals (AsIII, As Total, Cd, CrVI, Cu, Pb, Mn, Hg, Ni and Zn) given its
66

robustness, high level of accuracy and has 97% correlation of Atomic Absorption
Spectrometry (GF-AAS) estimation (Bhardwaj et al., 2020).

Figure 3.29. Equipment used in the chemical analysis of water (a) Reagent kit (b)
Chemical analysis process.

3.4.3 Data Analysis

The data obtained by the various techniques applied in this research were analyzed
individually and then combined. In this way, the geophysical data were combined with
internationally accepted models to study the risk of contamination by plumes from
dumps. The data of soil physical, mineralogical, and geochemical parameters were
compared with internationally accepted references as well as with local background
values to assess the contamination levels and the respective most impacted areas. The
spatial differentiation of contamination levels allowed the identification of areas that
represent a potential risk for human health.
67

3.4.4 Descriptive analysis

To understand the distribution of contamination processes in the study area, a univariate


analysis of the data was conducted. The univariate analysis allowed us to detect,
compare and regionalize different levels of contamination and risk indicators. Generally,
in environmental contamination studies, the systematization of data in graphs,
histograms, tables with mean values, medians and standard deviation is relevant to
understand the different levels of contribution of each element in environmental
contamination (Ott, 1995).

3.4.5 Correlations analysis (r)

Spearman and Pearson methods are commonly applied to assess the correlation
between variables (Schober et al., 2018). Spearman and Pearson methods are usually
applied to assess a correlation between variables. Spearman is classified as a non-
parametric method, it is used to measure the degree of association between two
variables, while Pearson is a parametric method that should only be used with a normal
distribution, and is used between variables that are linear (Yadav, 2018). In both
correlation techniques, correlation coefficients between variables can range from [-1.00,
+1.00], with -1.00 representing a perfect negative correlation, +1.00 a perfect positive
correlation and 0.00 represents no correlation (Yadav, 2018).

Thus, soil physical parameters such as EC, OM, pH, Moisture, Color, Granulometric and
the PTEs (Cr, Cu, Mn, Ni, Pb and Zn) were subjected to extensive relationship analysis
through these correlations. The correlation data were calculated using IBM SPSS
Statistics 20® software.

3.4.6 Principal Component Analyses (PCA)

PCA is a data dimensionality reduction technique that aims to explain most of the
variations in the data with a small number of independent variable data (called "principal
components") (Hou et al., 2017). Candeias et al. (2014) and Perona et al. (1999) referred
that currently the technique is widely used in environmental impact studies, by
elucidating relationships between variables and identifying common underlying
processes.

The main objective of PCA is to provide a small number of independent factors (principal
components) that synthesize the associations between variables, the referred factors (or
PCs) being orthogonal linear combinations of the variables. The first PC explains most
68

of the total variance of the data set, and each successive PC explains a smaller part of
the remaining variance. The different PCs are then related to common processes
affecting the variables through expert knowledge of the situation under analysis.

The number of significant principal components for interpretation is selected based on


the Kaiser criterion with an eigenvalue greater than 1 (Kaiser, 1960) and a total explained
variance of 70% or more.

The PCA results will be discussed in the characterization and analysis of soil physical
parameters and geochemical evaluation section. However, the results showed a strong
association between OM, EC, and moisture associated with a contamination by leachate
originated from the dumpsite. The same was notable for some potentially toxic elements
Cr, Cu, Mn, Ni, Pb, Zn, Zr), which were associated to the Hulene – B waste dump source.

PCA analysis was performed using SPSS® v.25 software (IBM, USA).

3.4.7 Analysis of Variance (1-way ANOVA)

One-way analysis of variance (ANOVA) is used to determine whether statistically


significant differences exist between the means of three or more independent (unrelated)
groups Eq. 3 (Olilima et al., 2021).

𝐻0 : 𝜇1 = 𝜇2 = ⋯ = 𝜇𝑘 =𝜇 (3)

Where µ = group mean and k = number of groups. If, however, the one-way ANOVA
returns a statistically significant result, we accept the alternative hypothesis (HA), which
is that there are at least two group means that are statistically significantly different from
each other.

One-way analysis was relevant for the determination of the statistically significant
elements, from which the PTEs showing the highest significance in the 2020 samples
(Cr, Cu, Mn, Ni, Pb, Zn, Zr in soils).

3.4.8 Spatial distribution

In the study of soils, one of the most applied methods for understanding the spatial
differences of their characteristics is mapping. In this study, soil physical characteristics
(EC, OM, pH, Moisture, color) and contamination risk indices (Potential ecological risk
69

index (PERI), pollution Load index (PLI) and Soil Nemerow index calculated) were
mapped. In this way, spatial distribution classes were defined.

pH classes were defined de accordance with USDA (1998): 6.6 - 7.3 neutral,7.4 - 7.8
slightly alkaline, and 7.9 - 8.4 moderately alkaline. Organic matter classes were defined
in accordance with USDA (2001):>4% high content, 2 – 4% medium content, and <2%
low content. Electrical conductivity classes were defined based on internationally
adopted reference values for sandy soils (Lund, 2008; USDA, 2011): ≤ 100 μS/cm natural
medium conductivity, 101 - 200 μS/cm high conductivity, and > 201 μS/cm very high
conductivity. Soil color classes were defined taking in consideration color intensity:
blackish, grayish, and brownish. Moisture classes were defined according to the average
of the local background value considered standard, 0.41-0.82 % high and 0.82- 1.9%
very high.

3.5 Ecological risk assessment (soils, road dust)


3.5.1 Potential ecological risk index (PERI)

PERI has been widely used to assess the degree of potential ecological risk of heavy
metals in soils, which was originally used in sediment samples (Candeias et al., 2022).
Potential ecological risk factor is defined as the sum of the monomial potential ecological
risk factors (EF), which quantitatively defines the potential ecological risk of a
contaminant in a sample (Hakanson, 1980) and is calculated by Eq.4:

PERI = ∑7i=1 EFi = ∑7i=1 CFi × TF (4)


where PERI is the Potential Ecological Risk Index, EFi is the monomial potential
ecological risk factor of each variable; CFi is the single contamination factor, and TF is
the heavy metal toxic-response factor for each element. Soil toxic-response factors were
computed for the seven selected elements (Hakanson, 1980). The TF values obtained
were Cr = 25, Cu and Zr = 10, Ni and Zn = 5 and Mn and Pb = 2. Five EF classes and
four PERI degrees are presented in Table 3.3.
70

Table 3.3 Monomial potential ecological risk factor (EF) and Potential ecological risk
index (PERI) classification levels (Hakanson, 1980)

EF Classification PERI Classification

0 ≤ EF < 40 Low PERI < 150 Low


40 ≤ EF < 80 Moderate 150 ≤ PERI < 300 Moderate
80 ≤ EF < 160 Considerable 300 ≤ PERI < 60 Considerable
160 ≤ EF < 32 Very high 600 ≤ PERI Very high
320 ≤ EF Very high

3.5.2 Soil Nemerow index (Pn)

PN is widely used to assess pollutants contribution to soil contamination (Xiao et al.,


2021). Nemerow pollution index synthesizes the single pollution index values of all
elements and can comprehensively reflect the degree of soil pollution and highlight the
impact of elements concentration (Ciarkowska et al., 2022).

The index is calculated as Eq. 5:

𝑃𝑛 = √[(1/𝑛 ∑𝑛𝑖=1 𝑃𝐼)2 + (PI𝑚𝑎𝑥 )2 ]/2 (5)

where PI is the single element pollution index, PImax the PI maximum value in all
elements, and n the number of elements studied.

Soil quality is classified as, non-polluted PN ≤ 0.7, warning line of pollution 0.7 < PN ≤
1.0, low level of pollution 1.0 < PN ≤ 2.0, moderate level of pollution 2.0 < PN ≤ 3.0, and
high level of pollution PN > 3.0. Soil Nemerow index was calculated using both
background (PNbkg) and world soils (PNRC)(Reimann et al., 2005)

3.5.3 Pollution Load index (PLI)

PLI provides a simple comparative means to assess level of enrichment (Kumar et al.,
2022). It was determined as the nth root of the product of the nPI, being n the number of
variables considered Eq. 6:

PLI = (PI1 × PI2 × PI3 × ⋅ ⋅ ⋅ × PIn)1/n. (6)

PLI > 1 implies environmental deterioration by elemental pollution.


71

For PLI, a class 1 < 1 uncontaminated, class 2 > 1 onwards implies environmental
deterioration by pollution (Luo et al., 2022).

3.5.4 Pollution Index water

The Single Factor Pollution Index was calculated as Eq. 7:

Pi = Ci/Si (7)

(Yan et al., 2015) where Pi is the pollution index of pollution indicator i, Ci is the
concentration of pollution indicator in water (mg/l), and Si is the permissible limit for the
pollution indicator in water. The Single Factor Pollution Index (Pi) is classified into five
grades, according to (Ajibare et al., 2022). < 0.4 Non-pollution; 0.4–1.0 Slight pollution;
1.0–2.0 Medium polluted; 2.0–5.0 Heavy polluted; > 5.0 Serious polluted. The water
pollution index was determined based on the measured concentrations of Hg, Pb and
Zn. Reference values were defined based on World Health Organization (WHO) Hg = 6,
Pb 10 and Zn =3,0 µg/l for drinking water (wells) Hg = 6, Pb 10 and Zn =3,000 and stream
water Pb=20, Hg = 5, Zn = ≤ 5.000.

3.6 Landfill pollution Index


3.6.1 Soils and groundwater

The landfill pollution risk index (Ip) was proposed by Mancini et al. (1999), is based on
the determination of vulnerability to aquifer pollution and hazard induced by the
quantified landfill as a specific parameter that allows (i) to identify suitable sites to host
new waste digestion sites, and (ii) to define the priorities in the control and remediation
operations to be fulfilled in case of potential hazard landfills or sites that have already
been compromised (Rapti-Caputo et al., 2006).

The integration of Ip assessment has been used in many studies to evaluate the risk of
groundwater contamination and understanding the properties of the soils where a landfill
is located and its surroundings has been highlighted as relevant once soil is considered
a superficial defense of the hydrogeological system, where several important processes
take place within the soil that make up the attenuation capacity (Civita et al., 2004).
72

The method applies eighth variables to assess risk index: (a) volume of deposited waste;
(b) leachate drainage; (b) type of waste; (d) waste physical condition; (e) waste
biodegradability; (f) monitoring system; (g) waste compaction; and (h) final coating
(Chaudhary et al., 2021; Liu et al., 2019; Nadiri et al., 2017). Risk factor (Ri) and risk
weight parameter (Wi) for each of the eight variables are described in Table 3.4. Landfill
risk index was determined using Eq.8:

𝑛
𝐼𝑝 = ∑ Wi𝑥𝑅𝑖 (8) (Rapti-Caputo et
𝑖=1

al., 2006; Liu et al., 2019)

where Ip corresponds pollution risk index; Wi is weight of the risk parameter (1 to 5), and
Ri the risk factor. Thus, Ip index < 3 represents low pollution potential; I p 3 - 7 suggest
risk and medium vulnerability; Ip 7 - 9 indicate low risk and high vulnerability, and Ip > 10,
high contamination risk and immediate intervention measures should be taken to mitigate
the impact on groundwater (Rapti-Caputo et al., 2006; Liu et al., 2019).
73

Table 3.4 Landfill risk index assessment (Rapti-Caputo et al., 2006).


Reduction factor
variables Weight (W) Single risk elements
(R)
< 10 t/day 1
Volume of waste 10 a 50 t/day 0.2
5
deposited 50 a 500 t/day 0.4
> 500 t/day 1
External and internal drainage 0.1
Leachate drainage Internal drainage 0.3
5
system Reuse of leachate in the system 0.5
Absent drainage 1
Inert 0.1
Urban 0.5
Type of waste 3
Industrial - non-hazardous 0.8
Dangerous 1
Solidified with inert matrix 0.1
Physical state of waste 3 Solid 0.2
Mud with humidity < 70 % 0.5
Mud with humidity > 70 % 1
Non-biodegradable 0.1
Stabilization typology 2 Aerobic 0.3
Aerobic and anaerobic 0.5
Anaerobic 1
Well and geomembrane monitoring 0.1
Geomembrane 0.3
Monitoring system 2
Monitoring well 0.5
Absent 1
Compacted with pneumatic
0.1
equipment
Compacting the waste 1 Compacted with bulldozer 0.2
Manually compacted 0.5
No compaction 1
Compacted soil with clay 0.1
Compacted clay 0.2
Final coating 1
Non-compacted soil 0.5
Absent 1

3.6.2 Surface water risk

The probability of contamination surface water (Pbci) by landfills are parameters used in
the environmental assessment of landfill contamination proposed by (Calvo et al., 2005).
In order to assess the probability of contamination, landfill variables are selected that
show high sensitivity to biochemical and physical processes that directly or indirectly
influence the surrounding environment (Calvo et al., 2005; Aryampa et al., 2021). In this
study, 8 factors have been selected for the environmental risk assessment with close
relation to soil physical parameters (degree of compaction, type and amount of cover
74

material, surface permeability of surrounding strata, landfill liner system, final cover,
leachate control, leachate location and operating system) and surface water (degree of
compaction, type of waste, final cover, precipitation, surface drainage systems,
permeability of surrounding soil, distance to surface water and leachate control). The
evaluation of these variables makes it possible to assess contamination risks in landfills
through the Pbcj, or Probability of Contamination for each contamination variable. This
probability has the following expression Eq.9:

𝐶𝑖𝑥𝑊𝑖
Pbci = (9) (Calvo et al., 2005)
8

where Wj is a variable for parameter. Cj is the classification of variable j. depends on the


state of the variable and provides information on the status of the interaction between
the disposal processes and environmental characteristics related to the variable. Values
obtained range from 0 to 4 (Table 3. 5).

Table 3.5 criteria for assigning weights and ranking

Weight (Wi) Ranking Rating value (Ri)


Very high 4
High 3
Medium 2
low 1
1 Very low 0
Very high 4
High 3
2 Medium 2
low 1
Very low 0

The values of the weights are assigned according to the characteristics of the factor
evaluated in relation to the risk of contamination. Wj is the weighting of variable. When
the variable is related to the structural elements it will provide a greater weight (2)
Otherwise, it acquires the value 1 (Calvo et al., 2005).

The evaluation of the variables facilitates the assessment of contamination risks in


landfills (Aryampa et al., 2021). This evaluation is carried out through the concept of
Probability of Contamination Parameter (Pbci), which has the following mathematical
expression eq. 10. (Calvo et al., 2005)
75

∑ Pbc𝑖
Pbci = , (10)
𝑁

where N is the number of variables of each parameter. Pbci obtain a final value between
0 and 1 with classifications of maximum (Calvo et al., 2005). Thus, Pbci = 0 Nonexistent;
0 < Pbci < 0.3 Low; 0.3 Pbci <0.6 Mean; 0.6 < Pbci ≤ 1 High (Calvo et al., 2005).

3.7 Human health risk assessment

Long-term exposure to soil and high doses to water, vegetation and dust contaminated
by PTE poses a potential risk to public health (Ávila et al., 2017; Khan et al., 2018; Khan
et al., 2018; Candeias et al., 2022)

3.8 Soils (surroundings area of Hulene -B Dump)

Human health risk assessment for residents living around the Hulene - B landfill was
based on the assumption that residents, both children and adults, are directly exposed
to soil through three main routes (a) ingestion; (b) dermal absorption and (c) inhalation
of soil particles in air (Thongyuan et al., 2021; Candeias et al., 2014).

Ingestion of soil (i) occurs by eating soil particles and/or licking contact surfaces (e.g.
hands). Children are assumed to have a higher rate of ingestion due to habit of ingesting
soils/ or taking hand-to-mouth with soils (Ihedioha et al., 2017). Dermal absorption (ii)
occurs through exposed skin, while soil is inhaled (iii) through both mouth and nose
during respiration (Candeias et al., 2020). Particulates <10 µm (PM10) are the most
relevant in this process, although larger fractions of inhaled soil are likely to be broken
down in the gastrointestinal route (Candeias et al., 2014). It is assumed that all
contaminants are absorbed through both the gastrointestinal tract and the lung (Van den
Berg, 1995; Candeias et al., 2014; Gujre et al., 2021)

Equations were used to estimate the chronic daily intake for each exposure route
considered (Eq. 11) (Candeias et al., 2014) supplemented by specific quantitative
information (Table 3. 6).

𝐼𝑛𝑔𝑅 𝑥 𝐸𝐹 𝑥 𝐸𝐷
CDIing= Csoil x x 10-6
𝐵𝑊 𝑥 𝐴𝑇

SA 𝑥 SAF𝑥 DA 𝑥 EF X 𝑥 ED
CDIderm = Csoil x 10-6
𝐵𝑊 𝑥 𝐴𝑇
76

InR 𝑥 𝐸F 𝑥 ED
CDIinh = Csoil (11)
𝑃𝐸𝐹𝑋𝐵𝑊 𝑥 𝐴𝑇

Table 3.6 Variables used to assess human health risks

Parameters Meaning Child Values Reference


Adult
ABSgi fraction of contaminant Cr, Cu =1.00;
absorbed in gastrointestinal Cr Zr; Pb, Ni,
tract Mn,
Zn =1.00
ABSdrm fraction of contaminant 0.001
absorbed dermally from soil
ATc (d) averaging time for LT × 365
carcinogenic effects
ATnc (d) averaging time for non- ED × 365
carcinogenic effects
BW (kg) average body weight 15 60 (Nhantumbo, et
al., 2012)
Csoil (mg·kg−1) concentration of the element in
soil
DA dermal absorption factor 0.001 for all
CDIing (mg·kg−1·d−1) chronic daily intake dose -
through ingestion
CDIdrm chronic daily intake through
(mg·kg−1·d−1) dermal contact
CDIinh (mg·m−3) (nc), chronic daily intake through
(µg·m−3) (c) inhalation
CSFing chronic oral slope factor
((mg·kg−1·d−1)−1)

CSFdrm chronic dermal slope factor CSFing/ ABSgi


ED (yr) exposure duration 6 24
EF (d·yr−1) exposure frequency
ET (h·d−1) exposure time
IngR (mg·d−1) soil ingestion rate 100 200
InhR (m3·d −1) inhalation rate 7.6 20
IUR ((µg·m−3)−1) chronic inhalation slope factor
LT (yr) lifetime expected at birth 60 (INE, 2020)
PEF (m3·kg−1) particle emission factor 1.36 × 109 (USEPA, 1989;
RAIS, 2021)
SA (cm2) exposed skin area 2800 5700 (Candeias et al.,
2022)
SAF (mg·cm−2) skin adherence factor 0.2 0.07 (Candeias et al.,
2021a)
RfD ing (mg·kg−1·d−1) chronic oral reference dose Cu 4 × (USEPA,2013)
10−2; Zn 0.3
RfDdrm chronic dermal reference dose RfDing × ABSgi
RfDinh (mg·m−3) chronic inhalation reference
dose

The carcinogenic and non-carcinogenic side effects for each PTE were computed
individually, as toxicity calculation uses different computational methods (Zhang et
al., 2021). For each element and pathway, the non-cancer toxic risk was estimated by
77

computing the Hazard Quotient (HQ, also known as non-cancer risk-Equation (12)) for
systemic toxicity (Candeias et al., 2020). If HQ exceeds unity, it indicates that non-
carcinogenic effects might occur. (Candeias et al., 2020).

To estimate the overall developing hazard of non-carcinogenic effects, it is assumed that


toxic risks have additive effects. Therefore, it is possible to calculate the cumulative non-
carcinogenic hazard index (HI), which corresponds to the sum of HQ for each
pathway (Equation (13) (Candeias et al., 2020). Values of HI < 1 indicate that there is no
significant risk of non-carcinogenic effects. While, values of HI > 1 imply that there is a
probability of occurrence of non-carcinogenic effects, and are enhanced with increasing
HI values (Candeias et al., 2014)

The toxicity levels for each element were taken from The Risk Assessment Information
System (RAIS). The probability of an individual developing any type of cancer over a
lifetime, as a result of exposure to the carcinogenic hazards, was computed for each
pathway according to Equation (14) (de Souza et al., 2017; Ihedioha et al., 2017). The
carcinogenic risk was estimated by the sum of total cancer risk (Equation (12). A cancer
risk below 1 × 10−6 is considered insignificant. The result of 1 × 10 −6 is classified as
the carcinogenic target risk. If the cancer risk is above 1 × 10−4 it is qualified as
unacceptable (Candeias et al., 2014; Thongyuan et al., 2021).

CDIpathway
HQ = RfD
(12)

𝐻𝐼 = ∑𝐻𝑄 = 𝐻𝑄ing + 𝐻𝑄drm + HQinh (13)

Riskpathway =CDIpathway X CSFpathway

RISK = ∑Riskpathway = Risking + Riskdrm + Riskihn

CDI drm X CSF ing


= CDIing X CFSing x IUR + (14)
ABSgi

3.8.1 Water health risk assessment (WRHA)

Wells water health risk was determined for Zn, Hg and Pb content (Venkatayogi, 2018;
Yin et al., 2021). Exposure was determined with reference values established by USEPA
(2002), WHO (2005, 2017), and Nhantumbo et al. (2012), of Hg 2 µg/l, Zn 3.000 µg/l and
Pb 1 µg/l . Adjustable parameters used were daily water consumption of 1.5 and 2 l/day
78

for children and adults, respectively; exposure frequency number of days exposed in a
year of 365; total years of exposure of 15 and 60 for children and adults, respectively;
average exposure time 5475 and 21900 for children and adults, respectively; dermal
permeability coefficient of Hg, Pb and Zn = 0.0001; surface area exposed of 852.5 and
1610 cm2 for children and adults, respectively; skin adherence factor of 0.2 and 0.07 for
children and adults, respectively. Carcinogenic Risk and non- carcinogenic Hazard were
determined using similar formulas described for soils (Ahmad et al., 2021).

3.8.2 Soil/plant transfer factor (TF)

TF allows to assess transfer of elements from soils to edible parts of plants. Was
calculated as follows eq. 13:

TF = Cplant/Csoil (13)

where Cplant and Csoil are the individual element concentration (mg/kg) in edible plants
and soils samples, respectively. TF > 1 indicates a significant accumulation of elements
in edible plants (Gupta et al., 2021).

3.8.3 Plant health risk assessment

To assess the health risk posed by edible plants consumption, were used the daily intake
of elements (DIM), the hazard risk index (HRI), the targeted hazard quotient (THQ), and
the hazard index (HIplant). The DIM by adults via consumption of edible plants does not
take into consideration the body metabolic response to PTEs, nevertheless can indicate
a possible ingestion rate of a selected PTE. DIM was calculated as follows Eq.14:

DIM = Celement × Dintake/Bweight (14)

where DIM is the daily adult elemental ingestion (kg/day), Celement the plants dry weight
elemental concentration (mg/kg), Dintake the daily consumption of edible plants (all the
vegetables studied were estimated at an average consumption of = 0.025 kg/day which
is the average daily consumption in sub-Saharan African countries(FAO, 2020) and
Bweight the average body weight (60 kg/person) (Ogunwale et al., 2021).

The HRI by ingestion of edible plants for each was calculated by: HRI = DIM/RfD,
where DIM is the daily intake of elements, and RfD the reference oral dose (Cr = 0.003,
Cu = 0.037, Fe = 18, Mn = 0.14; Ni = 0.2, Pb = 0.014, Ti = 0.7, Zn = 0.3, Zr = 4.2 mg/
kg/day) (USEPA, 1989)
79

THQ is the ratio between the PTE dose to a reference dose (dimensionless) indicating
the non-carcinogenic health risk through edible plants ingestion. It is calculated by
Eq.15:

THQ = (EF × ED × Dintake × Celement)/(RfD × Bweight × AT) (15)

where EF is the exposure frequency (365 days/year), ED the exposure duration (40
years), and AT the average

exposure time (ED × 365 days/year). The hazard index (HIplant) is the summation of all
the PTEs considered, i.e. EQ. 16

HI =Σi n=1 THQn , i = 1,2,…,n (16) (Guadie et al., 2021; Lee et al., 2021).

HRI, THQ and HI values < 1.0 suggest no risk posed by plant consumption, while values
≥ 1.0 indicate a certain level of risk of adverse health effects to occur (Gebeyehu et al.,
2020, and references within).
80
81

Part II - Results and Discussion


82
83

4 Paper 1 (Appendix 2)
Characterization of the Dynamics of Leachate Contamination Plumes in the
Surroundings of the Hulene-B Waste Dump in Maputo, Mozambique

Abstract
The contamination of areas around solid urban waste dumps is a global challenge for
the maintenance of environmental quality in large urban centers in developing countries.
This study applied a geophysical method (electrical resistivity) to identify leachate
contamination plumes in the subsoil and groundwater, as well as to describe their
temporal dynamics (2020 and 2021) in the surroundings of the Hulene-B waste dump,
Maputo, Mozambique. Eight 400 m electrical resistivity profiles were performed, four
profiles in January 2020 and four profiles in May 2021 overlapped, and the data were
inverted with RES2D software. The electrical resistivity models predominantly indicate
an E-W movement of large contamination plumes that are successively diluted with
saturated media and groundwater, creating zones of less resistive anomalies (<4.2–8.5
Ω·m) possibly contaminated at the two analyzed seasons, between 2020–2021. The
thickness of the contamination plumes was higher in summer (2020) for profiles 1 and 2,
and we associate it with the production and migration mechanisms of leachate that are
intense in the hot and rainy season. Southwest of the dump, profile 4b showed the
propagation of anomalous areas on the surface and at depth, which are associated with
the production of leachate resulting from the continuous decomposition of waste that is
continuously deposited in a new area southwest of the dump, thus generating a slow and
continuous migration of leachate at depth, mainly in winter (2021). The spatial distribution
of contamination plumes during both seasons was reduced significantly farther away
from the waste deposit, revealing the attenuating effect of groundwater and lithological
substrate.

Keywords: plumes; dynamics; resistivity; contamination; groundwater

Reference: Bernardo, B.; Candeias, C.; Rocha, F. Characterization of the Dynamics of


Leachate Contamination Plumes in the Surroundings of the Hulene-B Waste Dump in
Maputo, Mozambique. Environments 2022, 9, 19.
https://doi.org/10.3390/environments9020019
84

4.1 Geophysical Studies


Electrical Resistivity The interpretation of the electrical resistivity models allowed for the
understanding of the leachate formation areas, dynamics, and dispersion of
contamination plumes in the groundwater as well as for comparing the variations of the
resistive anomalies (2020–2021). The following anomalous areas were distinguished: (I)
areas of possible leachate formation and enrichment; (II) contamination plumes in
subsurface and groundwater.

4.1.1 Profile 1: 2020 (a) and 2021 (b)

The two profiles (a) and (b) extend in the S–N direction and are parallel to the western
boundary of the dump (Fig. 4.1).

Figure 4.1. Electrical resistivity model of profile 1, 2020, (a) and 2021 (b).

Zone 1 - Along the first 200 m of both profiles, no significant changes are noted near
the surface, and in general, the resistivity is higher and associated with the rubble and
de-bris of old houses that were built in this space until 2018 (VOA, 2018). At depth, slight
differences in resistive layers are noted, which may be associated with the effect of
variation in rainfall and moisture, more abundant in (a) and less in (b). The zone of
resistivity < 19.2 Ω·m in (a) and < 15.4 Ω·m (b), which we consider as the leachate
concentration zone, resulting from the vertical migration of leachate from the surface
washing of the dump by precipitation being channelled in the unprotected ditch and
parallel to these profiles (Fig. 4.2). Wu et al., (2021) and Ololade et al., (2019) showed
that leachate flow in non-isolated areas can cause migration to great depths and
85

groundwater contamination. The leachate accumulation zones, < 19.2 Ω·m (a) and <
15.4 Ω·m (b), show variable thicknesses, which may be associated with the effect of the
relatively saturated lower layers, < 16.1 to 11.7 Ω·m (a), being more extensive in
summer, which causes slow vertical migration at depth and the production of thick
plumes at great depth, < 8.46 Ω·m in profile (a) and < 8.56 Ω·m in profile (b). The leachate
migration layers can be considered as contaminated, < 16.1–11.7 Ω·m (a) and < 13.2–
10.7 Ω·m (b). At great depth, at the southern end of both profiles, there is a localized
anomalous zone that we interpret as being influenced by groundwater contamination by
plumes resulting from vertical leachate migration.

Zone 2 - In both profiles, anomalous zones are evident near the surface at 280 m on-
wards, which are more extensive in profile (b), representing residues humidified by sur-
face water (leachate producer) < 15.4 Ω·m, which are more visible in the north of both
pro-files and more extensive in profile (b). The clearly visible plumes in both profiles, in
the surface water, correspond to the contamination in the natural receiving basin by
surface leachates and plumes mobilizing at depth in the E–W direction, described in
profile 2.

Figure 4.2. Geophysical surveys: (a) profile 1 in 2020 the arrow indicates the surface
leachate concentration ditch parallel to the profile 1; (b) drainage ditch with uninsulated
surface leachate 2021; (c) southern section of profile 1 in 2021.

4.1.2 Profile 2 2020 (a) and 2021(b)

Profile 2, with a west–east orientation, and northern boundary of the dump (Fig. 3).
86

Figure 4.3. Electrical resistivity model of profile 2, 2020 (a) and 2021 (b).

Zone 1 - From 0 to 80 m surface, in profile (a) we can notice a marked variation of


resistivities. The zones of low resistivity (< 24.5 Ω·m), we interpret as a wet and leached
waste production zone, given the new solid waste depositions recorded at this point to
the west of the basin. The same profile section (b) shows high resistivities, given the
scarcity of precipitation and low ambient temperature. The high temperatures, humidity
and age of the dump are primary factors in decomposition and leachate production ,
(Helene et al., 2021; Nta et al., 2020; Fatoba et al., 2021). From 8.35 m depth
downwards, in profile (a), we note an extensive saturated zone that extends in the
northern direction and connects to an extensive subsurface flow system. In turn, in profile
(b), this saturated zone appears more confined and with an expressive concentration of
anomalous values < 4.96 Ω·m (b), which we interpret as a plume migrating horizontally
in summer in the east–west direction and being confined at this point. (Cendón et al.,
2020) described the aquifer system of this region as semi-confined that seasonally binds
continuously, a fact that is visualized in these bands of the two profiles and that are
associated with the transfer of possible plumes at depth. From 80 to 200 m, superficial
in profile (a), we verify the alternation of high and low resistivities. The high resistivities,
we interpret as compact material on the surface, which alternates between debris,
rubble, and old house debris. From 160 to 200 m (a), we observe surface waters of the
natural leachate reception basin, which are constantly enriched by the surface leachates.
In this section at a depth of 8.35 to 47.7 m (a), there is an extensive plume of
87

contamination arising from the large mass of wet waste and producing leachates < 24.5
Ω·m (a) that migrate horizontally in an east–west direction. Complex mechanisms of
leachate movement from the surface and at depth are also observed. The vertical and
horizontal movement of the leachate produces an extensive plume migrating E–W, <
6.21 Ω·m (a). In profile (b), from 80 to 140 m, the resistivity is higher along a larger area
than in (a), given the significant reduction in the extent of surface water responsible for
the decomposition of the waste mass and consequent decrease in resistivity. At the
depth of the same section, confined groundwater receives leachate, which moves
horizontally from E–W < 16.8 Ω·m (b) and a vertical migration between 142.5 to 147.5
m. The contamination plume < 7.9 Ω·m (b) at this point is quite pronounced and may
indicate a high level of contamination, given the reduced dilution and migration
environment.

Zone 2 - In (a) between 200 m to 315 m, it shows alternating resistivities between the
less moist waste mass < 20.7 Ω·m to saturated zones < 10.6 Ω·m (b) which represent
pits and small surface depressions enriched by surface water, and the same
characteristics are noted in (b), but with lower moisture extent. From 320 m onwards, in
profile (a), an increase in resistivity > 30.8 Ω·m is noted which corresponds to waste
mass mixed with less moist soils and resistivity > 52.8 Ω·m which represents compacted
dry waste. In profile (b), to the same extent, the increase in resistivity is much more
noticeable, which indicates dry waste and soil (< 33.4 Ω·m) and compacted waste (>
33.4 Ω·m). At depth, in both profiles, the large mass of waste, < 30.94 Ω·m (a) and <
20.74 Ω·m (b), gains successive moisture in (a), establishing an extensive area of
leachate production < 24.5 Ω·m (a), whereas in (b), it is confined (< 16.8 Ω·m). These
differences are the result of the variation in precipitation and temperature in the two
seasons studied, which are responsible for the variation in waste mass decomposition,
production, and migration of leachate (Barry et al., 2021; Harjito et al., 2018).

4.1.3 Profile 3 2020 (a) and 2021 (b)

Profile 3, with an S–N orientation northwest of the dump (Fig. 4.4).


88

Figure 4.4. Electrical resistivity model of profile 3 in 2020 (a) and 2021 (b).

The execution area of this profile corresponds to the strip that temporarily floods with
run-off water from the dump. It was executed to understand the spatial dynamics of
possible contamination plumes in the northern direction of the surroundings of the dump
(reception basin). At 40 m, in both profiles, the influence of moisture in a localized band,
at depth, is noted, evidenced by resistivity < 23.5 Ω·m in profile (a) and 21.6 Ω·m (b). In
the first profile, the strip occupies a relatively larger area due to the abundant precipitation
in this period that infiltrates to the deeper layers. At 140 m depth, there is an anomalous
zone in both profiles, which we interpret as a plume of contamination, < 7.18 (a) and <
7.34 Ω·m (b). We consider that this anomaly corresponds to the northern limit of the large
plume described in profile 2. From 160 m onwards, in both profiles at all depths, the
resistivity tends to be equal, and we consider as typical of local strata and anomalous
(positively) zones, > 38.2 Ω·m (a) and > 33.7 (b), corresponding to compacted soils or
rubble, given the strong movement of cars at these points. The similarity of the data in
both profiles and the absence of low depth anomalous zones show the decreasing effect
of the contamination plumes on the subsoil and the subsurface environment, as one
move away from the dump to the North, due to the natural process of attenuation by
dilution and dispersion of the plumes in the natural receiving basin. Similar situations
have been described in areas around several dumpsites in Africa, Morocco by El Mouine
et al. (2021) and Touzani et al. (2021); Nigeria by Fatoba et al. (2021) and Burkina Faso
by Barry et al. (2021).
89

4.1.4 Profile 4 in 2020 (a) and 2021 (b)

Profile 4 has SW–NE orientation (a), and the NE–SW (b) profile both overlap (Fig. 4.5).

Figure 4.5. Electrical resistivity model of profile 4 in 2020 (a) and 2021 (b).

Zone 1 - Profile 4 (a) from the beginning to 120 m shallow shows zones with anomalous
resistivities. The resistivity < 13.7 Ω·m was interpreted as a zone of leachate production
and dispersion, resulting from new waste deposits in the southwest of the natural basin.
These leachates migrate horizontally in the NE–SW direction and vertically until they mix
with groundwater. In profile 4 (b), these anomalies occupy a large area in the dry period
and extend over an area of about 240 m surface as well as at depth, < 16.3 Ω·m. The
extensive anomalous area at depth in profile 4 (b) (< 16.3 Ω·m) may be associated with
a vertical and continuous migration of leachate accumulated at the end of the rainy
season in the southeast of natural reception basin.

Zone 2 - In profile 4 (a) from 200 to 400 m, the surface resistivities are generally higher
and we interpret them as plastic waste, rubble, and debris from old, buried houses. The
same occurs in profile 4b, from 140 m to the end of the profile. The resistivity
corresponding to possible contamination plumes were interpreted as < 6.42 Ω·m in (a),
<8.9 Ω·m in (b), which at great depth, did not show great changes in the two seasons of
the years studied. This reality can be associated with the local aquifer system, which is
described as semi-confined in clay layers (Cendón et al., 2020).
90

4.2 Spatial Distribution of Possible Leachate Plumes (2020–2021)

The spatial distribution of the plumes in the study period (2020–2021) (Fig. 4.6), shows
predominantly an east–west movement of the deep contamination plumes and surface
leachates in profiles 1 and 2, which mix with the surface waters of the natural basin to
the west and migrate vertically to the groundwater, causing resistive anomalies at a
larger scale in winter (1). The same flow direction was described by (Bernardo et al.,
2022a). The flow is conditioned by the arrangement of the local topography (soft dune
E–W), which collects all the leachate from the subsurface compression of the waste,
since it is at a higher altitude than the local surface. Another relevant factor in this
direction of contaminant flow is the recent construction of leachate drainage channels,
which create natural leachate flows in the E–W direction (Figure 4b). However, the
dilution process of leachate plumes shows a varied dispersion in the natural reception
basin and groundwater (Figure 6. < 4.26 – < 8.5 Ω·m), suggesting that contamination
migrates in several directions. Similar anomalous values were interpreted as plumes in
groundwater by Harjito et al. (2018)3–9 Ω·m near a Bantul waste dump in Indonesia,
Bichet et al. (2016) refer between 5–12 Ω·m at a landfill in Belfort (France), and Ugbor
et al. (2021) registered values of 3.12–8.7 Ω·m in the surroundings of the Onitsha waste
dump in southeast Nigeria. To the southwest of the natural basin reception, there is a
subplume of contamination resulting from new depositions producing leachate that
migrates to the southwest during the wet season and in all directions during the dry
season (profile 4). The spatialization of the plumes shows that as one moves away from
the dump, the anomalies tend to disappear (profile 3), suggesting the attenuation of the
contamination by the local lithology. The role of lithology and groundwater in the
attenuation of contamination by leachate plumes in the surrounding of waste dumps has
been described in many studies, for example by (Fatoba et al., 2021) and (Biosca et al.,
2021).
91

Figure 4.6. Possible flow directions of contamination plumes 2020–2021, modified from
Bernardo et al. (2022a)
92
93

5 Paper 2 (Appendix 3)

Integration of electrical resistivity and modified DRASTIC model to assess


groundwater vulnerability in the surrounding area of Hulene-B waste dump,
Maputo, Mozambique

Abstract

In this study, electrical resistivity was applied in six 400 m profiles around Hulene-B
waste dump (Mozambique). After, an inversion was performed by RES2Dinv. The use
of electrical resistivity method allowed to characterize in detail some underlying aspects
for DRASTIC index, by identifying anomalous zones considered to be permeable and
prone to leachate migration. The modified DRASTIC index revealed high values in areas
near contaminated surface groundwater and surface layers of the vadose zone,
characterized by low resistivities. Areas with lower index results were characterized with
high resistivity on surface layers and high depth at which groundwater was detected. The
overall modified DRASTIC index result revealed medium vulnerability. However, high
vulnerability index values were detected in areas with higher surface elevation,
suggesting groundwater contamination by horizontal dilution of leachates from the
surrounding area of Hulene-B waste dump.

Keywords: Resistivity; anomalous zones; modified DRASTIC model; groundwater


vulnerability

Reference: Bernardo, B; Candeias, C; Rocha, F. (2022). Integration of Electrical


Resistivity and Modified DRASTIC Model to Assess Groundwater Vulnerability in the
Surrounding Area of Hulene-B Waste Dump, Maputo, Mozambique. Water 14, no. 11:
1746. https://doi.org/10.3390/w14111746
94

5.1 Results and discussion


5.1.1 Resistivity models and potential contamination risk

To the analysis of the profiles, the resistivity values of the profiles were adjusted to the
same scale so that each the color of the contour in the resistivity model implies the same
resistivity value (Fig. 5. 1). In this study, the electrical resistivity models were analyzed
to identify the possible influence of leachate on groundwater contamination. Thus,
anomalous zones that may reflect leachate migration and contamination process were
identified: (i) Leachate generation and migration areas (7.99 – 16.8 Ω.m); (ii) saturated
zones contaminated by leachate (4.96 - 7.99 Ω.m), (iii) groundwater and surface water
contaminated by leachate (1.535 - 7.99 Ω.m) (iv) Waste and lithologies local (> 16.8 Ω.m)

Figure 5.1. Calibrated color scale showing the resistivity range of materials and their
characteristics.

Profile 1, from south to north of the dump (Fig. 5.2.a), is superficially characterized by
high resistivity associated with rubble, house debris and compacted waste in non-wet
environments from 0 to 35 m. From 240 to 280 m there were zones with a rather
heterogeneous anomalous resistivity, which were considered as leachate generation
and migration zones, (7.99 - 16.8 Ω.m) given the strong accumulation of surface
leachates resulting from the E-W movement (profile 2) that are diluted with the surface
waters, causing possible contamination (280 to 400 m) (1.535 - 7.99 Ω.m ). At medium
depths 30 - 56.5 we note the predominance of a vast layer throughout the length of the
profile that we interpret as TPv lithologies, less resistive (12.9 - 16.8 Ω.m) given the
possible influence of horizontal migration at depth described in profile 2. At deeper levels
between 47 and 56.5 m, anomalous zones are observed, which were considered as
lithologies influenced by horizontal leachate migration and possible contaminated
groundwater (> 45 m depth) (1.535 - 7.99 Ω.m).

Profile 2, along W-E direction (Fig.5.2.b), is quite heterogeneous, with high surface
resistivities from 0 to 140 m, which represent for rubble, old house debris, waste buried
in non-wet environment west of the dump, followed, from 140 to 160 m, by a zone of
possible migration of surface leachate into the subsurface environment, causing an
95

extensive zone of subsurface anomalies, which were consider as lithologies


contaminated by strong horizontal migration of leachate with E-W direction (7.99 - 16.8
Ω.m) and satured zone of contaminated groundwater (1.535 - 7.99 Ω.m), differentiated
levels of semi-confinement of aquifers. From 245 to 400 m depth there were local
anomalies and lithologies with higher resistivities, corresponding to highly compacted
waste with diverse contents.

At depths ranging 30 to 160 m, semi-confined anomalies (< 8.35 m depth) and aquifers
(> 10 m depth) separated by semi-saturated layers were noted, demonstrating the
existence of a possible continuous connection between the two. These characteristics
were described as conducive for groundwater contamination at various depths, mainly
in the surroundings of Hulene-B waste dump, where surface leachate flows were noted
with successively enrichment of lithological layers (7.99 - 16.8 Ω.m).

Profile 3 (Fig. 5.2.c) north of the dump in S-N direction, to study possible dynamics of
groundwater contamination, ~300 m away of the dump. The profile at surface level
exhibited generally high resistivities, alternating between rubble and highly compacted
soils. At the deeper level (> 40 m) in the southern end an anomalous zone was found,
considered as contaminated lithologies (7.99 - 16.8 Ω.m). This anomaly was associated
with the horizontal and vertical migration of leachate described in profile 2.

Profile 4, in the NE-SW direction (Fig. 2.d), from the starting point to 150 m, showed
generally high resistivities associated with compacted residues and rubble of old houses.
From 150 m depth to the end of the profile, a continuous decrease in resistivity was
observed, which can be associated with saturated and wet areas with origin in natural
receiving basin where new waste deposits were observed, a localized source of dilution,
and vertical migration of leachate and groundwater contamination (< 7.99 Ω.m).

Between 300 and 310 m, relatively high resistivities were found, associated with waste
deposited in the intra dune depression with low surface moisture level. From 320 m to
the end of the profile, a shallow aquifer was evident (< 7.99 Ω.m), as this area
corresponds to the end of the depression (SW) with a surface covered by humid soils,
evidenced by low surface resistivities. Between 6 - 56 m in depth, was noticed a large
anomaly (16.3 - 7.99 Ω.m) related to surface influence by leachates propagated to
greater depths (> 37.4 m) causing possible groundwater contamination (< 7.99 Ω.m).

Profile 5 with west-east orientation, was at the southern end of the dump (Fig. 2.e). The
electrical resistivity results didn’t displayed significant changes at surface level, and
showed higher resistivities associated to compacted soils and rubble (including road
96

asphalt where the profile was executed). This road, besides being in the southern
boundary of the dump, is an access to the interior of Hulene-B neighborhood. Between
130 and 145 m (> 40 m depth) extended a resistive zone of low anomalous values, which
were interpreted lithologies that might influence leachates migration (7.99 - 16.8 Ω.m
Ω.m) and possible water contamination by the vertical movement of leachates (1.535 -
7.99 Ω.m).

Profile 6, this profile was taken at the eastern limit of the dump, along S-N direction,
between the dump and Julius Nyerere Avenue (Fig. 5.5.f). From the starting point to 200
m, there were resistivities with average profile values (12.9 – 33.4 Ω.m). These
resistivities suggested soils with different levels of compaction and surface moisture that
may be associated with various activities south of the dump. The resistivity 33.4 – 53.8
Ω.m corresponding to a superficial but thick layer, between 160 to 200 m, indicated
compacted soils at the entrance of the dump. From 8.6 m depth, higher resistivity values
(> 70.3 Ω.m) were noted which may be related to sandstone stratum with different levels
of alteration, typical of the TPv formation (Momade et al. 1999). Onwards, in the northern
direction, resistivity starts to decrease successively (< 27.05 Ω.m), along thick layers with
moisture levels that increase until to groundwater (< 10.44 Ω.m).

The saturated area occupied a large space, revealing the existence of an E-W
groundwater flow parallel, to the dune slope where the leachate is located. The
groundwater contamination process at this point may be occurring horizontally due to
leachate diffusion, causing localized anomalous resistivities (< 10.445 Ω.m) close to the
groundwater resistivity (< 7.99 Ω.m). From 240 m, resistivity begins to decrease,
generating localized anomalous zones in the subsurface, which are associated with
vertical leachate migration, pointing the occurrence of two isolated "hot spots". Between
240 and 280 m, below the first "hot spots", there was a tendency for a significant increase
in resistivity, that may correspond to less saturated layers up to the least conductive
stratum (> 33.4 Ω.m). Profile data showed the existence of two mechanisms of possible
satured zone and groundwater contamination which were, horizontal dilution in the south
and center of the profile, and vertical migration (< 16.8 Ω.m) and retention of leachate in
localized "hot spots".
97

Figure 5.2. Profiles of the variation of the electrical resistivity in the study area
98

5.2 Modified DRASTIC index


5.2.1 Depth to water table (D)

Aller et al. (1987) refer that the depth of the water table determines the depth through
which a contaminant moves before reaching the aquifer and determines the contact time
with the surrounding media. Thus, a greater possibility of contamination mitigation occurs
when the depth of the water table is greater because deeper water table implies more
travel time and less vulnerability to contamination (Gemail et al., 2017). Thus, the deeper
the phreatic table implies more travel time and less vulnerable to contamination (Iddrisu
et al., 2021; Koda et al., 2017). On the eastern, southwestern, and northern boundary of
the dump (area covered by profiles 6,4 2 and 1), Subsurface waters and groundwater
was detected between 1.5 - 4.6 m, and on the southern, north and western boundary at
depths > 30 m (area covered by profiles 5 and 3). Results for the eastern border
southwest and northwest were similar to those published by (Muchimbane, 2010), who
classified the predominant aquifers in Maputo city as shallow and phreatic aquifers,
estimating their average depth between 1.5 and 9.3 m. DRASTIC parameters of 9 and 1
respectively were assigned, and D = 5 (Tables 5. 1 and 2).

5.2.2 Net Recharge (R)

Net recharge is the amount of surface water that infiltrated into underground and reaches
groundwater (R. Ghosh et al., 2021), indicating the amount of water from precipitation
that was available for vertical transport, dispersion, and dilution of pollutants from a given
application point (Marino et al., 2020; Khattak et al., 2021). Recharge water in dumps
surrounding is a source of contaminants transport within the vadose zone to the
aquifer(Islami et al., 2020). The greater the recharge, the more vulnerable the
groundwater (Hasan et al., 2021). Given the small area analyzed in this study, data used
according to Momade et al.(1996) and Vicente(Vicente, 2011), that estimate the value of
groundwater recharge in Maputo city between 165 and 185 mm/year for the entire
Hulene-B dump surrounding area. Value of 8 was assigned to the area and R = 4 (Tables
5. 1 and 2).

5.2.3 Aquifer media (A)

Aquifer media refers to a rock in the ground that serves as water storage (Ersoy et al.,
2013). It indicates materials property that controls pollutant attenuation processes based
on permeability of each layer (Paul et al., 2021). The attenuation characteristic of the
aquifer material is reflected by the mobility of contaminants through aquifer media (Asfaw
99

et al., 2020). In the surroundings of the Hulene-B dump, two types of semi-confined (west
of profile area, 2) and shallow (south-west of profile area 1, 4 and area covered by profile
6) aquifers were assumed to exist, which have been described by Momade et al.(1996),
Vicente (2011), and Cendon et al. (2020), composed of inland dune sands and semi-
permeable sands, and recharge occurs mainly by precipitation given the permeable
surfaces such as dune sands. The value of 7 was assigned to the area around the
Hulene-B dump and A = 3 (Tables 5. 1 and 2).

5.2.4 Distance of the anomalous surface layer and groundwater (S).

The distance between anomalous surface layer (low resistivities) generally represents
surfaces contaminated by leachates in areas close to landfills (Lau et al., 2019).
Anomalous surface to groundwater bands areas were characterized by intense leachate
migration, which was evidenced by transected profiles 2 (< 10.445 Ω.m) and 6 (< 16.8
Ω.m). The area north of profile 1 and southwest of profile 4 show intense anomalies that
we interpret as a shallow aquifer (4) and surface soils enriched by leachates (1) that
accumulate successively to the west that can easily migrate into the confined aquifer
described in (2). However, other profiles did not showed bands with the continuous
connection of anomalies and groundwater. The area covered by profiles 1,2, 4 and 6
was assigned the value 10 and other areas value 1, and S = 2 (Tables 5. 1 and 2).

5.2.5 Topography (T)

Topography refers to the slope of an area (Asfaw Mengistu, 2020). It controls the
probability of a pollutant to be transported by runoff or to remain in the soil where it may
be infiltrated (Aller et al., 1987). The softer the slope (slope of 0-2 %), higher the water
and/or pollutant holding capacity, while in slopes > 10%, lower water and/or pollutant
holding capacity occurs (Aller et al., 1987). In the surroundings of the dump, the relief is
heterogeneous (Fig. 2). The area covered by profiles 1, 3, 4 and 6 did not present slopes.
However, the area covered by profile 2 and 5 have slope > 10%. Thus, the area covered
by profiles 2 and 5 was assigned the value 5 and the other areas were assigned the
value is 10, and T = 1(Tables 5. 1 and 2).

5.2.6 Vadose (I)

Vadose zone is the unsaturated zone that lies below soil horizon and above water table
(Asfaw et al., 2020). It determines the attenuation characteristics of the contaminants
(Aller et al., 1987). The movement of contaminants into the saturated zone is controlled
100

by this parameter. Aller et al. (1987) and Asfaw et al. (2020) refer that if the flood zone
consists of sand, the potential risk of contamination of the aquifer is very high. The
surroundings soils of Hulene-B dump are dune sands. However, the combination of
electrical resistivity data, showed specific characteristics, which allowed to classify in
detail the environment of the dump. The area covered by profiles 3 and 5 was
characterized by having very resistive surface layers, showing low infiltration rate, so a
value of 4 was assigned. Profile 1 was heterogeneous, with half of the area covered
being very resistive(south) and with very low resistivities, the assigned value was 8.
Profiles 2, 4 and 6 had a lower resistivity, showing higher infiltration, which is typical of
sandy soils, so a value of 8 was assigned, and I = 5 (Tables 5. 1 and 2).

5.2.7 Hydraulic conductivity (C)

Hydraulic conductivity is described as the ability of materials to transmit water to aquifers,


in turn controlling the rate of groundwater and contaminant material flow under a given
hydraulic gradient (Gemail et al., 2017). It controls contaminant migration and dispersion
from the injection point within the saturated zone (Asfaw et al., 2020).The surroundings
of the Hulene-B waste dump, the hydraulic conductivity was estimated by Momade et al.
(1996) as 1-5 m/d. For the whole studied area, the value of 2 was assigned, and C = 3
(Tables 5. 1 and 2).

Table 5.1 Parameter values considered in the DRASTIC Index

Factor Characteristics of the surroundings of P1 P2 P3 P4 P5 P6 Mean


the waste dump
D Depth of groundwater level 45 45 5 45 5 45 31.6
R Recharge capacity 32 32 32 32 32 32 32
A Sands 21 21 21 21 21 21 21
S Distance of anomalous surface layer 14
20 20 2 20 2 20
T and groundwater
I Plan, soft dune and interdune 8.3
10 5 10 10 5 10
C depression
Sands 40 40 20 40 20 40 33.3
Hydraulic conductivity 6 6 6 6 6 6 6
DRASTIC index: 174 169 96 174 91 174 146.3
P – Profile
101

5.3 Descriptive statistics of electrical resistivity and DRASTIC index

The electrical resistivity values of the areas covered by the profiles were projected with
the vulnerability index values (Table 5.2). In general, the areas covered by profiles 2, 4
and 6 were classified as having high DRASTIC index. The profile areas 2 and 4, with
mean resistivity values of 20.2 and 18.1 Ω.m, respectively, suggested predominance of
lower resistivity across profile surface area which extends into the groundwater,
suggesting successive leachate migration.

Table 5.2 mean, maximum, minimum, standard deviation of resistivity values (Ω.m) and
DRASTIC index

ID min max mean SD modified DRASTIC


P1 8.64 42.36 20.78 5.72 174
P2 5.37 119.1 20.22 9.31 169
P3 10.12 50.8 31.01 7.3 96
P4 6.98 41.19 18.1 4.64 174
P5 1.04 477.1 37.35 31.86 91
P6 3.06 207 49.2 32.78 174
SD – standard deviation

The area covered by profile 6 had an average of 49.2 Ω.m, minimum of 3.06 Ω.m and
maximum of 207 Ω.m. The average resistivity value was relatively higher but showed a
much higher standard deviation (32.78 Ω.m), revealing resistivity heterogeneity, marked
by the existence of surface anomalous bands that connect with the groundwater at
various points, which may be associated with a greater migration of leachates and
groundwater contamination.

Areas covered by profiles 3 and 5 revealed low DRASTIC index and heterogeneous
electrical resistivity. Thus, the area covered by profile 1 showed heterogeneity in surface
and subsurface anomalies that may be associated with vertical and horizontal migration
of contaminants with minimum resistivity values of (1.535 - 7.99 Ω.m) revealing a higher
risk of contamination of the semi-confined aquifers described in profile 2. However, the
greater depth at which the groundwater was detected reveals a natural attenuation
mechanism of contamination by the underlying lithologies. The area covered by profile 3
showed an average resistivity of 31.01 Ω.m, with ranging 10.12 to 50.8 Ω.m. These
results suggest the predominance of high resistivity values, which is translated by
reduced predominance of resistive surfaces that were interpreted as less permeable and
less contaminated substrates. The minimum resistivity at great depth (7.99 - 10.12 Ω.m)
may be associated to a saturated area or localized influence of horizontal migration of
102

contaminants, described in profile 2. The area covered by profile 5, presented an


average resistivity of 37.35 Ω.m, ranging 1.04 to 477.1 Ω.m. Groundwater was detected
at depths > 40 m, with less risk of contamination by vertical migration.

The low DRASTIC values in areas covered by profiles 3, and 5 resulted from the
combination of two factors, high resistivities prevailing in the surface lithologies
suggesting low infiltration, which greatly reduces the risk of vertical migration of leachate
to deep layers, and greater depth at which, resistivities interpreted as groundwater, were
found. In general, data showed areas covered by profiles, with high average resistivity,
with lower DRASTIC index (profiles 3 and 5) (Fig. 5.3). Exceptionally, the area covered
by profile 6 showed high DRASTIC index, with relatively high mean resistivity values due
to its higher standard deviation and predominance of resistivities < 10.45 Ω.m in a large
surface strip that was associated to leachate migration to groundwater at low depth than
in all profiles. However, in area covered by profiles 1, 2 and 4, the presence of resistive
anomalies was considered as semi-confined and shallow aquifers and surface anomalies
that are conducive to contaminant migration were determining factors for high DRASTIC
index.

Figure 5.3. Electrical resistivity values (minimum, maximum, mean, standard deviation
Ω.m) and modified DRASTIC Index. Maximum value of profile 5 (477.1 Ω.m) was set to
(200) to maintain the scale within limits of perceived index.

5.4 Data Integration

The resistivity data of the areas covered by the profiles 1, 2, 4 and 6, with predominance
of surface and groundwater resistivity anomalies, was described in other studies
suggesting leachate migration (Parvin et al., 2021), and groundwater contamination (El
103

Mouine et al., 2021), mainly in areas surrounding non-isolated dumps where leachate
can freely circulate through adjacent lithologies and subsequently affecting vadose zone
and groundwater (Wysocka et al., 2017). Surface circulation and leachate infiltration is
well evidenced in profiles 1, 2, 4 and 6, by surface and subsurface resistive anomalies,
indicative of contamination (Netto et al., 2021).

In the area covered by the profiles 1(in north), 2 and 4, continuous anomalies from
surface to groundwater were noted, described as lithological migration bands of leachate
to groundwater (Yap et al., 2022; Chetri et al., 2021). The area transacted by the profile
6, besides showing an extensive layer with low resistivities (< 10.445 Ω.m), has been
considered, in similar studies, as saturated zone and groundwater under leachate
influence (Adamo et al., 2020; Brahmi et al., 2021). In the north of the dump, two localized
points of resistive anomalies, were identified as "hot spots" (Ololade et al., 2019;
Marques et al., 2021), resulting from vertical migration and localized accumulation of
contaminants Feng et al. (Feng et al., 2021), and may be associated to confined aquifer
system, described in this zone as quite vulnerable to contamination given its proximity to
the surface and sandy characteristics of the vadose zone (Cendón et al., 2020).

The electrical resistivity data of the areas covered by the profiles 3 and 5 were
characterized by high resistivities, and associated to low contaminant infiltration capacity
(Gemail et al., 2017), given high surface compaction (Udosen, 2021). The area
transacted by the profile 3 is located far of the waste dump (> 350 m), revealing that
contaminants may not be reaching this area. Morita et al. (2021) and Touzani et al.
(2021) demonstrated that resistivity increase away from dumpsites represents a
significant decrease in contamination, due to attenuating role of soils and groundwater.
Groundwater at great depths in area covered by profiles 1, 3 and 5 has been described
in similar studies as a determinant to low contamination risk (Oke, 2020; George,
2021)which partly explains the low DRASTIC index in these areas. Pau et al. (2021) and
Boumaiza et al. (2021) mentioned that the depth of groundwater and the characteristics
of the infiltration zone are the most important factors determining the DRASTIC risk.

Gemail et al. (2017), Shah et al. (2021) and Nasri et al. (2021) integrated electrical
conductivity and DRASTIC data and concluded that areas with highest DRASTIC index
around dumps, generally exhibit low electrical resistivity associated with the migration of
contamination through adjacent lithologies that may subsequently reach groundwater.

The spatial projection of Vulnerability Index in the surroundings of the Hulene - B dump
revealed to be higher to the east, southwest and northwest of the dump, with high and
104

transient relief, and lower to the south and west of the dump with low elevation, except
for the southern area which is transient (Fig.5.4). Tan et al. (2020) and Blarasin et al.
(2021) have reported that contamination of the upper levels of water tables leads to the
dispersion of contaminants to larger areas.

Figure 5.4. Spatial projection of the modified DRASTIC in surroundings of the Hulene -
B dump: dashed in white represents low vulnerability and in black high vulnerability

The combination of electrical resistivity data for the assessment of groundwater


vulnerability indices are an important tool for DRASTIC factors assessment, such as:
groundwater depth, and intrinsic characteristics of vadose zone. The modification of
"soil" factor in the original DRASTIC by the "distance between the anomalous surface
layer and groundwater" is important in areas with potential contaminant migration to
groundwater, as well as in studies aiming to outline remediation measures in areas of
groundwater contamination flow.

Measured groundwater depth in northern well (WN) was 6.5 m, and southern well (WS)
was 5.1 m similar to electrical resistivity model (Table 5. 3). However, The WS location
was further south of the profile 4 area and results showed that groundwater levels tended
to be more at surface when approaching intradunar depression. Groundwater pH and
PO43– was 6.1, 1.33 mg/l in WS and, 8.4 and 0.43 mg/l, respectively. Phosphate results
were above natural groundwater reference value of 0.005 to 0.05 mg/l (WHO, 2017).
Previous studies suggested that active waste dumps have a phosphate production of 1-
100 mg/l, that can be incorporated into groundwater by leaching (Akinbile et al., 2011
and Pan et al., 2021). The prolonged consumption of water contaminated with high levels
105

of phosphates can damage blood vessels, damage kidneys, cause osteoporosis and
induce ageing processes (Isiuku et al., 2020)

Table 5.3 Data validation of electrical resistivity and modified DRASTIC

RfD
ID Depth pH PO43− RfD (WHO, 2017)
(WHO,2017)
WS 5.1 m 6.1 1.33 mg/l
6.5–8.5 0.005–0.05 mg/l
WN 6.5 m 8.4 0.43 mg/l
RfD - Natural reference value
106
107

6 Paper 3 (Appendix 4)

Soil properties and environmental risk assessment of soils in the


surrounding area of Hulene-B waste dump, Maputo (Mozambique)

Abstract

Soils in areas surrounding landfills are constantly being enriched by heavy metals
contained in the leachates, which can subsequently migrate to groundwater. The present
investigation aims to characterize soil properties of 71 soil samples collected in the
surroundings of Hulene-B waste dump and to determine the landfill pollution index (Ip).
Soils properties studied were texture, pH, electrical conductivity, organic matter, color,
and moisture. Results revealed that soils properties in the surroundings of Hulene-B
waste dump were significantly altered when compared to local background. Ip index
classified these soils with very high pollution, indicating a possible migration of
contaminants to subsoil and groundwater, suggesting the need for intervention to
mitigate the impact.

Keywords: soils; waste dump; soil properties; environmental risk assessment.

Reference: Bernardo, B., Candeias, C., Rocha, F. (2022). Soil properties and
environmental risk assessment of soils in the surrounding area of Hulene-B waste dump,
Maputo (Mozambique). Environ Earth Sci 81, 542 https://doi.org/10.1007/s12665-022-
10672-7
108

6.1 Results and discussion


Studied soil and background samples were classified as sand (Fig. 6.1). Sand fraction
ranged 90.52 to 99.32 % in studied samples, and 92.84 to 94.20 % in background
samples (Table 6.1). An One-way Anova analysis showed significant differences
between studied soils and background samples in sand and clay fraction (p = 0.000).
The spatial distribution of sand, silt and clay fractions is presented in Figure 6. 2, being
the scale divided in classes of 20 % of each fraction range. Samples with higher
percentage of coarser particles are distributed in the center of the studied area,
corresponding to remobilized soils for construction purposes. Lower sand percentages
were found in samples collected in the dump limits. Dakheel et al. (2022) and Škrbić et
al. (2022) suggested that areas with nearby contamination sources, sandy soils are more
vulnerable to leaching and vertical migration of heavy metals to groundwater. Silt
fraction, ranged 0.44 to 7.69%, presented higher contents in samples collected near the
dump (Fig.6.1). Clay and silt fractions presented similar patterns, with higher clay
contents in samples on the dump proximity. This tendency can be associated to leaching
due to topographic context, with E-W leachates circulation (Bernardo et al., 2022).

Figure 6.1. Soil texture classification of studied soils (blue) and background (yellow)
samples.
109

Table 6.1 Granulometric statistical information (in %).

studied soils (n = 71) Background (n = 10)


fraction
min max mean±SD min max mean

sand 90.52 99.32 97.03±1.89 92.84 95.36 94.20±0.78

silt 0.44 7.69 2.112±1.41 2.27 3.36 2.74±0.34

clay 0.22 2.60 0.85±0.52 2.31 3.80 3.06±0.45

SD – Standard deviation

Figure 6.2. Granulometric spatial distribution of sand, silt and clay fractions of the
studied soils (in %; n = 71).
110

Soil pH applies to the concentration of ions (H+) present in the soil solution, being strongly
attracted to negative charges and have capability to replace other cations(Altaf et al.,
2021; Soubra et al., 2021). This parameter control soil adsorption and distribution of
heavy metals (Naveed et al., 2020). Sparling (2020) suggested that soil pH has a major
influence on heavy metals solubility. The pH of the studied 71 soil samples sand fraction
ranged 6.7 to 8.01, with a mean of 7.4; and in silt fraction varied between 7.0 and 8.4,
with mean 7.8 (Fig. 6.3). Spatially, soils pH in the two fractions, were classified as
moderately alkaline (7.9 - 8.4) to slightly alkaline (7.4 - 7.8). Moderately alkaline class
were found in sand fraction samples towards the west of the dump while silt fraction, in
samples in the immediate edge of the dump. Neutral pH (6.6 - 7.3), in silt fraction, was
found in samples across the western strip of the dump, while in sand fraction in the
northwest direction of the dump. Background samples presented acid pH, ranging 4.2 to
6.3, and mean 5.25 in sand fraction, and in 5.7 to 6.3, in silt fraction. Studies suggested
that heavy metals mobility tends to be lower with increasing pH, except for metal (loid)s
As, Mo and Se (Škrbić and Miljevic 2002; Zimik et al.,2021). Alkaline pH decreases Cu,
Co, Fe, Mn, and Zn bioavailability, because of low solubility in this pH range, while Mo,
Se, V, As and Cr are more available (Alexakis, 2021). In soils with pH > 6, occurs the
dissociation of H+ from OH groups in organic matter and Fe and Al oxides, increasing
metal adsorption and subsequent precipitation, reducing their bioavailability (Chen et al.,
2022; Choppala et al., 2018). However, Odom et al., (2021) found high levels of Zn
concentration in soils around the Dompoase landfill in Ghana under alkaline pH
conditions (8.1) and associated it with strong soil enrichment by leachates. In a similar
study, (El Fadili et al., 2022) reported that the soils in the surroundings of the landfill of
the city of Benguerir presented high levels of contamination by Zn and Cu in alkaline pH
and associated the process of adsorption of these metals that is remarkable in alkaline
pH. In tropical soils, such as those in the surroundings of the Hulene-B dump, heavy
metal retention can occur under high pH conditions (Campos, 2010), due to the
predominance of oxidic (Al, Fe and Mn) and kaolinitic mineralogy in the clay fraction,
which increases the metal adsorption capacity (Alleoni et al., 2005).
111

Figure 6.3. Soil samples pH in sand and silt fractions.

Soil samples EC ranged 43.1 to 725 µS/cm in sand fraction and 37.5 to 217 µS/cm in silt
fraction (Fig.6.4). All samples showed higher EC values when compared to average
background samples, of 17.7 μS/cm for sand fraction, and 13.9 μS/cm for silt fraction.
Sand fraction, with minimum EC 43.1 μS/cm was found on the northwest area and
maximum of 725 μS/cm in the immediate dump boundary to the northwest of the dump.
Spatially, the lowest values (43.1 – 100 μS/cm) were randomly distributed, in the center
and north of the sampling area and well evidenced in a strip parallel to the dump,
corresponding to a disturbed area during construction of the leachate drainage channel.
EC ranging 101 to 200 μS/cm, were found in samples collected near the dump, mainly
at the eastern end of the dump towards the center and north of the main sampling area.
Higher EC (201 – 725 μS/cm) were found in samples on the immediate western boundary
and with some dispersion to the southwest, west and in the northern direction of the
sampling area. Silty fraction, revealed minimum EC 37.5 μS/cm in samples at the
northern boundary, corresponding to a transition area between dump upper and lower
parts (characterized by intense leaching).It is also associated with the fact that this
sampling period was abundant rainfall, characterized by a constant dissolution of salts
in the soil and consequently a decrease in soil EC (USDA, 2011; Chaaou et al., 2022).

Maximum EC 217 μS/cm was found in the northern direction of the sampling center.
Spatially lower EC (37.5-100 μS/cm) was predominant throughout the sampling area,
followed by EC varying 101 – 200 μS/cm, predominant at the immediate western
boundary and sporadically on the south and north of the sampling area. In both fractions,
112

samples collected on the immediate western boundary exhibited higher EC, suggesting
an impact of waste discharges on surrounding soils. The eastern boundary of the Hulene
- B dump characterized by high conductivity was described by (Bernardoet al.,2022c)as
areas with shallow phreatic levels, which may be a factor in the capillary rise of salts and
water into the topsoil and consequently an increase in EC(USDA, 2011).Hussein et al.
(2021),found that soils with high EC values were associated with high soil contamination,
by As, Cd, Pb and Cr in a study of 6 dumpsites in Malaysia. Similar findings by Wu et al.
(2021), on a study of waste dumps in China, and Fatoba et al. (2021) on a waste dump
in Nigeria, with contamination by metal-ion enriched leachates.

Figure 6.4. Soil samples EC in sand and silt fractions.

Soil samples OM distribution is presented in Figure 6. 5. In studied samples, OM content


is scarce (mean 1.1 %), with few samples with >2 % in soils collected closer to the dump,
in the eastern and western sections, and in western area limit (> 2%, maximum of 4.2
%).The higher OM content, suggest an enrichment by contaminants resulting from the
migration of surface leachate from the dump and/or by the aeolian deposition of ash from
waste burning (Nisari et al., 2021). Lower OM content was found in the center of the
sampling area, corresponding to a depression, that might increase OM leachate and
surface water migration to the east-west direction, especially during rainy season.
Samples OM content, in the western end of the sampling, may be related to vegetation.
113

Figure 6.5. Soil samples OM content, color, and moisture content.

Soil samples color distribution is presented in Figure6. Blackish samples were found
predominantly in areas near the landfill site and the northwest. Brown samples were
predominant in the immediate eastern, western, and south-western boundaries of the
landfill. Greyish samples were more prevalent throughout the western sampling area.
Factors such as plants fixation, may contribute to darker colors in some samples distant
from the dump. Color has been associated to OM content, which exhibits an affinity for
heavy metal uptake (Dregulo and Bobylev, 2020; Sparling, 2020). The background
samples revealed reddish color in the TPv Formation and reddish brown in the QMa
114

Formation samples. Studies reported that the sandy soils may have color changes,
resulting from accumulation processes of OM transported by water, which subsequently
settle in the interstitial spaces, and may fixate some contaminants (Seidl et al.,
2021).Hussein et al. (2021)suggested that ash resulting from burning solid waste that is
transported and deposited changes soils color and enriches them with Hg. Color and OM
showed a similar trend, being good indicators for locating areas with possible
contamination (Dregulo and Bobylev, 2020).

Studied soil samples moisture content is presented in Figure 6. Background samples


mean content was 0.4%, ranging 0.2 to 0.6%. Samples from dump outskirts revealed an
average moisture content of 0.4%, ranging 0.0 to 1.9%. Samples collected in the
immediate edge of the dump showed very alternating moisture contents, with a
predominance of high and very high values (0.4 to 1.9 %). In the center of sampling area,
occurred a prevalence of moisture contents < 0.4%, alternating with higher 0.4 to 0.8%.
One sample in the immediate north-western boundary of the dump, showed the highest
concentration of 1.9%, being the lowest concentration of 0.0% in the southern boundary
and central sampling area. In general, sandy soils have low water retention capacity
(Zeitoun et al., 2021) and in the surroundings of dumps with surface drainage of leachate
whose substrate is sandy soils, the process of leaching and migration of leachate in
depth predominates (Zhang et al., 2021), being pointed as the cause of subsoil and
groundwater contamination in many studies (Stefania et al., 2019; Wu et al., 2021;
Przydatek et al., 2021).

6.1.1 Principal Component Analysis (PCA)

Principal component analysis reduces a set of variables into a smaller one called
principal components (PC), attempting to reveal variables correlation structure and
interpret parameters that influence soil samples. In this study, PCs were extracted from
71 samples (sand fraction) and 8 variables with Promax rotation. Three principal
components were considered, accounting for 73.14 % of the total variance. First
component (PC1), explained 40.86 % of total variance, defined two groups of variables:
sand negatively related to silt and clay variables (Fig. 6.6), with high negative loadings
(> 0.95). Second component (PC2) explained21.61 % of total variance, with positive
loading composed by OM, EC, and moisture. Third component (PC3), with 13.66 % of
variance, groups pH and color variables.
115

The opposition in PC1 between sand and smaller fractions (Silt and clays), is justified by
the high content of sand fractions > 90.52 % in the samples from the dumpsite
surroundings. Silty and clayey particle size was noticeable in few samples at the
immediate limits of the dump.

PC2 groups OM, EC, and moisture, influenced by incorporation effects of the metallic
ions contained in the leachates, as well as deposition of incineration ashes of
contaminated waste from the Hulene-B dump. The samples from the immediate limits of
the Hulene - B dump showed a spatial distribution that expresses the highest possible
correlation of these soil properties, characterized by high EC, OM, and Moisture.

PC3, on the one hand, suggested heterogeneous alteration conditions associated with
the movement and accumulation of leachates that successively cause heavy metal
contamination, and, on the other hand, the predominance of sandy granulometry that is
conducive to vertical leaching of contaminants.

Figure 6.6. PCA variables distributed by component.

Spearman's correlation (Table 6. 2), in line with PCA showed the same trend. The pairs
silt/sand, clay/sand, moisture/sand, pH/clay, color/pH (p < 0.01), color/sand, and
color/EC (p < 0.05) revealed a significant negative correlation. This suggests
heterogeneous processes that can be associated with the geo-environmental conditions
surrounding the dump, such as predominance of leaching in sandy soils which is
intensified by the circulation of leachate-enriched surface water, responsible for
significant EC increase and soil color alteration. Pairs clay/silt, EC/OM, pH/sand, pH/EC,
moisture/silt, moisture/clay, color/clay (p < 0.01) and moisture/OM (p < 0.05), revealed a
116

significant positive correlation. The impact of Hulene-B waste dump on soil properties,
by incorporation of metal ions from leachate and contaminated ash, alter soil properties
in a combined way at the immediate limits of the dump and in a localized manner at
various points in the surroundings of the dump

Table 6.2 Spearman correlation

sand silt clay OM EC pH moisture

silt -0.968** 1

clay -0.952** 0.868** 1


OM -0.126 0.151 0.135 1

EC 0.139 -0.038 -0.18 0.532** 1

pH 0.303** -0.202 -0.341** 0.149 0.527** 1

moisture -0.355** 0.377** 0.349** 0.282* 0.206 -0.11 1

color -0.279* 0.176 0.348** -0.148 -0.266* -0.367** 0.049


** p < 0.01; * p < 0.05

6.2 Landfill risk index assessment (Ip)


The analysis of the geo-environmental and structural context of the dump, allowed to
assess risk factor variables weights (Table 6.4). For the risk factor volume of deposited
waste, a maximum score of 1 and corresponding weight is 5 was assigned, taking in
consideration a deposition > 1000 t/day (Table 6.3). For the leachate drainage system,
a score of 1 and corresponding weight is 5 was assigned once leachate is dispersed on
soil in the surroundings of the dump (Fig.8). For the type of waste, a value of 0.5 was
assigned and a corresponding weight 3. For the stabilization typology was considered
0.3 and a weight of 3, due to dump successive accumulation of waste and a
heterogeneous process of waste biodegradation, and continuous accumulation causing
a reduction of oxygen in the layers below. The monitoring system was assigned a
maximum of 1, with a corresponding value of 2 since there is no monitoring mechanism
at the dump. The compaction of the waste is done by bulldozer, so a value of 0.2 was
assigned and a corresponding weight of 1. For the final coating, a value of 1 and a
corresponding weight of 1was given since the dump has no coating mechanism.
117

The factors with the highest scores were the volume of waste deposited (>1000 t/day)
and the conditions of its storage (open deposition, no isolation of the leachates from the
surrounding environment) and monitoring. These factors were described in many studies
as conducive to soil contamination in the surroundings of the dumpsites (Rapti-Caputo
et al., 2006; Lote, 2017; Liu et al., 2019; Rapti et al., 2021).

Table 6.3 Risk index (Ip) of the studied area

Risk factors Situation R W R*W

Landfilled waste volume > 1000 t/day 1 5 5


Drainage system Drainage absents 1 5 5
Type of waste Urban 0.5 3 1.5
State of the waste Solid 0.2 3 0.6
Waste biodegradability Anaerobic and Aerobic 0.5 2 1
Site monitoring Non-existent 1 2 2
Compacting the waste with bulldozer 0.2 1 0.2
Final coating material Absent 1 1 1

Ip 16.3
R – Risk factor; W - Weight parameter; R*W – Risk factor final weight.

Risk factors analysis allowed to classify Ip on the surroundings of the de Hulene-B dump
as very high (16.3), an indicator of possible groundwater contamination and need of
immediate intervention for impact mitigation (Rapti-Caputo et al., 2006). Structural
measures were adopted by local authorities along with the present study sampling
campaign, such as the construction of a surface leachate drainage system that could
have reduced the risk factor to 2.5 and the overall risk to 13.8. Nevertheless, the
constructed system does not systematically control leachate and doesn’t isolate the
surrounding environment (soils, groundwater, surface water and cultivated fields in the
vicinity of the dump) (Fig.6.7).
118

Figure 6.7. (a) Leachate drainage channels without isolation, (b)dispersed leachate, (c)
uncontrolled leachate flow.

Soils properties and landfill pollution risk assessment data combination suggested
complementarity to understand soil and groundwater contamination risk by leachate.
Results showed that soil properties were strongly impacted by the landfill site, as well as
spatial areas with greatest changes. The assessment of the landfill pollution risk was
relevant, as the factors associated to the management of the landfill site that influence
the change of soil properties around the Hulene-B dump were identified. The
combination of these results proved to be a knowledge base to be applied in the search
for measures to mitigate the impacts of the Hulene-B dump on soil properties and the
risk of groundwater contamination.
119

7 Paper 4 (Appendix 5)

Soil properties temporal variation and risk assessment in the surroundings


of Hulene-B dump, Maputo (Mozambique)

Abstract

The present study considers the temporal variation of soil properties in surroundings of
Hulene-B waste dump (Maputo, Mozambique), between a rainy season (2020) and a dry
season (2021). For that purpose, 71 superficial soil samples were collected in the two
climatic seasons. Soil properties analyzed were texture, pH, EC, MO, color, and
moisture. Dry season samples result suggested higher tendency for superficial
accumulation of potentially toxic elements. The surface water contamination risk index
(Pbci) was estimated, taking in consideration the characteristics of the dump and its
surroundings, revealing high values indicative of contamination by leachates. The study
highlighted the need for mitigation measures for soil and surface water contamination,
adjusted to the geoenvironmental dynamics of Hulene-B dump surroundings, taking in
consideration the fragilities of the environmental components and the main vectors of
contamination (leachates and waste incineration ashes).

Keywords: soil properties, waste dump, temporal variation; surface water, risk
assessment

Reference: Bernardo, B., Candeias, C., Rocha, F. (2022). Soils in the surroundings of
Hulene-B dump, Maputo (Mozambique) – properties temporal variation and risk
assessment. Submitted.
120

7.1 Results and Discussion

Textural distribution of the studied samples and background is presented in Figure 7.1.
Background samples and most of the soil samples collected in the surroundings of
Hulene-B were classified as sand, with the exception of samples P1II, P2II, P7II, P11II,
P4II, P21II, classified as loamy sand, due to their silt content. Samples with higher clay
content were ranked P1II > P67II > P2II > P11II > P21II > P66II > P55II, ranging 5.4 to
2.6 %. Sand content, dominant in all soil samples ranged 97.9 (P25II) to 81.9 (P1II) %
samples located at the W and E boundaries of the dump. Background soils sand content
varied 95.4 to 92.1 %, silt between 3.8 to 2.3 %, and clay between 3.8 and 2.3 % (Table
Fig.7.1). These data show significant differences with those published by Bernardo et al.
(2022c), on the study of soil texture in the rainy season, who classified the soils around
the Hulene-B dump as sandy. The presence of the silty and clayey fraction in most
samples in the dry season (4.48% and 1.56 % Table 7.1) and in the wet season (2.1%
and 0.22% Bernardo et al. 2022c) may be associated with the low predominance of
precipitation in the dry season, which is pointed out as a relevant factor in the leaching
of fine particles and ash from waste incineration that accumulate on the soil surface
(Dakheel et al., 2022). Clay and silt fraction is also pointed out in many studies as an
important factor for the adsorption of heavy metals in soils (Shaylinda, 2020). Textural
characteristics of the soils in the dry season show a higher metal adsorption capacity in
the soils surrounding the Hulene-B dump. In turn, the results of the background samples
showed no differences with the sandy and fine content classification referred to by
Bernardo et al. (2022c) 94.2% sandy, 2.7% silt and 3.1% in the rainy season and in the
dry season 93.84%, 3.09% and 3.07% respectively.

Figure 7.1 Soil texture classification of studied soils (blue) and background (yellow)
samples
121

Table 7.1 Granulometric statistical information (in %).

studied soils
fraction
min max mean SD
sand 81.92 97.86 93.94 3.84
silt 1.34 12.71 4.49 2.98
clay 0.59 5.37 1.56 0.93

In general, the background samples did not reveal a significant variation of the properties
like pH, EC, OM and moisture with those of the rainy season published by Bernardo et
al. (2022c), suggesting

reduced impact factors that could significantly modify the analyzed parameters (Table
7.2). However, samples collected in 2021 showed slightly higher pH, EC, and OM, than
those of 2020. These differences can be associated to the fact that 2021 samples were
collected in a period of scarce rainfall, which is a relevant factor in the variation of these
properties.

Table 7.2 Soil physical parameters of background (Bkg) 2020 and 2021 samples

2020 Bkg (n = 10) 2021 Bkg (n = 10)


var
min max mean min max mean
Sand 92.4 95.4 94.2 95.36 93.84 0.79
Silt 2.3 3.4 2.7 3.81 3.09 0.41
Clay 2.3 3.8 3.1 3.78 3.07 0.39
pH (< 2 mm) 4.2 6.3 5.3 5.2 6.2 5.6
pH (< 63 µm) 5.7 6.4 6.2 6.1 6.7 6.4
EC (< 2mm) 9.8 28.1 18.3 11.8 34.3 21.1
EC (< 63 µm) 10.8 16.5 13.9 11.0 19.0 13.5
OM 0.3 0.6 0.5 0.4 1.2 0.8
Moisture 0.2 0.6 0.4 0.2 2.0 0.7
var – variables; Bkg – background; min –
minimum; max – maximum; SD – standard
deviation; Sand, Silt, and Clay in %; EC –
electrical conductivity (in μS/cm); OM –
organic matter (in %); Moisture in %.
Hulene-B dump surrounding soils collected in 2020 showed, in sand fraction, a pH
ranging 6.7 to 8.01, with an average of 7.4, and in silt fraction ranging 7.0 to 8.4, with an
122

average of 7.8, being predominantly alkaline to slightly alkaline (Bernardo et al. 2022c).
The same parameter in 2021 showed that sandy fraction ranged 7.5 to 10.5 with mean
of 8.52, while and silt fraction varied between 7.6 and 8.9, and mean of 7.93. Silt fraction
revealed a predominantly slightly alkaline pH in most of the study area (Fig. 7.2 a and
b). Soil pH conditions in the surroundings of the Hulene-B dump in the two study periods,
classified as predominantly slightly to strongly alkaline were described in previous
studies as favorable to migration of potentially toxic elements (Rieuwerts et al., 2006).
However, in soils of tropical regions, heavy metal retention can occur under high pH
conditions (Campos, 2010), due to the predominance of oxidic (Al, Fe and Mn) and
kaolinitic mineralogy in the clay fraction, which increases the metal adsorption capacity
(Alleoni et al., 2005). Bernardo et al. (2022d) suggested a predominance of Al, Fe and
Mn oxides and a slight predominance of clay minerals in Hulene-B dump surrounding
soils, which can be an important factor in the adsorption and absorption of potentially
toxic elements even under mildly alkaline to moderate conditions. In a similar study
carried out near Mavallipura (India) by Naveen et al. (2017), alkalinity levels of
contaminated soils were detected and linked to successive accumulation of ash from
waste incineration and circulation of leachate from waste biodegradation, which are
potential factors in contaminating and altering soil pH. Škrbić et al. (2004) suggested that
a greatest heavy metal retention in soil occurs at high pH given the lower solubility metals
have in this pH range.

Figure 7.1. Soil samples pH variation (2020 vs. 2021) in (a) sand, and (b) silt fractions.

Soil EC of the 2020 sandy fraction samples showed ranged 43.1 to 725 µS/cm, with an
average of 220.59 µS/cm, and in silt fraction ranged 37.5 to 217 µS/cm, and an average
of 88.4 µS/cm (Bernardo et al., 2022c). Higher EC values were detected in samples
123

collected on the immediate limit of Hulene-B dump (Bernardo et al. 2022c). Samples
collected in 2021 showed higher EC ranging 90 to 9510 µS/cm, and average of 1150
µS/cm on sandy fraction, and ranging 85 to 1720 µS/cm and average 378 µS/cm on silt
fraction. Spatially, in 2021, sandy fraction presented higher EC values in the whole area
around the dump, with higher values between heterogeneously detected in some
samples collected in the north section of the sampling area, and in one sample in the
southwest of the dump. Lowest conductivities were detected in few samples located in
the north and southwest of the dump (Fig. 7.3a). Silt fraction higher EC was detected in
samples located on the W of the dump while lower EC was detected in one sample on
the north of the sampling area (Fig. 7.3b). Comparison of the two sampling periods
revealed that sand fraction presented higher mean EC than silt fraction. Previous studies
suggested that silt in general is more conductive than sand fraction (Nassereddine et al.
2013). However, organic matter in the sandy fraction is pointed out as a factor that raises
EC in sandy soils due to its capacity to adsorb contaminants (Nisari et al., 2021).

The 2021 EC results showed higher values than 2020 what is related to the different
climatic seasons of the sampling periods, the rainy season in 2020, and dry season in
2021. Studies suggested that precipitation is a factor in leaching and migration of
contaminants, influencing soils EC (Li et al., 2022). Results suggested a strong soil
contamination during the dry season (Siddiqi et al., 2022). Ashraf et al. (2022) compared
the contamination levels of soils in the surrounding of a landfill in Lahore, having similar
conclusions. Bernardo et al. (2022a) suggested in a comparative study between 2020
and 2021, using the electrical resistivity method, that in the dry period the resistivity
anomalies were located at a greater depth when compared to the rainy season and
associated this phenomenon to the poor superficial leaching in the dry period.

Other studies showed that during dry periods, characterized by scarce rainfall, there was
higher accumulation of heavy metals in the topsoil layer (Azeez et al., 2011; Chaudhary
et al., 2021). Wu et al. (2022) revealed that areas with contamination sources, e.g., waste
dumps, during dry period present higher transport and deposition of contaminants on the
soil surface. This suggested a predominance of higher EC in soils around the dump
during the dry season (Naveen et al., 2017).
124

Figure 7.1 (a) Soil samples EC variation (2020 vs. 2021) in (a) sand, and (b) silt
fractions.

Organic matter (OM) of 2020 samples ranged 0.0 to 4.2 %, and average of 1.09%
(Bernardo et al. 2022b). Samples collected in 2021, background samples ranged 0.4to
1.2 %, with an average of 0.72. Samples collected around the dump in 2021, ranged 1.8
to 8.6 %, and a mean of 1.87 %, higher OM content than 2020 samples (Fig. 7. 4a).
Spatially, OM content > 4 % and between 2 to 4 % were found in samples located in the
immediate eastern and western boundaries of the dumpsite, a similar pattern to 2020
study. Samples collected on the western boundary of the dump presented higher OM
content, occupying a larger area than 2020 samples (Bernardo et al. 2022 c). Lower OM
content were found in the central and northern areas of the sampling site, similar pattern
to 2020 samples (Bernardo et al. 2022 c). Temporally, 2020 samples showed a lower
OM content with a range of 0.0 to 4.2 % (Bernardo et al. 2022c) when compared to 2021
with a range of 1.8 to 8.6 %. Studies suggested that OM in sandy soils is susceptible of
leaching (Siddiqi et al., 2022), explaining the relatively higher predominance of OM
content in 2021 (scarce rain season). However, it is thought that high OM content in
topsoil is a factor that may lead to the uptake of contaminants, once heavy metals bind
easily to OM (USDA, 2001) and consequently decreases toxicity levels (Nisari et al.,
2021).
125

Figure 7.1 (a) Soil samples variation (202 vs. 2021) (a) OM content, (b) color, and (c)
moisture content.

Soil color, in 2020 samples, were characterized by blackish colors at the immediate limit
of the dump in a very narrow band, and slightly dark in the transition area between the
dump limit and the west center of the sampling area where lighter color were predominant
(Bernardo et al. 2022c). The 2021 collected samples color distribution was more
heterogeneous (Fig. 4b), with darker soil samples found at the immediate limit of the
dump, occupying a larger band than 2020 samples. Soil brownish color was found on a
wide band near the darkened samples and a band on the northwest. Greyish presented
a similar pattern to the 2020 samples (Bernardo et al. 2022c). Background samples
126

showed no color variations in both sampling periods. In general, blackish and brownish
colors are associated with higher levels of organic matter, indicative of a greater capacity
of the soil to absorb contaminants (Nisari et al.,2021). Greyish soil color, in areas with
contamination sources such as leachates have been associated to leaching processes,
which is consistent with the present study, since grayish soil samples were found in 2020
sampling, during dry season. In the surrounding of dumpsites, color is also influenced by
ash deposition resulting from solid waste incineration, which is a vector of contamination
mainly by Hg and Pb (Tao et al., 2020). In Hulene-B dump the practice of waste burning
is characteristic and may be associated with the blackish and brownish colors of the
predominant soils in the 2021 samples. Weibel et al. (2017) demonstrated that the
burning of waste contributed to the dispersion of Hg-contaminated ash, which was
subsequently deposited in the topsoil causing its enrichment and alteration of the soil
color.

Samples collected in 2020 in the surrounding of the landfill showed alternating soil
moisture content, ranging 0.0to 1.9 %, and average of 0.4% (Bernardo et al. 2022c). Soil
samples collected in 2021, showed higher moisture content (Fig. 4c) when compared to
those of 2020, ranging 0.18to 8.7 %, and average 0.94 %. Highest content was recorded
at the center to the west of the study area, and slightly elevated content at the immediate
edge of the dump. In the two study periods, background levels were slightly different,
with higher moisture content in 2021 samples. In the surrounding of Hulene-B dump
samples, in general, showed natural (low) moisture content. Low moisture content is
characteristic of sandy soils and is an indicator of strong leaching and migration of
contaminants, especially in soils surrounding dumpsites, that may be affected by the
leachates (Zeitoun et al., 2021). Möller et al. (2005), Hussein et al. (2021) suggested that
low soil moisture with active sources of contamination (urban areas, waste dumps) lead
to increased accumulation of contaminants in the topsoil. This can be evidenced by the
dry season results of soil properties in the surroundings of the Hulene-B dump. (Fig. 4c).

In general, soil properties data showed a certain alternation in the two studied periods.
The EC in the 2021 samples evidenced a much higher accumulation of contaminants on
the soil surface than in the rainy season, while pH, moisture, color, OM showed no
significant changes.

7.2 Principal Component Analysis (PCA)

Principal components were extracted from 71 samples (sand fraction) and 7 variables
(sand, silt, clay, OM, EC, pH, and moisture), with Promax rotation. Two principal
127

components were considered, accounting for 73.85 % of the total variance. First
component (PC1), explained 56.53 % of total variance, defined two groups of variables:
sand negatively related to silt, clay, OM, EC, pH, and moisture (Fig.7.5), with high
negative loading (> 0.972). Second component (PC2) explained 17.32 % of total
variance, with positive loading composed by pH (0.885).

The opposition in PC1 between silt sand, clay, OM, EC, pH, and moisture is justified by
the higher predominance of sandy soils in all samples (> 93.94%). However, the OM
contents were predominantly lower, being more prevalent in the samples close to the
dumpsite and associated to the incorporation of organic matter contained in the
leachates and resulting from the addition of ashes from the incineration of solid waste.
The EC was predominantly higher in all samples evidencing higher values than the world
average for sandy soils (< 100 µS/cm) and the pH was predominantly alkaline in all
samples.

The second component (PC2) with positive charge composed by pH, is justified by the
little alternation of pH which was predominantly alkaline in all samples. The high pH of
the soils in the surroundings of the waste dumps is used as an indicator of soil
contamination by heavy metals in many studies Škrbić et al. (2004)) and in particular in
soils of tropical zones with a higher predominance of Al and Fe oxides which are fixers
of heavy metals in soils (Campos, 2010).

Figure 7.2. Principal component analysis variables distributed by component.

Spearman correlation is presented in table 7.3. The pairs silt/clay, clay/sand, OM/sand,
EC/sand (p< 0.01), pH/sand, pH/clay, pH/OM, pH/EC, moisture/sand, and moisture/pH
(p< 0.05) revealed a significant negative correlation, suggesting that properties silt, clay,
sand, OM, EC and pH behave opposed. The opposition suggested different soil
128

alteration processes in the surrounding of Hulene-B dump. The opposition between clay
and sand is possibly linked to increased leaching and consequent PTEs migration.
Studies reported that pH is the most important factor in controlling mechanisms of heavy
metal contamination, determining their mobility in depth in the areas around dumpsites
(Naveen et al. 2017). Despite low silt, clay, and OM content, high EC was observed
which can be evidence of contamination by metal ions contained in leachates or solid
waste incineration ashes. The pairs clay/silt, OM/silt, OM/silt, EC/silt, EC/clay EC/OM,
moisture/silt, moisture/clay, moisture/OM (p <0.01) and moisture/EC (p < 0.05) showed
a possible correlation suggesting that the properties clay, silt, OM, moisture are
properties that change in the same trend, which can be evidenced by the samples that
had higher silt content showed higher clay and OM content thus evidencing a higher
adsorption and absorption capacity of heavy metals in the vicinity of Hulene-B dump,
especially in samples from the immediate limits of the dump which had high levels of
these contents. Samples relatively distant from the dump showed the same clay, silt, OM
trend, that may be associated with superficial and vertical leaching processes associated
with the free circulation of leachates and the contaminants migration to lower soil layers.
In similar studies, Zhang et al. (2010) and Pikula et al. (2022) suggested that reduced
clay, silt, and OM in sandy soils promoted the migration of contaminants in depth. OM,
EC, and moisture were properties with higher alterability in the vicinity of dumps, given
the circulation of leachate and ash from waste incineration that can increase these
properties (OM, EC, moisture), wtih moisture that can be associated with the
accumulation of surface water enriched by leachate to the west of the Hulene-B dump.

Table 7.3 Spearman correlation

sand silt clay OM EC pH


silt -0.995** 1
clay -0.954** 0.928** 1
OM -0.738** 0.746** 0.690** 1
EC -0.578** 0.601** 0.453** 0.637** 1
pH 0.225 -0.231 -0.201 -0.282 -0.048 1
moisture -0.650** 0.647** 0.614** 0.601** 0.315* -0.254
** p < 0.01; * p < 0.05
129

7.3 Risk of surface water contamination (Pbci)


The following weights (w) and ranking levels (c) were assigned to the studied variables,
based on the characteristics of the landfill and its surroundings (Calvo et al., 2005): (i)
compaction, given the precariousness of the compaction processes at the Hulene-B
dump, characterized by being partial, it was assigned a w = 2 and c = 3; (ii) cover
material, once is an open-air landfill without any mechanism to cover the waste, it was
assigned w = 2 and c = 4; (iii) leachate control, at the northern, southern and western
borders of the landfill, to where natural leachate drains and been opened and cyclically
overflowing, with no underground isolation being likely that contributes to migration of
leachate in depth, therefore a maximum w = 2 to surface water and w = 1 to soil, has
and c = 4; (iv) age and percentage of organic matter, the landfill dates from 1973 (~50
years age), this assumes little leachate generation but with >1000 tons of miscellaneous
waste deposited daily, and an estimated 50 % of organic, therefore w = 2 and c = 4; (v)
impermeability of discharge point, the landfill has neither been waterproofed nor has its
surroundings, increasing the potential risk of soil, ground and surface water
contamination, therefore a w = 1 for soils, and w = 2 for surface water, and a maximum
c = 4 has; (vi) point located in flood-prone areas, the surroundings of the dump are
susceptible of inundation, given the inexistence of mechanisms to adequately control
surface flows leading to widespread flooding by surface leachates from the swift, w = 4
and c = 2 were assigned; (vii) rainfall, average annual precipitation of ~789 mm, and
once higher precipitation influences decomposition and leachate production, w = 2 and
c = 2 were assigned; (viii) existence and distance of surface water, west of the dump
surface water occurs occupying the surface of the dune depression every year and other
cyclically floods at more distant areas, w = 2 and c = 4.

Contamination risk of surface water was 0.91 (Table 7.4), being almost all factors
contributing with a higher value (8), except for the compaction factor (6), given the
process of waste compaction by bulldozer that takes place in recent years.
130

Table 7.4 Contamination risk of surface water around Hulene-B waste dump

variables W*C
Compaction 6
Daily Cover 8
Control of leachate 8
Final coverage 8
Type of waste and % organic matter 8
Waterproofing the discharge point 8
Rainfall 4
Existence and distance of surface water 8
Contamination risk High 0.91

High risk of surface water contamination, associated with Hulene-B dump characteristics
and contamination processes occurring by leachate free circulation at various points of
the W boundary (Fig. 7.6c), associated with ashes deposition of incinerated waste, are
vectors of contamination and alteration of soil properties. Bernardo et al. (2022b),
suggested a high risk of soil and groundwater contamination mainly by leachate in the
surrounding area of Hulene-B waste dump. Naveen et al. (2018), showed high levels of
surface water and soil contamination in the surrounding a landfill in Bangalore (India),
and associated it to the contaminants contained in leachates.

Figure 7.3. (a) Leachate at the W boundary of Hulene-B; (b) leachate at the S
boundary of Hulene-B; and c) surface water.
131

8 Paper 5 (Appendix 6)

Soil Risk Assessment in the Surrounding Area of Hulene-B Waste Dump,


Maputo (Mozambique)

Abstract

Soil contamination in areas close to unplanned dumpsites represents an increasing risk


to the ecosystems and human health. This study aimed to evaluate soil quality in the
area surrounding the Hulene-B waste dump, Maputo, Mozambique, and to estimate
potential ecological and human health risks. A total of 71 surface soil samples were
collected in the surrounding area of the dump, along with 10 samples in areas considered
not impacted by the dump. Chemical and mineralogical analyses were performed using
XRF and XRD. Quartz was the most abundant mineral phase, followed by feldspars,
carbonates, clay minerals, and Fe oxides/hydroxides. Results showed a significant
contribution to ecological degradation by PTE enrichment, ranked as Zn >> Cu > Cr > Zr
> Pb > Ni > Mn. Carcinogenic risk for both children and adults was significant due to Pb
soil content. Soil sample concentrations of Cr, Cu, Mn, Ni, Pb, Zn, and Zr, posing a risk
especially in children, suggested the need for continuous monitoring, as well as the
definition and implementation of mitigation measures.

Keywords: soils; waste dump; potential toxic elements; risk assessment

Reference: Bernardo, B.; Candeias, C.; Rocha, F. Soil Risk Assessment in the
Surrounding Area of Hulene-B Waste Dump, Maputo (Mozambique). Geosciences 2022,
12, 290. https://doi.org/10.3390/geosciences12080290
132

8.1 Results and Discussion


8.1.1 Mineralogical and Chemical Characterization

Mineralogical analysis showed that quartz (SiO2) was the dominant mineral, ranging
93.1% to 99.6%, except in samples P40 (87.0%) and P60 (68.4%). Other mineral phases
present were ranked as follows: feldspars (K feldspar (KAlSi3O8) and plagioclase
((Na,Ca)[(Si,Al)AlSi2]O8); 0.2% to 6.2%) > carbonates (calcite (CaCO3), dolomite
(CaMg(CO3)2), and siderite (FeCO3); 0.2% to 5.7%) > kaolinite (Al2(Si2O5)(OH)4; 0.1% to
1.4%) > alunite (KAl3(SO4)2(OH)6; 0.1% to 1.4%) > anhydrite (CaSO4; 0.1% to 1.1%)
(Figure 8. 1). Other minerals present in a few samples were magnetite–maghemite
(Fe3O4-γ-Fe2O3), zeolites, opal C/CT (SiO2·nH2O), vaterite (CaCO3), and melanterite
(Fe2+(H2O)6SO4·H2O). Local background samples revealed a similar trend but showing
some ilmenite (Fe2+TiO3; 0.2% to 0.3%). These results are similar to those described by
Momade et al., (1996) who considered the soils surrounding Hulene-B as heterogeneous
as a result of remobilization and mixing of the Ponta Vermelha (90% to 95% quartz) and
Malhazine formations (~98% quartz).
133

Figure 8.1 Mineral phases identified on the studied soils and local background
samples. Quartz was not considered.

The chemical composition of the areas surrounding the waste dump and local
background samples is presented in Table 8.1. Cd was below the limit of detection (LOD
3.88 mg/kg) in 93% of the samples. A one-way ANOVA revealed significant differences
between soil and local background sample concentrations of Al, Fe, K, Mn, Ni, Rb, Si, V,
and Zr (p < 0.05). Higher concentrations were found in local background samples, except
for Si and Zr, with lower content. The high concentrations of Al, Fe, K, Mn, Ni, Rb, and V
in the local background samples could be associated with the mineralogical and
geochemical composition of Ponta Vermelha formation, characterized by high levels of
Al and Fe oxides, continually remobilized to the Malhazine formation (Momade et al.,
1996). Additionally, the Malhazine formation in local background areas does not exhibit
high erosional disturbance or surface runoff, as it is relatively flat. Soils in the surrounding
of dump are distributed between a dune slope and a depression (heterogeneous area of
sediments from the Malhazine and Ponta Vermelha formations), characterized by strong
134

sediment mobilization and cyclic retention of surface water from the dump and the
unchanneled flow of Julius Nyerere Avenue in the east. Studies by Ghasera et al. (2022);
Han et al. (2022) suggested that Al, Fe, K, Mn, Ni, Rb, and V concentrations in soils with
strong surface water circulation and high porosity are prone to strong depletion due to
high oxide mobility. The relatively high concentrations of Si in soil samples could be
associated with the soil mineralogical composition, with predominance of silicate
minerals. The feldspar presence was similar to the results in (Momade et al., 1996).
Studies by Kabata-Pendias (2000) suggested that Fe and Al oxides have the ability to
adsorb and retain PTEs. The authors of Bai et al., (2021) suggested that SiO2 mineral
also has the aptitude to adsorb PTEs, such as Pb and Cu, while clay minerals, such as
kaolinite, can adsorb and retain PTEs in soils (Srivastava et al., 2005).
135

Table 8.1 Descriptive statistics of studied soil and local background samples (in mg/kg)

Soils (n = 71) Local Background (n = 10)


Var
Min Max Mean SD Min Max Mean SD

Al 6399 24,510 11,453 3974 19,101 32,200 22,406 4061

Ba 64 287 117 41 72 134 106 23

Br 0.5 12.1 3.7 2.4 1.0 3.5 2.3 0.8

Ca 486 31,240 9167 8630 243 843 471 217

Cr 10.3 238.0 41.3 41.1 20.0 59.4 33.5 11.5

Cu 3.3 1470 59.7 189 6.8 14.0 9.2 2.1

Fe 1364 27,839 4655 4204 5316 9204 6801 1255

K 3512 10,601 6278 1719 6409 14,129 9001 2263

Mg 10 3 028 979 530 609 965 675 104

Mn 23 310 83 57 95 194 140 34

Na 386 3769 1091 780 519 1803 974 343

Ni 0.5 16.1 3.4 3.3 2.8 5.2 3.9 0.7

P 79 3 029 686 677 127 284 197 53

Pb 6.4 506 30.2 59.6 6.6 9.5 7.8 0.8

Rb 8.0 28.8 13.1 3.4 13.9 23.8 18.0 3.8

Si 363,290 451,075 430,505 18,882 398,497 433,284 422,573 13,835

Sn 0.5 110 7.0 12.9 0.5 4.9 3.4 1.6

S 50 3228 544 578 55 175 103 31

Sr 9 89 29.5 21.8 10.0 18.0 13.6 2.7

Ti 779 2812 1624 514 1121 3333 1754 696

V 3 18.4 7.3 3.3 9.6 19.9 13.6 3.3

Zn 2 1077 92 162 3 4 3 1

Zr 65 341 161 65 81 200 123 40

Var—variable; Bkg—local background; min—minimum; max—maximum; SD—standard


deviation.

Principal component analysis (PCA) performed on the studied soil samples identified
four main components: (a) Group 1, with 61.2% of the total variance explained, grouping
Br, Ca, Mg, Mn, Ni, P, S, and Sr with positive loading and Si with negative loading; (b)
Group 2, explaining 12.4% of the total variance, with variables Al, Ba, Cr, K, Na, Rb, and
136

V; Group 3, explaining 8.0% of the total variance, with elements Cu, Fe, Pb, Sn, and Zn;
(d) Group 4, with 4.4% of the total variance grouping variables Ti and Zr.

Group 1 suggested the existence of two enrichment sources, the anthropogenic Hulene-
B dump with elements Br, Ca, Mg, Mn, Ni, P, S, and Sr and a natural source of Si, with
negative loading. The waste dump elemental contribution can be evidenced by higher
mean concentrations in studied samples when compared to local background mean
contents (Table 8. 2). The high Si concentration may be associated with mineralogical
phases identified, with >98% quartz (SiO2) (Momade et al., 1996).

Group 2 variables suggested heterogeneous sources. High Cr concentrations might be


related to anthropogenic enrichment (e.g., dump and Julius Nyerere Avenue degraded
asphalt). Group 3 variables presented high concentrations of Cu, Pb, Sn, and Zn when
compared to local background samples, suggesting anthropogenic sources (e.g., dump,
Julius Nyerere Avenue degraded asphalt, and airport traffic flow). The low Fe content
can be linked to leaching of Fe oxides given the topographic nature of the sampling area
(Bernardo et al., 2022 b) promoting water circulation and infiltration. The low Ti
concentration (Group 4) could be associated with geogenic context and the Ti oxide
leaching processes. The higher Zr concentration (Group 4), when compared to local
background samples, suggested an anthropogenic source (dump).

To access elemental sources, a Spearman correlation analysis (Table 8.2) was applied
to elements Cr, Cu, Mn, Ni, Pb, and Zn, considered by several studies as PTEs (See
appendix 1 for a detailed description of each element – page 194). The analysis revealed
statistically significant correlations between pairs Cr/Cu, Mn/Cr, Mn/Cu, Ni/Cr, Ni/Cu,
Ni/Mn, Pb/Cr, Pb/Cu, Pb/Mn, Pb/Ni, Zn/Cr, Zn/Cu, Zn/Mn, Zn/Ni, Zn/Pb, and Zr/Cr (p <
0.01), suggesting a common source.
137

Table 8.2 Spearman correlation between the selected PTEs of the soil samples.

Cr Cu Mn Ni Pb Zn

Cu 0.640 ** 1

Mn 0.747 ** 0.709 ** 1

Ni 0.770 ** 0.699 ** 0.792 ** 1

Pb 0.626 ** 0.759 ** 0.669 ** 0.711 ** 1

Zn 0.695 ** 0.813 ** 0.791 ** 0.799 ** 0.734 ** 1

Zr 0.377 ** 0.14 0.218 0.152 0.175 0.155


** p < 0.01.

Cluster analysis was conducted on selected PTEs in order to assess Cr, Ni, Pb, Mn, Zn,
Cu, and Zr common sources. Two main variable groups were formed (Figure 8. 2).
Cluster 1 grouped variables Cr, Ni, Pb, Mn, Zr, and Zn, suggesting a common
anthropogenic source (Hulene-B waste dump), with elemental concentrations higher
than local background samples (p < 0.05). Contaminants may be associated with the
dispersion of leachates from the dump, ash from burnt waste, degraded pavements and
traffic-related materials from Julius Nyerere Avenue, and the international airport (west
of the dump). Cr pollution has been associated with paints, varnishes, organic solvents,
oils, and vehicle brakes (Fan et al., 2022; Al-Salem et al., 2021) where as Ni pollution
has been associated electronic wastes deposited on the dump (Fiala et al., 2021). Zr
and Mn contamination could be associated with paints, cosmetics, pharmaceuticals,
textiles, electrical waste, traffic associated emissions, and electronic equipment (Islamd
et al., 2021) while Pb could be associated with electronic wastes such as rechargeable
batteries, paint cans, varnishes, organic solvents, glass waste, fuel combustion, and air
transport (Alghamdi et al., 2021). Cluster 2 grouped elements Cu and Zn, characterized
by higher concentrations than local background samples, suggesting a common origin
(the dump site). These PTEs were linked to electronic waste, paints, bottle caps, tire
wear, and asphalt degradation. Studies by (Gujre et al., 2021) found high concentrations
of heavy metals in soils around landfills and considered an association with leachates.
Furthermore, the authors of (Obiri-Nyarko et al., 2021; Chen et al., 2022) suggested that
high levels of Pb, Zn, Hg, and Cd in soils around a dumpsite can be associated with ash
deposition in soils, resulting from contaminated waste burning.
138

Figure 8.2. Hierarchical cluster analysis of the selected PTEs.

8.2 Environmental Risk Assessment


The PTEs Cr, Cu, Mn, Ni, Pb, Zn, and Zr were selected to estimate the impact on the
environment using the potential ecological risk index (PERI), pollution load index (PLI)
and soil Nemerow index (PN) (Figure 8. 3). PLI results showed that 15.5% and 2.8% of
all samples were classified with high and very high contamination, respectively. A greater
representative number of samples (56.3%) were classified with moderate contamination,
while 25.4% were classified with low contamination. The PLI individual element
contribution was ranked as Zn >> Cu > Pb > Cr = Zr > Ni > Mn. Spatially, the
environmental contamination levels varied significantly. PLI results < 1 were found in
samples located at the central, western, and northern boundaries of the dump. In turn,
samples suggesting high levels of environmental deterioration (PLI > 1), due to Zn >>
Cu > Pb > Cr concentrations, were located at the eastern boundary and at the immediate
western boundary of the dump. Two samples with extremely high levels of deterioration
(PLI > 6) were located at the southwestern (densely inhabited) boundary and at the
immediate northwestern boundary of the dump.
139

Figure 8.3. Potential ecological risk index (PERI), pollution load index (PLI), and soil
Nemerow index calculated with local background soils and with world soils proposed by
Reimann and Caritat (1998).

PERI results revealed that 59.2% of all samples presented a low ecological risk, while
22.5%, 11.3%, and 7.0% of the samples were classified with moderate, serious, and
severe potential ecological risk, respectively. The monomial potential ecological risk
factor was ranked as Zn >> Cu > Cr > Zr > Pb > Ni > Mn. The samples with PERI index
<150 were concentrated in the center and north of the sampling area. Medium and high
risk was found to the immediate east of the dumpsite and to the east of the sampling
area. Extremely high risk (PERI > 600) was predominant in samples to the east, west,
and southwest of the dump (densely inhabited).
140

The soil Nemerow index was calculated using both local background (PNbkg) and world
soils (PNRC) proposed by (Reimann et al., 1998). The spatial distribution of (PNRC)
showed that only four samples, on the eastern boundary of sampling area, were
classified as nonpolluted. The warning level of pollution was found in random samples in
the center, east, west, and north of study area. Samples with low to moderate levels of
pollution were predominant at the eastern and northern boundaries, while a high level of
pollution was identified in samples at the immediate east and west boundaries of the
dump and in the extreme southwest (densely inhabited). The PNbkg suggested that only
three samples were classified with a low pollution level, whereas four samples were
classified with a moderate pollution level, located at the eastern boundary and in the
northern direction of the dump.

The eastern and western boundaries of the dump were characterized by a high pollution
level, suggesting a strong contribution of the dump to soil contamination. Pollutants were
ranked as Zn >> Cu > Pb > Cr = Zr > Ni > Mn for PNbkg and Cu > Pb > Zn > Cr > Zr > Mn
= Ni for PNRC. About 90.1% of studied samples were classified as seriously polluted, with
no samples in the safety or precaution domains. PNRC revealed 46.5% of the samples to
be slightly polluted and 23.9% to be seriously polluted. The variation between local
background and world soil reference indices evidenced a bias resulting from the differing
biogeochemical characteristics of the soils (Reimann et al., 1998).

Dumps have been reported to be a source of enrichment of pollutants in surrounding


soils (El Fadili et al., 2022). Variables Cr, Cu, Mn, Ni, Pb, Zn, and Zr found in the studied
samples could be associated with different anthropogenic sources, such as the dump
wastes, surrounding traffic and pavement degradation Candeias et al. (2022),
resuspension Brtnický et al. (2020), and airport proximity (Perks et al., 2021). Aluminum
has been associated with untreated disposal of cans, household utensils, cosmetics, and
laminated packaging (Morita et al., 2021). Cr has been associated with packaging waste
from paints, varnishes, organic solvents, and oils (Fan et al., 2022 ; Al-Salem et al.,
2021) Cu, Pb, Zn, and Ni have been associated with waste electronic equipment,
rechargeable, paint cans, varnishes, organic solvents, and glass waste (Safonov et al.,
2022; Fiala et al., 2021) and Zr has been associated with hospital wastes, explosives,
and lamps (Perks et al., 2021) all deposited in the Hulene-B dump.

The spatial distribution of indices (Figure 8.4) suggested very high levels in the
immediate eastern, western, and southwestern limits of the dump, with medium to low
141

levels predominant in the central and northern areas. The high contamination levels
found in the southwest area of the dump may be associated with three contamination
mechanisms, namely, incineration of wastes ash deposition, a recent sporadic waste
disposal area that contributes significantly to leachates, especially during rainy season,
with mobile metals such as Mn, Zn, Ni, and Zr (Aydi et al., 2021), and the Maputo
International Airport in the vicinity of the study area (Figure 1). Contamination found on
samples collected west of the study area may be associated with leachate circulation
originating from the dump, an important source of enrichment of mainly Zn, Cu, Pb, and
Cr, which are easily incorporated into the leachate (Zaynab et al., 2022). High levels of
contamination on the eastern border may be associated with leachate dispersion and
intense traffic flow on Julius Nyerere Avenue (Figure 1), due to degraded asphalt, a
source of Pb, Cu, Mn, Ni, Zn, and Zr. Recently, the authors of Bernardo et al. (2022)
suggested that the area around the Hulene-B dump is characterized by anomalous
resistive surfaces associated with heavy-metal contamination, e.g., Cr, Cu, Mn, Ni, Pb,
Zn, and Zr, as well as pollutants reported in leachates originating from landfills (Aghili et
al., 2018).

8.3 Human Health Risk Assessment


Areas with contaminated soils pose a health risk to the exposed population
Zaynab et al. (2022); Ceballos et al. (2021). For example, Cr chronic exposure
can cause nasal irritation, bleeding, ulcers, convulsions, kidney, and liver damage
and even death (Aghili et al., 2018). The mean Cr concentration (41.3 mg/kg)
found in the studied soils was below maximum allowable soil (MAS) target
proposed by (EU, 2004) (150 mg/kg). Nevertheless, a maximum of 238 mg/kg was
found in a sample located in the eastern end of the waste dump, near Avenida Julius
Nyerere. The high concentration levels may be associated with traffic-related waste,
such as tires and brakes, as well as emissions from Julius Nyerere Avenue traffic,
including degraded road pavement. The toxic effects of Cu chronic exposure include
anemia and disorders of the central nervous system and cardiovascular system. The
average Cu concentration found in this study was 59.7 mg/kg, below the MAS (150
mg/kg) (Ceballos et al., 2021) and a maximum of 1470 mg/kg was found in a sample at
the immediate northwestern boundary the dump. This area is characterized by intense
tire burning. Tires are considered a source of Cu enrichment. Zinc toxicity leads to loss
of appetite, dehydration, weakness, weight loss or gain, diarrhea, and jaundice (Stuckey
et al., 2021). The mean content found in the studied samples was 92 mg/kg, below the
MAS (250 mg/kg) (Ceballos et al., 2021); however, a maximum of 1077 mg/kg was found
142

in a sample at the immediate northwestern boundary the dump. Zinc enrichment may be
associated with batteries, metal alloys, paint cans, cosmetics, pharmaceuticals, textiles,
and electrical and electronic equipment, accumulated over time (Fiala et al., 2021).
Nickel chronic exposure has been associated with gastrointestinal and neurological
effects including lung cancer (Stuckey et al., 2021). The average concentration in studied
samples was 3.4 mg/kg, below the MAS (45 mg/kg) (Ceballos et al., 2021). Ingestion or
inhalation of Pb can disturb almost all organs and the nervous system, in addition to
causing kidney damage, brain damage, miscarriage, and death (Naveed et al., 2020).
The average concentration found in the studied samples was 30.2 mg/kg, below the MAS
(300 mg/kg) EU, (2004), with a maximum of 506 mg/kg in a sample at the immediate
northwestern boundary the dump. Mn is classified with low toxicity (Sattar et al., 2021);
however, chronic exposure to high levels can cause memory problems, hallucinations,
Parkinson’s disease, pulmonary embolism and bronchitis, apathy, schizophrenia, muscle
weakness, headaches, and insomnia (Li et al., 2022). The average concentration found
was 83 mg/kg, below the MAS (330 mg/kg) (Ceballos et al., 2022), and a maximum of
310 mg/kg was observed in a sample at the immediate northwestern boundary the dump.
Zr is classified with low toxicity, but chronic exposure can cause respiratory tract irritation,
dermatitis, and pulmonary fibrosis, with a few cases reported in nonindustrial settings
(Stewart et al., 2022). The limits of its concentration in soils are still in debate (Stewart
et al., 2022).

Considering that the surroundings of the Hulene-B dumpsite are densely populated (INE,
2017), and that the dumpsite is used daily by adult and children waste-pickers (Matsinhe
et al., 2020), the noncarcinogenic hazard, systemic toxicity, and carcinogenic risk
associated with selected PTE concentrations were calculated. The hazard index (HI) for
systemic toxicity (Figure 5), as expected, was higher in children than adults, due to the
lower body mass and higher absorption factor in children (Yap et al., 2022; Jones et al.,
208). Elements that most contribute to HI are Zr and Al. The higher exposure of children
to PTEs present in soils around the Hulene-B dump was recently reported by Matsinhe
et al. (2020), who found that children around the dump were collecting solid wastes
without using proper protection, thus increasing exposure.
143

Figure 8.4. Hazard index for systemic toxicity for children and adults, as well as
adjusted for both children and adults.

Carcinogenic risk, for both children and adults, only posed a significant risk for Pb
ingestion in samples P13 (6.19 × 10−6) and P33 (2.94 × 10−6), remaining lower than 1.00
× 10−4, the target level. Sample P13 was collected at the immediate northwestern
boundary of the dump, ~20 m away from dwellings, where waste pickers are common,
and sample P33 was collected in a residential area, southwest of the dump, representing
a risk especially for children, due to hand-to-mouth habits, associated with the fact that
Pb accumulates at the soil surface in the first 2–5 cm (Islamd et al., 2020). Studies by
Lian et al. (2022); USEPA, (2017) described that Pb is very toxic to children and pregnant
women, who are more susceptible to adverse health effects. The population around the
Hulene-B dump is young and active (14–44 years), representing ~65% of the inhabitants
INE, (2020), who spend most of their time working on solid waste recycling inside the
dump, often without adequate protection (Matsinhe et al., 2020). In both samples, total
risk and ingestion risk were similar, with inhalation and dermal contact considered
negligible.
144
145

9 Paper 6 (Appendix 7)

The contribution of Hulene-B waste dump (Maputo, Mozambique) for the


contamination of rhizosphere soil, edible plant, and stream and
groundwater

Abstract

The contamination of ecosystems in areas around waste dumps is a major threat to the
health of surrounding populations. The aim of this study is to understand the contribution
of the Hulene-B waste dump (Maputo, Mozambique) to the contamination of edible
plants, rhizosphere soils, stream waters, and groundwater, and to assess human health
risk. Soil and plant samples were analyzed by XRD and XRF for mineralogical and
chemical composition characterization, respectively. Mineral phases identified in
rhizosphere soil samples were ranked, calcite (CaCO3) > quartz (SiO2 ) > phyllosilicates
(micas and kaolinite) > anhydrite (CaSO4 ) > K feldspar (KAlSi3O8 ) > opal (SiO2 ·nH2O)
> gypsum (CaSO42H2O), suggesting potential toxic elements low mobility. Soil
environmental indices showed pollution by Pb > Cu > Zn > Zr. The chemical composition
of edible plants revealed contamination by Ni, Cr, Mn, Fe, Ti, and Zr. Groundwaters and
stream waters showed a potential health risk by Hg and, in one irrigation water sample,
by Pb content. The health hazard index of rhizosphere soils was higher by ingestion, with
children being the ones more exposed. Results suggested a combined health risk by
exposure to edible plants, rhizosphere soils, stream waters, and groundwaters.

Keywords: soil rhizosphere; edible plants; surface and groundwater; contamination; risk
assessment

Reference: Bernardo, B.; Candeias, C.; Rocha, F. (2023d). The Contribution of the
Hulene-B Waste Dump (Maputo, Mozambique) to the Contamination of Rhizosphere
Soils, Edible Plants, Stream Waters, and Groundwaters. Environments, 10, 45.
https://doi.org/10.3390/environments10030045
146

9.1 Results and discussion


9.1.1 Risk assessment of rhizosphere soils

Rhizosphere soils pH, electrical conductivity (EC), organic matter (OM), and color are
presented in Table 9.1. Soil pH is an important property once controls metals chemical
reactions (Lee et al., 2022). Rhizosphere soils revealed a moderately alkaline pH (8.1 -
8.4), confirming Naveen et al. (2017) findings that, in general, areas impacted by heavy
metals from dumps tend to have high pH. Gajaje et al. (2021) study, in soils around
Morupule dumpsite (Botswana), found an average pH of 8.3 in rhizosphere soils and
associated it with heavy metal contamination (Cu, Cr, Mn, Pb). Skrbic et al. (2002)
suggested that the PTEs highest retention in rhizospheres occurred with high pH, due to
lower metals solubility in this pH range. Alkaline pH can affect plant growth because of
the difficulty in absorbing essential plant nutrients, as well as, e.g., Fe, Mn, Cu and Zn
(USDA, 1998).

Table 9.1 Rhizosphere Proprieties

ID pH EC OM Color

R1 8.1 510 3.6 blackyish

R2 8.4 571 1.7 greyish

R3 8.2 310 0.98 greyish

R4 8.4 810 2.3 greyish


EC - electrical conductivity, in μS/cm; OM –
organic matter, in %.

EC showed high values, when compared to sandy soils guidelines (≤ 100 μS/cm; Lund,
2008). In this study, rhizosphere soils were ranked R4 > > R2 > R1 > R3. Bhardwaj et
al. (2020) suggested that rhizosphere soils in areas near waste dumps, usually exhibit
higher EC values, being associated to contamination by metal ion enriched leachates
(Hussein et al., 2021), and deposition of contaminated ashes resulting from wastes
incineration (Li et al., 2022). Recently Bernardo et al. (2022) reported the predominance
of high EC in Hulene-B dump surrounding soils and associated it with leachates.
Chemical analyses of surface soils showed high Cr, Cu, Mn, Ni, Pb, Zn, and Zr
concentration (Bernardo et al., 2022c). Another factor that may be associated with high
EC is the presence of shallow water table (USDA, 2011), thought to increase EC
(Chaaou et al., 2022). In sandy soils, the presence of water retained during rainfall, and
periodic enrichment by surface leachates can promote (Bernardo, 2019), retaining
147

rainwater, and periodically enriched by surface leachates from Hulene-B dump


(Bernardo et al., 2022a). Studies related EC to crop yields and found that high EC values
can reduce yields (Cheng et al., 2021). However, for horticultural crops, EC (salinity)
results were considered within the recommended limit (< 1000 μS/cm) (USDA, 2011).

Soil OM content was ranked R1 > R4 > R2 > R3, coherent with other studies sandy soils,
in general with low OM (Huang et al., 2020). The slight differences in OM content may
be due to different uptake capacities of PTEs OM (Cheela et al., 2021). Some studies
have shown that in contaminated areas, high OM concentration in rhizosphere, can lead
to heavy metals uptake, that bind easily with OM (USDA, 2001) and consequently
decrease their toxicity levels from the rhizosphere to crops (Wen et al. 2018; Nisari et
al., 2021).

Rhizosphere soils color was classified as very dark in all samples, with sample R1 being
the darker, such as its OM content. Soil color is influenced, in general, by OM content,
which can be incorporated in the interstitial spaces of sandy soils (Seidl et al., 2021).
Momade et al. (1996) suggested that soils surrounding the Hulene-B dump were
predominantly sandy with greyish coloration. Bernardo et al. (2022d) linked this soils
color slight variation to factors such as leachates circulating freely on the soil surface and
waste incineration ashes, associated with heavy metal contamination of rhizosphere
soils (Alexakis, 2021), that transferred to edible plants (Alexakis, 2020).

Mineral phases identified on studied rhizosphere soils, where quartz (SiO2), where
(calcite (CaCO3), phyllosilicates (micas and kaolinite), anhydrite (CaSO4), K feldspar
(KalSi3O8), opal (SiO2·nH2O), and gypsum (CaSO4·2H2O) (Fig. 9.2). results suggest
differentiated results between soils in the intra dune area and in other strips of the area
surrounding the Hulene-B dump predominantly composed by quartz > 90% (Bernardo et
al., 2022b). Quartz content (18.1 to 40.5 %) is similar to the mineralogy of intra dune
quaternary deposits (Momade et al., 1996). Samples R2 and R4 showed a higher
percentage of calcite because they are in influence of soils remobilized at depth during
the construction of the leachate transport channels. Samples R2 and R4 showed a higher
percentage of calcite because they are in influence of soils remobilized at depth during
the construction of the leachate transport channels.

Phyllosilicates (3.0 to 28.8 %) may be linked with the predominant aeolian and superficial
deposition processes of the intra dune area (Momade et al., 1996). Previous studies
suggested that soils with significant phyllosilicates content have a higher PTEs
adsorption capacity (Asowata, 2021). Opal, under conditions of exceptional morphology
148

fracture, can adsorb PTEs such as, Pb, Cu, Cr (Bai et al., 2021). Calcite has been
reported to have the ability of immobilize heavy metals such as, Pb, Fe and Cd, in
contaminated soils near areas such as dumpsites (Sharma et al., 2022).

Figure 9.1 Rhizosphere mineral content.

Chemical composition (Table 9.2) revealed that, in general, all analyzed PTEs (Cr, Cu,
Fe, Mn, Ni, Pb, Ti, Zn and Zr) presented higher concentration than background (Bernardo
et al., 2022). However, when compared to world reference soils used for agriculture
(Carlon, 2007), only Ti presented higher content (table 2). This difference may be
associated with the bias between local and international references which is widely
reported in many environmental contamination studies (Guerra-García et al., 2021).

Table 9.2 Chemical composition of rhizosphere (R1 to R4), background (Bkg), and
world soils (in mg/kg).

Var R1 R2 R3 R4 Bkga Worldb


Cr 18.5 13 6.8 8.9 33.5 400
Cu 184 88 24 351 9.2 200
Fe 15966 5954 3195 5480 6801 12600
Mn 153 120 93 113 140 850
Ni 14.6 4.2 2.6 4.9 3.9 45
Pb 156 39 9.7 156 7.8 300
Ti 2782 2116 1775 1703 1754 82.4
Zn 35.7 38.7 3.7 9.8 3.0 200
Zr 289 207 170 148 123 219
a
Bernardo et al. 2022b; bCarlon, 2007.

Environmental rhizosphere soil risk assessment of sample R1 revealed a low pollution


index taking in consideration background samples (PIbkg) only for Cr, while Cu, Pb, Zn
149

and Ni showed very high and high PIbkg, and Zr, Fe and Ti presented moderate PIbkg
(Table 9.3). Pollution index calculated with world soil references (PIW) showed that only
Ti concentration contributed to a very high PIW, and Fe and Zr to a moderate PIW. Sample
R2 Cu, Pb, and Zn content contribute to a very high PIbkg, and Ni, Ti, and Zr to a moderate
PIbkg, while only Ti contributed to a high PIW. Sample R3 Cu, Pb, Zr, Ti content classified
soils with moderate PIbkg, while for PIW only Ti presented a high concentration and Zn a
moderate PIW. Sample R4 Cu and Pb content accounted to a very high risk PI bkg, Zn a
high PIbkg, and Zr a moderate PIbkg, while PIW Zr and Ti concentration presented a very
high PW, and Cu a moderate PIW.

Table 9.3 Pollution index considering background (PIbkg) and world soils (PIW) and
pollution load index for background soils (PLIbkg) and world soils (PLIW) of the studied
soil samples.

R1 R2 R3 R4
Var PLIbkg PLIW
PIbkg PIW PIbkg PIW PIbkg PIW PIbkg PIW
Cr 0.6 0.1 0.4 0.0 0.0 0.0 0.3 0.0 0.0 0.0
Cu 20.0 0.9 9.6 0.4 2.6 0.1 38.2 1.8 4747.7 0.0
Fe 2.4 1.3 0.9 0.5 0.5 0.3 0.8 0.4 0.2 0.0
Mn 1.1 0.5 0.9 0.4 0.7 0.3 0.8 0.3 0.1 0.0
Ni 3.7 0.3 1.1 0.1 0.7 0.1 1.3 0.1 0.8 0.0
Pb 20.0 0.5 5.0 0.1 1.2 0.0 20.0 0.5 618.1 0.0
Ti 1.6 33.8 1.2 25.7 1.0 21.5 1.0 20.7 0.5 96454
Zn 11.9 0.2 7.2 0.1 0.4 1.2 3.3 0.1 26.4 0.0
Zr 2.4 1.3 1.7 1.0 1.4 0.8 1.2 67.6 1.6 17.1
Pollution index (PI) showed higher contamination rates in samples 1 and 4. Sample 4
was collected at the SW edge of the dump, where surface dispersion of leachate
predominates and was a contamination factor. Leachates have been identified as source
of PTEs in the surrounding of dumps (Kazemi et al., 2021; Hesami, 2021), and
associated to higher OM content that may influence the fixation of contaminants and
possible transfer (Cheela et al., 2021). Sample R1 was located in the NW edge of the
dump, strongly influenced by the open-air wastes incineration, with ashes regularly
deposited in this location. In the surrounding of dumps, incineration of urban wastes,
such as, electronics, hospital waste, tires, and lamps, have been pointed out as sources
of Cr, Cu, Mn, Ni, Pb and Zn (Itai et al., 2014; Li et al., 2022). Pollution load index
considering background soils (PLIbkg) was ranked Cu >> Pb >> Zn >> Zr > Ni > Ti > Fe >
Mn > Cr and considering world soils (PLIW) Ti >> Zr with other variables not posing a risk
of pollution, suggesting an environmental deterioration in the surrounding of the
150

dumpsite. Rhizosphere soils high level of contamination were similar to those reported
by Bernardo et al. (2022b), associated to dispersion of surface leachates and ashes.

Hazard quotient (HQ) by ingestion and dermal contact for children was considered due
to playing on the ground and hand-to-mouth behaviors in young age (Fig. 9.2). All
considered elements presented HQ > 1 by ingestion, except for Zr, being Mn the element
posing higher hazard followed by Cu, Ti, and Zn. Dermal contact HQ was > 1 except for
Pb in sample R1 to R3, and Zr in all samples, being Ni and Mn the PTEs that presented
higher hazard. Toxicity soils with high levels of Mn might induce hallucinations,
pulmonary embolism, and bronchitis (Kazemi et al. 2021), while Ti poisoning affect lung
function, breathing difficulties, coughing, skin or eye irritation (Heller et al., 2018), and
Cu chronic exposure can induce disorders of the central nervous and cardiovascular
systems (Stuckey et al. 2021).

Figure 9.2 Rhizosphere soil hazard quotient (HQ) by ingestion and dermal contact for
children (in %).

9.2 Water risk assessment


Water pH presented higher results in stream waters, being higher than reference values
(Table 9.4). Water sample WS presented a pH slightly below recommended for
consumption (WHO, 2017). Drinking water contaminated by Hg, Pb and Zn if it has an
acidic pH, its bioavailability and toxicity to the human organism is indicated as high
(Githaiga et al., 2021). Yesil et al. (2021), reported higher bioavailability and toxicity of
Hg and Pb in slightly acidic to neutral waters, such as WS, suggesting greater health risk
in WS sample, despite presenting relatively lower levels of contamination than WN.
151

Irrigation water presented higher concentrations than reference values, and prolonged
contact with water with alkaline pH level can induce dermal irritation (WHO, 2005).

Table 9.4 Water samples pH, and Pb, Hg and Zn concentration and reference values
(in µg/l).

samples ID guidelines
var drinking stream
WS WN IWN IWS reference
water water
pH 6.1 8.4 9.1 10.3 6.5−8.50 5.5 −7.5 WHO, 2017
Pb 6 nd 79 nd 10 20 WHO, 2017
Hg 36 76 nd 25 6 5 WHO, 2017
WHO, 2017;
Zn 63 13 7 9 3 ≤ 5.000
CCME, 2018
nd – not detected.

Water pollution index revealed severe Hg Pi in the two wells analyzed (Table 9. 5), while
Pb Pi showed a slight pollution in sample WS and heavy Pi in IWN. For Zn Pi was
classified as medium polluted in both wells and slightly polluted in IWS. Mercury can
incorporate into ground and surface waters through leachates produced at waste dumps
by anaerobic decomposition of wastes or by particulate material resulting from burning
of contaminated wastes (Ruengruehan et al., 2021; Githaiga et al. 2021). Waste burning
practice at Hulene-B, may be influencing the dispersion of ash enriched by heavy metals
that were successively deposited in the vicinity of the waste dump and contribute to the
contamination of ground and surface waters. Lead enrichment in areas surrounding
landfills has been associated to deposition and burning of electronic equipment, and
rechargeable batteries wastes (Ibrahim et al., 2021). In the surroundings of Hulene-B
dumpsite, other potential Pb sources may be linked to, e.g., traffic related materials
(Julius Nyerere Avenue), international airport. Zinc enrichment has been associated with
batteries, alloys, paint cans, cosmetics, pharmaceuticals, textiles and electrical and
electronic equipment waste that is continuously deposited and burnt in Hulene-B dump
(Islamd et al., 2021).
152

Table 9.5 Pollution index (Pi) of groundwater and surface water.

ID Pb Hg Zn
WS 0.6 6.0 2.1
WN nd 12.7 4.3
IWS 4.0 nd 1.8
IWN nd 5.0 0.6
nd – not detected.

9.3 Edible plants


Potentially toxic elements detected in edible plants showed higher concentration than
reference values for daily consumption (RfD) (Table 9.6), suggesting a contribution from
Hulene-B dump. Studies suggested that waste dumps can be PTEs sources for adjacent
crops (Alfaro et al., 2022), being contamination associated to transfer factor (TF) of
contaminants from soil to plant (Hermans et al., 2021). Another mechanism of edible
plants contamination in the vicinity of landfills is ash resulting from waste incineration,
transported, and deposited on crops (Weibel et al., 2017), and its consumption may
represent a potential health risk (Candeias et al., 2022b). Plants PTEs concentration was
ranked RP1 > RP3 > RP2 > RP4. Sample RP1 (Amaranthus spinosus) was collected in
the NW of the dump, an area influenced by surface water enriched in leachates, and by
waste incineration ash deposition. Associated rhizosphere soil showed high
contamination level, when compared to background. However, Zr was not detected
edible leaves despite its high concentrations in rhizosphere soils, suggesting its low TF
under alkaline pH conditions (Odom et al. 2021; Yesil et al. 2021). Sample RP3 (Brassica
oleracea L) located SW, relatively distant from the dump, showed high PTEs levels.
However, rhizosphere soil contamination was relatively low, suggesting other
contamination mechanisms (Bernardo et al., 2022b). Sample RP2 (Ipomoea batatas)
was collected relatively close to the dump, with PTEs content on rhizosphere soil higher
than background, and plant edible leaves revealing a PTEs concentration above
reference values. This suggested a possible capacity of Ipomoea batatas to absorb
metals from soil. Kacholi et al. (2018) showed that Ipomoea batatas in contaminated soils
presented higher capacity to absorb heavy metals such as Cu Fe, Pb and Zn. Sample
RP4 (Cucurbita pepo), was collected at the S boundary close to the landfill, and its
rhizosphere soil presented higher PTEs concentration than local background. Plant
edible leaves showed content higher than reference values for consumption.
153

Table 9.6 Chemical analysis of potentially toxic elements in edible plants.

Var RP1 RP2 RP3 RP4 AL


Cr 974 852 2968 1161 2.3a
Cu 1142 1245 778 1099 10-40a
Fe 56324 30031 32643 18687 18b
Mn 1030 1183 1436 1544 500a
Ni 108 238 1128 538 10a
Pb 394 56 190 nd 0.3a
Ti 4990 5910 8392 3519 0.7c
Zn 5730 2596 1511 3519 50a
Zr nd 761 708 198 4.2d
AL – acceptable limit; aGebeyehu et al., 2020; bHurrell et
al., 2010; cJovanović et al., 2018; dUSEPA, 2010.

9.4 Data integration

Transfer factor (TF) from soil to plants is an important component for toxicity, being soil
pH and OM relevant in this process. Studied samples TF was > 1 (Fig.9.3), except for
Pb in R4P and Zr in R1P, not detected in plant samples. High TF values suggested a
high capacity of the plants to absorb contaminants (Balasooriya et al., 2021; Alfaro et al.,
2022). Transfer factor were ranked RP3 > RP4 > RP2 >RP1. Sample RP3 presented the
highest TF for all PTEs, suggesting a greater risk to consumers health, being Pb the
most representative element. Lead can be absorbed from plants leaves by ashes or road
dust deposition, being this plant location on SE wind direction over the Hulene-B landfill,
and Julius Nyerere Avenue on the E (Bernardo et al., 2022b). Businelli et al. (2009), and
Wang et al. (2022) suggested that plant leaves grown in the surrounding of dumps are
constantly being enriched by PTEs from ashes resulting from waste incineration. Sample
RP4 Zn and Mn TF showed higher percentage rate. High TF values in all samples
suggested potential toxicity to humans.
154

Figure 9.4 Transfer factor (TF) from soil to plants.

Daily elemental intake (DIM), and hazard risk index (HRI), are presented in Table 9.7.
Sample R1P (Amaranthus spinosus) DIM was ranked Fe >> Zn > Ti, being other PTEs
neglectable. The HRI was high in all elements, except Ni, and Zr, being ranked Cr > Cu
> Pb > Zn > Mn > Ti > Fe. Sample RP2 (Ipomoea batatas) DIM was > 1 in Fe, Zr and Ti,
while HRI > 1 was ranked Cu > Cr > Mn > Pb. Sample RP3 (Cucurbita pepo) DIM >1
was ranked Fe > Cr > Ti and HRI Cr > Cu > Pb > Ti > Mn > Ni > Zn. Sample RP4
(Brassica oleracea L) DIM > 1 was ranked Fe > Zn > Ti while HRI Cr > Cu > Pb > Zn >
Mn > Ti > Fe. Consumption of plants contaminated by PTEs can induce various diseases
and even cancer (Ashraf et al., 2021). Chromium toxicity in humans has been linked to
damaged blood cells, livers, nervous systems, and kidneys (Kennou et al., 2015).
Chronic Cu consumption may result in gastrointestinal irritation, diarrhea, and liver failure
(Ali et al., 2017), while Fe toxicity can cause vomiting, diarrhea, and damage to the
intestine and other organs (Hurrell et al., 2010), and Mn is linked to memory problems,
hallucinations, and Parkinson’s disease (USEPA 2016). Nickel chronic ingestion has
been associated with gastrointestinal and neurological effects including lung cancer
(CCME, 2015). Ingestion of Pb can disturb almost all organs and the nervous system, in
addition to causing kidney damage, brain damage, miscarriage, and death (Wani et al.,
2015), Wani et al. (2015) reported that pregnant women who ingest high levels of Pb
may have pronounced intoxication levels, which increases the risk of premature birth or
low birth weight newborns. In the sample area, women and children are the ones who
155

grow and prepare the food, suggesting greater exposure and risk (Rogerson, 2017). Ti
can cause damage to DNA, brain, as well as liver and kidneys, and can cause cancer
(EFSA, 2018), while Zn excessive and prolonged intake can lead to anemia and affect
the immune system (Kennou et al., 2015; Price et al., 2021). Zirconium is classified with
low toxicity (Jones et al., 2017) despite the presence and retention in relatively high
amounts in biological systems, has not yet been associated with any specific metabolic
function (USEPA, 2010).

Table 9.7 Daily elemental intake (DIM), and hazard risk index (HRI).

RP1 RP2 RP3 RP4


Var
DIM HRI DIM HRI DIM HRI DIM HRI
Cr 0.41 135.29 0.36 118.37 1.24 412.17 0.48 161.26
Cu 0.48 128.58 0.52 140.23 0.32 87.61 0.46 123.68
Fe 23.47 1.30 12.51 0.70 13.60 0.76 7.79 0.43
Mn 0.43 3.07 0.49 3.52 0.60 4.27 0.64 4.60
Ni 0.05 0.23 0.10 0.50 0.47 2.35 0.22 1.12
Pb 0.16 11.73 0.02 1.67 0.08 5.65 nd nd
Ti 2.08 2.97 2.46 3.52 3.50 5.00 1.47 2.09
Zn 2.39 7.96 1.08 3.61 0.63 2.10 1.47 4.89
Zr nd nd 0.32 0.08 0.30 0.07 0.08 0.02
nd – not detected.

Targeted hazard quotient (THQ), and hazard index (HIplant) is presented in Figure 9.4,
being THQ ranked Cr > Cu > Pb > Zn > Mn > Ni > Fe > Zr. Sample RP1 (Amaranthus
spinosus) contribution to induce adverse health effects in children and adults was ranked
Cr > Cu > Pb> Zn > Mn > Ti > Fe, being nickel < 1 and not posing potential risk for
adverse health effects. Sample RP2 (Ipomoea batatas) Fe, Ni and Zr content do not
posed a potential health risk, being the main contributions posed by Cu > Cr > Zn > Ti >
Mn > Pb. Sample RP3 (Brassica oleracea L) the main contributors to potential health
outcomes were ranked Cr > Cu > Pb > Ti > Mn > Ni > Zn, being Fe and Zr content
representing < 1. While sample RP4 (Cucurbita pepo) PTEs were ranked Cr > Cu > Zn
> Mn > Ti > Ni, with Fe and Zr < 1. In all samples the highest THQ risk index was posed
for children. Similar findings were reported as a result of their lower body mass,
suggesting a higher likelihood of developing cancer throughout life (Candeias et al.,
2022).
156

Figure 9.4 Targeted hazard quotient (THQ) in edible plants.

HIplant was similar to THQ, being > 1 in all samples and higher for children (Fig. 9.5),
being ranked RP3 (6.41E+02 to 4.27E+02) > RP4 (3.68E+02 to 2.45E+02) > RP1
(3.59E+02 to 2.39E+02) > RP2 (3.36E+02 - 2.24E+02). Long-term consumption of
studied plants may induce cancer (e.g., kidney, bladder and respiratory tract), nervous
system disorders, memory problems, hallucinations, cardiovascular diseases and
diabetes. In children has also been associated to cognitive development problems and
aggressive behavior (WHO, 2021a; WHO, 2021b).

Figure 9.4 Hazard index in edible plants.

Wells water samples systemic toxicity (non-carcinogenic hazard) presented values


above 1.00E+00 (Fig.9.6). Results are linked to Hg concentration, being Pb and Zn
considerable negligible. Mercury is a toxic element, being ranked third most toxic among
all PTEs (Khattak et al., 2021). It was estimated that chronic Hg consumption of about
157

0.00023 mg/l per day can induce adverse renal, neurological, and respiratory effects
(Ekawanti et al., 2021; Fonge et al., 2021), and can irreversibly damage kidneys, liver,
and central nervous system (Fiala et al., 2021). Wells waters do not present a
carcinogenic risk for humans by direct consumption, with Pb showing the highest values
of 1.09E-07 and 6.55E-07 for WN and WS, respectively, both below the threshold of
1.00E-06. Lead implication on human health depends on intensity, duration, and level of
exposure, that can result in a range of toxic effects such as hematological, neurological,
psychological, renal, genetic mutation and reproduction are examples of the cumulative
effect on the body (Njoku et al., 2020). As expected, in wells water, children had a higher
health risk than adults, due to their lower body mass (Candeias et al., 2022). Lead toxicity
in children can induce neurological complications, including difficulties in learning,
concentration, and aggressive behavior (Wani et al., 2015). Water ingestion is
considered one of the main forms of Pb poisoning in children, accounting for about 7%,
(Isa et al.,2015).

Figure 9.4 Systemic toxicity of wells water samples.


158
159

10 Paper 7 (Appendix 8)

Maputo urban area (Mozambique) road dust characterization and risk


assessment

Abstract

Roads are important source sources of pollutants, generating by wind and turbulence
generated by cars road dust particles with impact in the environment and human health.
In this study, road dust samples, fractions < 63 and 2000 µm, were collected in Maputo
urban area (Mozambique). Dust was characterized on their physical, mineralogical
(XRD), and geochemical (XRF) content. Environmental and human health risk was
assessed. Mineralogical analysis revealed the predominance of quartz (SiO2) >>
phyllosilicates (mainly micas and kaolinite) > K-feldspars (KalSi3O8) > plagioclase
((Na,Ca)((Si,Al)AlSi2)O8) > carbonates (calcite (CaCO3), dolomite (CaMg(CO3)2) >
sulphates (alunite (KAl3(SO4)2(OH)6), anhydrite (CaSO4)), revealing a greater capacity
for airway destruction and adsorption of PTES. Contamination index taking in
consideration Potentially toxic elements (PTEs) Cr, Cu, Fe, Mn, Ni, Pb, Ti, Zn, and Zr
revealed concentration above guidelines. Carcinogenic risk of the <63 µm fraction was
due to road dust Ni and Pb concentration. Same fraction hazard quotient, or non-
carcinogenic hazard, suggested that ingestion was the main exposure route for children,
mostly due to, Cu, Fe, Ni and Pb content.

Keywords: road dust, mineralogy, geochemistry, risk assessment, Maputo

Reference: Bernardo, B; Candeias, C; Rocha, F: Maputo urban area (Mozambique)


road dust characterization and risk assessment. Submitted.
160

10.1 Results and discussion


Dust samples pH, electrical conductivity (EC), and color are presented in Table 10. 1.
Samples < 63 µm fraction presented a pH predominantly neutral, except for sample 9,
slightly acid, and samples 6 and 8, highly alkaline to slightly alkaline. Sample 9 was
collected on a roads intersection of entrance and exit of the city, with intense traffic.
Samples 6 and 8 were taken one smaller roads with squares, with roads with significant
degradation, being sample 8 from a relatively more degraded road. Fraction < 2000 µm
samples 1, 2, 5, and 9 presented neutral pH, samples 3, 9, and 7 slightly acidic, and
samples 6, and 8 highly alkaline. Sample 3 was collected near Hulene-B waste dump,
with pH reflecting its influence. Sample 7 was collected in National Road nº 1, with
intense traffic, and sample 9 also collected on degraded road with intense traffic. In
general, acidic and neutral pH was identified in samples collected on degraded pavement
roads and close to contamination sources such as, Hulene-B dump and International
Airport (samples 1, 2, 3, 4, 5, 7 and 9). Alkaline pH was found in samples located on less
degraded roads and relatively far from previous contamination sources. Previous studies
in road dust suggested a prevalence of acidic pH as an indicator of higher PTEs
concentration (Nematollahi et al., 2020).

Table 10.1. Dust samples < 63 and 2000 µm fractions pH, electrical conductivity (EC;
μS/cm) and color.

pH EC color
ID < 63 < 2000 < 63 < 2000
< 63 µm < 2000 µm
µm µm µm µm
Light brownish
1 6.86 7.16 251 425 Brown
gray
Grayish
2 6.80 7.09 292 660 Grayish brown
brown
3 6.52 6.35 310 1051 Dark gray Gray
4 7.13 7.48 130 224 Gray Pale brown
5 7.38 6.95 139 218 Dark gray Gray
6 8.62 10.41 252 610 Gray Gray
Grayish
7 6.97 6.41 128 362 Dark gray
brown
8 7.70 8.82 106 180 Dark gray Gray
Grayish
9 6.28 6.50 521 930 Gray
brown

Dust samples EC presented higher values in < 2000 µm fraction. Road dust is a complex
mixture that includes organic matter (OM), which is pointed out as a factor for EC
increase in sandy materials (Kasimov et al. 2019; Pariente et al., 2019). Samples with
lower EC values were the ones collected in less degraded roads (samples 4, and 5) and
161

widened roads (sample 8). Samples EC in < 63 µm fraction was ranked 9 > 3 > 2 > 6 >
1 > 5 > 4 > 7 > 8, and in fraction < 2000 µm 3 > 9 > 2 > 6 > 1 > 7 > 4 > 5 > 8, being the
highest EC found in samples collected in locations with degraded roads, intense traffic,
near International Airport, and Hulene-B waste dump. Dust samples color varied slightly
in both fractions analyzed (Table 10.2). Road dust color is influenced by particles origin
and composition (Yukioka et al. 2020). In general, darker colors are linked to the
presence of organic matter and clay materials, pointed as relevant for adsorption and
absorption of PTEs (Aryal et al., 2015; Vlasov et al., 2022). Dust samples granulometric
distribution is presented in Table 2. Sand fraction ranged 93.2 to 98.3 %, in samples 2
and 7, respectively, with all samples classified as sand. Predominance of the sandy sizes
particles is associated with source materials, and surrounding soils (Momade et al.,
1996).

Table 10.2 Dust samples granulometric distribution (in %).

1 2 3 4 5 6 7 8 9
sand 95.54 93.22 94.47 94.95 97.33 97.50 98.34 94.51 94.58
silt 2.67 4.38 3.27 3.04 1.36 1.54 0.80 2.34 3.33
clay 1.79 2.40 2.26 2.02 1.31 0.96 0.86 3.15 2.09

Identified mineralogical phases of the <2000, and 63 µm fractions, is presented in Figure


2. In all samples, quartz (SiO2) was the predominant mineral, ranked 97.6 > 91.0 > 90.6
> 89.7 > 89.5 > 87.3 > 86.4 > 77.3 > 73.7 % in < 2000 µm fraction samples 3, 1, 6, 9, 8,
7, 2, 4, and 5, respectively. In fraction < 63 µm, also quartz was the main mineral
identified, ranging 20.6 to 66.6 %. Quartz content, in all samples and fractions, is linked
to local soils and transported dusts mineralogical composition (Momade et al., 1996;
Bernardo et al., 2022). Silt fraction revealed smaller quartz content, suggesting that its
particles had higher sizes than 63 µm. Quartz presents higher structural hardness than
some other minerals, that helps preventing physical weathering and abrasion (Candeias
et al., 2022). However, Bai et al. (2021) suggested that, in smaller quartz particles with
structural cracks at temperatures > 20 ºC, adsorption of PTEs, such as Pb, can occur.
Prolonged contact with high quartz content dust can induce diseases such as silicosis
and pulmonary fibrose (Leung et al., Yu; Chen 2012; Koga et al., 2021). Respirable
crystalline silica poses a risk to human health when exposed over long term (Horwell et al.,
2012). Other minerals identified were, phyllosilicates (mainly micas and kaolinite)
detected all samples < 2000 µm fraction, with exception of sample 3, collected near
162

Hulene-B waste dump; K feldspars (kAlSi3O8), and plagioclase ((Na,Ca)((Si,Al)AlSi2)O8)


were detected in all samples < 2000 µm, with highest contents in samples 3 and 4. In
many studies it has been revealed that micas and kaolinite have a higher adsorption capacity for
PTEs (Murray, 2000; Odukoya et al., 2023), which can then be incorporated into the human body
through respiratory, inhalation and ingestion routes, creating various health complications
(Alghamadi et al., 2023).

In the < 63 µm fraction, K-feldspars were detected in considerable amounts in all


samples. Carbonates (calcite (CaCO3), dolomite (CaMg(CO3)2) were detected only on
the thinner fraction. Sulphates alunite (kAl3(SO4)2(OH)6) and anhydrite (CaSO4) were
detected in all samples, with higher alunite concentrations in samples 3, 5 and 2. Oxides
magnetite-maghemite (Fe2O3-γFe2O3), ilmenite (FeTiO3), and pyrite (FeS2) were
detected in almost all < 63 µm fraction samples, suggesting higher adsorption of Fe
oxides in the thinner fractions. Nsutite (Mn4+, Mn2+)(O,OH)2) was detected in sample 7
revealing a higher affinity of Mn oxides. Results showed a heterogeneous mineralogical

composition of the sampling areas, with remobilization and deposition of materials


transported from other locations (Momade et al., 1996).

Figure 10.1. Dust samples identified mineral phases of the < 63 and < 2000 µm
fractions relative distribution (m-m – magnetite maghemite). *Quartz is not projected

Independent samples T-test showed significant differences (p < 0.05) between fraction
< 2000 µm and < 63 µm, with higher concentration in the thinner fraction (Table 10.3).
Previous studies reported that thinner fractions presented higher toxicity, with higher
ability to adsorb PTEs (Al-Shidi et al., 2021; Bian et al., 2015; Lin and Wu, 2015). This
163

fraction is also easily resuspended by traffic and wind, allowing incorporation and
dispersion in the air, and dermal adherence, by exposed population (Aguilera et al.,
2021; Al-Shidi et al., 2021; Candeias et al., 2021, Paula et al., 2021).

Table 10.3 Dust samples potentially toxic elements chemical composition of the < 63 and
2000 µm fractions (in mg/kg).

ID Cr Cu Fe Mn Ni Pb Ti Zn Zr
< 63 µm
1 10400 1210 304020 5030 1040 1360 39450 1080 5860
2 4420 1130 246970 3940 690 nd 46770 1850 4850
3 19660 2750 321950 4160 1320 2240 38680 2320 3500
4 19850 1970 317010 7800 900 2240 42760 2340 7240
5 11090 1620 311600 6340 1410 1320 34540 2160 4860
6 12880 1470 306410 6220 1090 4450 38640 1350 4160
7 5510 1520 385810 8270 3540 nd 36810 420 830
8 2770 800 332150 5130 1260 nd 53440 nd 440
9 2590 840 196420 3280 690 nd 47540 nd 800
< 2000 µm
1 2560 370 102580 1220 740 nd 15410 nd 300
2 2450 450 118330 1600 760 nd 30620 420 2300
3 7820 1070 138060 1270 600 nd 18140 390 1240
4 5510 870 135230 2990 570 720 23250 nd 1370
5 3230 730 84400 1520 620 nd 12630 380 1650
6 3140 510 76700 1720 nd 1360 16520 nd 1150
7 1200 nd 65530 1720 nd nd 15850 nd 870
8 6520 1010 121730 1530 580 nd 18370 620 2940
9 980 430 90660 1490 nd 830 21040 530 3770
GV 75.8a 27.1a 12600a 615.6a 39.1a 19.6b 1618a 82.4a 219.0c
nd – not detected; GV – guide values; aAli et al., 2017; bZhao et al., 2021; cZhang et al.,
2019.

Lowest Cr concentration, in both fractions, wase found in sample 9, collected on an


intense traffic road, while highest values were detected in samples 3, and 4, collected on
degraded asphalt roads near Hulene-B waste dump, and near the airport, respectively.
Cr has been associated with intense traffic, asphalt degradation, and wastes incineration
that contain paints, varnishes, organic solvents, and oils (Chaudhary et al., 2021).
Copper presented similar pattern, with sample 3 revealing the higher concentration also
on the < 63 µm fraction, above the 27.1 mg/kg guideline (Table 3), while samples 9 and
8 with lower Cu content. This suggests a common enrichment source, e.g., degraded
pavements, brakes on bus stops, tires (ZnO and Cu/Zn layers formed during
vulcanization). Particulate wear and resuspension have been considered main sources
of Zn, and Cu in urban areas (Valotto et al., 2015). Samples with highest Fe
concentrations in both fractions were collected near the waste dump (sample 3), and
degraded pavements and heavy traffic (samples 4 and 8), suggesting anthropogenic
sources of enrichment, when compared to local soils content (Bernardino et al., 2022).
164

Iron has been linked to metal alloys production, that are commonly used in vehicles
components, e.g., main component on steel and associated rust). Samples with higher
Mn concentration were collected on roads with less degraded pavement but intense
traffic, i.e., National nº 1 road, the main road for entry and exit of Maputo city. Mn has
been associated with tire and brake wear and is used to prevent corrosion and
deformation of car components in (Candeias et al., 2020). In all samples Ni content was
higher than guide values of 39.1 mg/kg (Table 3), and differences between analyzed
fractions may be associated to a higher adsorption capacity by thinner materials (Bisht
et al., 2022). Ni has been used in the production of stainless steel and other Ni alloys,
giving high corrosion and temperature resistance in automobiles, and in waste electronic
equipment, rechargeable batteries, paint cans, varnishes, organic solvents, and waste
glass (CCME, 2015), materials that are deposited in Hulene-B waste dump (Bernardo et
al., 2022). Lead higher concentration was found in samples < 63 µm fraction samples 6
and 3, located in areas with intense traffic and degraded asphalt, both near Hulene-B
dump and Maputo international airport, considered anthropogenic sources of Pb in urban
areas (Bernardo et al.,2022b). However, when detected, Pb content was higher than
guideline 19.6 mg/kg, suggesting differentiated anthropogenic sources. PbO4 is a
component used in brake friction materials, and a gasoline additive used in Mozambique
until 2005. In samples 2, 7, 8 and 9, Pb was below detection limit, what may be
associated with leaching processes and topsoil removal by wind, given the greater
openness of these square and the strong influence on the continuous removal of surface
dust at these points by wind. Kasimov et al. (2019) a similar study in Moscow showed
that more open streets have a greater capacity to remove and leach out heavy road dust
enriched with heavy metals such as Pb. Zn was detected mainly in < 63 µm fraction,
being samples 3 and 4 the ones with highest content, collected in degraded asphalts
with intense traffic, higher than the guideline (82.4 mg/kg). Zn has been used in road
pavement materials, and as an engine oil additive (Sternbeck et al., 2002), suggesting
that fuel combustion may be a significant source. Zirconium concentration in urban
environments has been linked to vehicle exhaust emissions, as mixed Zr oxides
(CeO2/ZrO2) have become an essential component of three-way catalysts (Piatak and
Bedinger, 2017).

Nemerow Index was calculated for < 63 µm fraction, using both soils background (PNbkg),
and soils surrounding Hulene-B dump (PNHB) mean concentration (Bernardo et al.,
2022), to assess the contribution of soils to road dust Cr, Cu, Mn, Ni, Pb, Zn, and Zr
concentration (Fig. 3). Dust PN using background values revealed higher contamination
165

than when considering Hulene-B surrounding soils, suggesting an anthropogenic


contribution of the Hulene-B dump associated with leachates and waste incineration
ashes. Another fact to consider is that the high levels of PTEs may be associated with
the sampling season (June) marked by scarcity of precipitation in the study area, which
is pointed out as a factor for heavy metal accumulation in road dust in many studies (Men
et al., 2023; Davoudi et al., 2021). Samples 3, 4 and 7 PNbkg showed the higher index
values, revealing the influence of the dump to dust contamination. Sample 7 PNHB was
the highest among all samples suggesting the influence of other anthropogenic sources
to the PTEs content, e.g. intense traffic, high degradation of Lurdes Mutola avenue
pavement. Sample 9 presented the lowest values in both indexes, which may be
associated to resuspension processes, a factor that disperse contaminants (Valotto et
al., 2015).

Figure 10.2. Dust samples Nemerow Index using soils background (PNbkg), and soils
surrounding Hulene-B dump (PNHB) mean concentration (Bernardo et al., 2022d).

Individual contributions of Cr, Cu, Mn, Ni, Pb, Zn, and Zr in the Nemerrow index were
rather heterogeneous (Fig. 4). Based on the background soils, i.e. considered as soils
with little influence of anthropogenic action, the variation was very heterogeneous. In
samples 1 to 6, the contributions in descending order were Zn> Ni> Cr>, Cu> Pb>Zr>Mn
and in samples 7 and 9 Ni> Cu>Cr>Mn>Zr>Zn. This reveals the existence of diversified
sources of enrichment. Pb, however, in samples 8 and 9 showed a lower contribution.
Based on the Hulene-B background, in samples 1 to 6, all elements showed significant
contributions in the Nemerrow index. In these samples the contributions in descending
166

order were Zn>Cr>Ni>Pb>Zr> Mn. In samples 7 to 9, Ni>Cu>Cr>Mn>Zn>Zr showed the


highest contributions in the Nemerrow index. This reveals that samples 1 to 6 have varied
sources of enrichment by PTEs. In samples 7 to 9, a Pb contribution may be associated
with the intense leaching that these samples are exposed to and the intense wind
depletion due to a larger aperture of the squares.
In many studies it has been reported that Pb accumulates on the soil surface (Binh et
al., 2021), which reveals its higher vulnerability to be mobilized in more open areas
(Kosheleva et al., 2019). It is also associated with the fact that the Hulene B reference
has higher Pb levels than the local background, which influences the Nemerrow index
results for Pb in this investigation.

Figure 10.3. Relative distribution of elemental contribution for Nemerow Index using soils
background (PNbkg), and soils surrounding Hulene-B dump (PNHB) mean concentration
(Bernardo et al., 2022d)

Risk assessment was calculated for fraction < 63 µm due to the higher PTEs Cu, Fe, Ni,
Pb, and Zr concentration (Fig. 5). However, all samples represent a health risk due to
their high contamination level than the reference values.

The hazard quotient (HQ), or non-carcinogenic hazard, suggested that ingestion was the
main exposure route, and children the ones with higher HQ, because of their hands-to-
mouth habit, while dermal contact and inhalation routes presented a low hazard. Another
aspect is the fact that children have a significantly lower body mass (Candeias et al.,
2020) and that young children absorb a greater proportion of the Pb ingested than adults
167

due to dietary deficiencies in iron or calcium, which are factors in greater absorption
(WHO 2021). In many studies it has been reported that women are very exposed to road
dust in many African cities (Friberg et al., 2021). In Maputo city in particular, women and
children do most of the commercial activities along the roads and in the main squares,
which can also increase their risk due to greater exposure to contaminated road dust
(Rogerson, 2017). In parallel, (Wani et al.,2015) reported that pregnant women with a
history of exposure to high lead-containing dusts may have high levels of lead in their
blood which can lead to the risk of premature birth or low birth weight babies.

Figure 10.4. Dust samples, fraction < 63 µm, hazard quotient by ingestion (HQing) in
children and adjusted for children and adults by potentially toxic elements and their
sum.

Carcinogenic risk was only identified in PTEs Ni and Pb (Fig. 10. 6), being Ni the one
representing a higher risk in samples 2, 7, 8 and 9, collected in locations with diverse
commercial activities, such as street food vendors. These spots are characterized by
intense road traffic. Begum et al. (2022) reported that nickel is emitted from inefficient
vehicles with difficulties in the total burning of fuel by engines. Zhao et al. (2021) reported
that Ni is mainly inhaled in the form of insoluble compounds, and nickel carbonyl
(Ni(CO)4) vapors that can cause chronic bronchitis, decreased lung function and cancer.
It has been reported that inhalation of doses of nickel carbonyl enriched dusts can
immediately cause severe lung diseases, including coughing, diffuse interstitial
pneumonia and cerebral hemorrhage (Begum et al. 2022). In dermal exposure allergies
and dermatitis are reported (CCME,2015). Nickel aerosols (e.g. Ni(II) sulphate) have
recently been found to be carcinogenic to the human respiratory tract (Schaumlöffel,
168

2012). Samples 1, 3, 4, 5 and 6 presented carcinogenic risk due to Pb content. These


samples were collected near populated places, including schools (samples 4, 5 and 6).
Kazemi et al. (2021), suggested that Pb exposure can induce seizure disorders,
aggressive behavioral disorders, chronic abdominal pain, anemia, and cancer. In turn,
WHO (2021) reported that children are particularly vulnerable to dust pollution by Pb,
and estimated that lead exposure is responsible for 30% of the global burden of idiopathic
development and intellectual disability.

Figure 10.5. Dust samples Risk by Pb, Ni and combination of both.


169

Part III Conclusions


170
171

11 Conclusions
The integration of geophysical (electrical resistivity) and geochemical prospection data
in the surrounding of the Hulene-B dump demonstrated a complementarity to understand
the environmental (soil, groundwater, surface water and edible plants) and human health
impacts of the dump.

Data acquisition and analysis at two stations was relevant to understand the different
mechanisms of contamination by leachate plumes in the subsoil and groundwater as well
as the contamination of soils and edible plants around the Hulene -B dumpsite.

The application of the electrical resistivity method to study the dynamics of contamination
plumes between 2020-2021 in the surroundings of the Hulene - B dump has allowed the
identification of areas of anomalous resistivity which we consider as leachate production
zones and migration of contamination plumes towards lithological substrates and
consequent dilution and contamination of groundwater.

The analysis of the dynamics of the possible contamination plumes showed that in
summer (2020) they occupy an extensive area and with strong horizontal and vertical
migration, which may be associated with the greater production of leachates given the
greater decomposition of waste in the hot and rainy period (Profiles 1a, 2a). On the other
hand, in winter (2021) the contamination plumes were thick and with a predominantly
vertical movement, which allows greater vertical migration to great depths causing
extensive subsurface anomalies in the lithologies above the groundwater (Profiles 1b
and 4b). In the north, the reference profile (3) showed, exceptionally, that the southern
part is partially affected by the large plume described in Profile 2, without significant
variations in both analysed stations of the areas occupied by anomalous resistivities.

The contamination plumes in the two study seasons (2020-2021) show a predominantly
E-W movement (Profiles 1 and 2). In the southwest (Profile 4), the plume assumes a NE-
SW movement and its enrichment is associated with the new deposition at the SW
boundary of the natural leachate reception basin. The spatialized electrical resistivity
models at the two studied stations show that the resistive anomalies in the subsurface
environment decrease as one moves away from the dump to the north (profile 3), which
reveals the attenuating effect of groundwater and lithological substrates. Thus, the
resistivity models proved to be efficient for the implementation of structural measures to
monitor possible leachate contamination fluxes in the lithologies, surface waters and
groundwater around the Hulene - B dump, Maputo, Mozambique.
172

The combination of electrical resistivity and modified DRASTIC models was effective in
describing hydrogeological features, estimating groundwater vulnerability in detail, and
identifying areas of possible leachate migration to groundwater in the vicinity of the
Hulene-B dump. The areas covered by profiles 1 (west), 2 (north) 4(northwest -
southwest) and 6 (east) showed strong evidence of possible groundwater contamination
by leachate plumes, the modified DRASTIC model was rated high in this area (169-174)
due to the proximity of groundwater to the contaminating surface (dump) and the
connection of continuous anomalous layers from the surface to the aquifers. In the area
of profiles 3 and 5 (northern and southern edge of the dump), the index was low (91-96)
due to the strong resistivity of the surface layers and the high depth of groundwater
detected (> 40 m). However, higher vulnerability in areas covered by profiles 1, 2, 4 and
6 located in sections with high altitudes (2 and 6) suggests that groundwater
contamination occurs by horizontal dilution throughout the area. Validation of the
geophysical data such as groundwater depths were similar to those of the electrical
resistivity models and chemical analyses showed high phosphate concentrations,
alkaline pH in the north, and high levels of Hg and Pb above the WHO recommended
average for human consumption. The study of soil properties between 2020 (rainy
season) and 2021(dry season), demonstrated the influence of climatic conditions on the
mechanisms of its environmental contamination of the surroundings of the Hulene-B
dump.

The results suggest a strong alteration when compared to background samples. This
suggests a possible contamination and conditions for leachate migration into the subsoil
and groundwater. Electrical conductivity showed more significant alterations, with values
considered high in all samples around the dump, what can be associated with possible
contamination of soils by heavy metals present in leachates and ashes from waste
incineration.

The predominance of sandy fraction suggested vertical migration of contaminated


leachate in depth. Soils pH classified from neutral to slightly alkaline, and a low OM
content suggested a higher leaching and migration capacity of contaminants in depth.
The landfill pollution index (Ip) was rated very high (16.3) suggesting a possible migration
of contaminants to groundwater, a similar result to soil properties. Results highlight the
importance of the dump monitoring and the systematic assessment of soil and
groundwater contamination levels.

Soil properties data for the dry period (2021) show better conditions for contaminant
accumulation in topsoil. The sandy fraction was predominant in both seasons. The OM
173

content was relatively high in the dry period and was characteristic of a larger number of
samples. The dark brownish color occupied larger bands in the winter 2021 samples,
evidenced low leaching. Soil EC was extremely high in the winter samples, suggesting
higher retention of contaminants at the surface. pH showed no significant changes,
suggesting that it is not a determining factor in contaminant retention and migration in
soils surrounding the Hulene-B dump. The risk of surface water contamination was high
(0.97), consistent with rainy season indices, suggesting that the dump is a constant
vector of contamination (soils, surface water and groundwater) in the surroundings of the
Hulene-B dump. Nevertheless, these results are good indicators for the need to take
structural measures in the control of the main contamination vectors, mainly leachates
and bottom ash from waste incineration.

Chemical and mineralogical analyses of the soils, collected in summer 2020, showed
highly variable PTEs concentration levels around the Hulene - B dump. Analysis of
environmental degradation levels using local background values and universal soil
references was extremely different. They were much higher in all local references and
heterogeneous in the global reference. Thus, Hazard indices PLI, PN, and PERI ranged
from low to moderate in samples collected in central west area of the dumpsite, whereas
samples closer to the dumpsite and from southwestern edge, showed a predominantly
high ecological hazard. Nemerov hazard using background mean concentrations (PNbkg)
revealed a predominant high hazard (90.1%) in studied area, suggesting high level of
environmental deterioration. Pollution load index indicated that individual elemental
contribution was ranked Zn >> Cu > Pb > Cr = Zr > Ni > Mn; with PERI Zn >> Cu > Cr >
Cr > Zr > Pb > Ni > Mn; and with PNbkg Zn >> Cu > Pb > Cr = Zr > Ni > Mn. Carcinogenic
risk for both children and adults was significant only due to Pb concentrations in the
immediate northwest area of the dump (6.19 x 10-6) and southwest of the densely
inhabited dump (2.94 x 10-6). Given the toxicity of Cr, Cu, Mn, Ni, Pb, Zn and Zr,
especially for children, measures to control and mitigate soil contamination in the
surroundings of the Hulene-B dump are pertinent.

Rhizosphere soils showed a higher pollution load for Pb, Cu, Zn, Zr and in edible plants
for Zn, Ni, Cr, Cu, Mn, Pb, Fe, Ti, Zr, suggesting contamination by soil transfer of Pb, Cu,
Zn and Zr and by possible deposition of ash contaminated for Ni, Cr, Mn, Fe, Ti and Zr.
Well water and irrigation water showed elevated levels of Hg, Pb and Zn showing
potential health hazards. The health hazard index of rhizosphere soils showed higher
levels by ingestion and in children along with edible plants. the water toxicity index was
high due to high levels of Hg and Pb contamination.
174

The study of road dust in Maputo city taking the local and Hulene-B background as
reference showed heterogeneity in physical, mineralogical, and chemical characteristics
of road dust. The evaluation of the physical characteristics of road dust pH, EC, OM and
particle size showed heterogeneous characteristics that confer greater absorption
capacity and toxicity to road dust. Mineralogical analysis showed the predominance of
Quartz (SiO2) >, phyllosilicates >, feldspars (KAlSi3O8) >, plagioclase
((Na,Ca)((Si,Al)AlSi2)O8)>, Carbonates (Calcite (CaCO3)>, dolomite (CaMg(CO3)2) >,
aluminite sulfates (KAl3(SO4)2(OH)6) >, anhydrite (CaSO4) which presents higher risks
of physical airway destruction and aggregation of PTEs. Chemical analyses showed
higher contamination of Cr, Cu, Fe, Mn, Ni, Pb, Ti, Zn, and Zr, which demonstrates a
higher capacity for human toxicity. Health hazard ratios show higher contribution of Cu,
Fe, Ni, and Pb, especially for children, showing a higher probability of associated
diseases. Carcinogenic effects were identified for Ni and Pb, especially for children,
whose extremely high concentrations were identified for all samples, suggesting the
need for continuous monitoring and the definition of mitigation measures for the
protection of the population.

The integration of geophysical and geochemical methods proved to be efficient to


understand the environmental impacts of the Hulene-B dumpsite and the risks to human
health. The results suggest the need for continuous monitoring and the definition of
measures to mitigate the impacts of the dumpsite, given the high levels of contamination
of the surrounding ecosystem and the higher exposure of the local population, especially
children, who represent the most vulnerable group to the harmful effects of potentially
toxic elements in the short and long term. However, further studies are needed, given
the heterogeneous mechanisms of environmental contamination that the Hulene-B dump
presents.

11.1 Future Studies


• to assess the areas of suspected vertical migration of leachate through chemical
analysis of the subsoil, especially the areas of suspected vertical migration of
leachate that binds continuously from the surface to the groundwater;
• to assess the chemical composition of incineration ashes in order to evaluate
their contribution to the contamination of soils, water and edible plants by Hulene-
B;
• to seasonally assess the levels of chemical elements in leachates and evaluate
their contribution to environmental contamination;
175

• to understand the bioaccessibility of metals in soils, edible plants, water and


identify bioaccumulation patterns in indicators such as biological uptake
coefficient for each element;
• to study the indoor dust in the surroundings of the Hulene-B dump and in
residential areas close to the roads studied ;
• An integrated model for contamination and remediation of environmental and
public health impacts needs to be designed.
176
177

12 References
Abdulrahman, A., Nawawi, M., Saad, R., Abu-Rizaiza, A. S., Yusoff, M. S., Khalil, A. E., Ishola,
K. S. (2016). Characterization of active and closed landfill sites using 2D resistivity/IP
imaging: case studies in Penang, Malaysia. Environmental Earth Sciences, 75(4), 1–17.
https://doi.org/10.1007/s12665-015-5003-5

ABEM. (2018). Instruction Manual Instruction Manual-Terrameter SAS 4000 / SAS 1000. In
International Business.
https://www.academia.edu/39478165/Instruction_Manual_Terrameter_SAS_4000_SAS_1
000 (accessed on 10 July 2020)

Adamo, N., Al-Ansari, N., Sissakian, V., Laue, J., Knutsson, S. (2020). Geophysical Methods
and their Applications in Dam Safety Monitoring. Journal of Earth Sciences and
Geotechnical Engineering, 11(1), 291–345. https://doi.org/10.47260/jesge/1118

Adimalla, N., Li, P., Venkatayogi, S. (2018). Hydrogeochemical Evaluation of Groundwater


Quality for Drinking and Irrigation Purposes and Integrated Interpretation with Water
Quality Index Studies. Environmental Processes, 5(2), 363–383.
https://doi.org/10.1007/s40710-018-0297-4

Afonso, R. (1978). A Geologia de Moçambique – Notícia Explicativa da Carta Geológica de


Moçambique, 1:2000000. Maputo.

Ahmad, W., Alharthy, R. D., Zubair, M., Ahmed, M., Hameed, A., Rafique, S. (2021). Toxic and
heavy metals contamination assessment in soil and water to evaluate human health risk.
Scientific Reports, 11(1), 1–12. https://doi.org/10.1038/s41598-021-94616-4

Aizebeokhai, A. P. (2010). 2D and 3D geoelectrical resistivity imaging: Theory and field design.
Scientific Research and Essays, 5(23), 3592–3605

Ajibare, A.., Ogungbile, P., Ayeku, P. (2022). Evaluation of water pollution monitoring for heavy
metal contamination: A case study of Agodi Reservoir, Oyo State, Nigeria. Environmental
Monitoring and Assessment, 194(10). https://doi.org/10.1007/s10661-022-10326-y

Akhtar, J., Sana, A., Mohammed, S., Gajendran, T., Parmeswari, C. (2021). Evaluating the
groundwater potential of Wadi Al -Jizi , Sultanate of Oman , by integrating remote sensing
and GIS techniques. Environmental Science and Pollution Research.
https://doi.org/10.1007/s11356-021-17848-x

Akinbile, C. O., Yusoff, M. S. (2011). Environmental Impact of Leachate Pollution on


Groundwater Supplies in Akure , Nigeria. International Journal of Environmental Science
and Development, 2(1). https://doi.org/10.7763/IJESD.2011.V2.101

Al Manmi, D. A. M. A., Abdullah, T. O., Al-Jaf, P. M., Al-Ansari, N. (2019). Soil and groundwater
pollution assessment and delineation of intensity risk map in Sulaymaniyah City, NE of
Iraq. Water (Switzerland), 11(10). https://doi.org/10.3390/w11102158

Alekseenko, E., Roux, B., Kuznetsov, K. (2022). water Wind-Induced Resuspension and
Transport of Contaminated Sediment from the Rove Canal into the Etang De Berre ,
France. 1–25

Alexakis, D. E. (2020). Suburban areas in flames: Dispersion of potentially toxic elements from
burned vegetation and buildings. Estimation of the associated ecological and human
178

health risk. Environmental Research, 183, 109153.


https://doi.org/10.1016/j.envres.2020.109153

Alexakis, D. E. (2021). Multielement contamination of land in the margin of highways. Land,


10(3), 1–13. https://doi.org/10.3390/land10030230

Alfaro, M. R., Ugarte, O. M., Lima, L. H. V., Silva, J. R., da Silva, F. B. V., da Silva Lins, S. A.,
do Nascimento, C. W. A. (2022). Risk assessment of heavy metals in soils and edible
parts of vegetables grown on sites contaminated by an abandoned steel plant in Havana.
Environmental Geochemistry and Health, 44(1), 43–56. https://doi.org/10.1007/s10653-
021-01092-w

Alghamdi, A. G., Aly, A. A., Ibrahim, H. M. (2021). Assessing the environmental impacts of
municipal solid waste landfill leachate on groundwater and soil contamination in western
Saudi Arabia. Arabian Journal of Geosciences, 14(5). https://doi.org/10.1007/s12517-
021-06583-9

Alghamdi, M. A., Hassan, S. K., Al Sharif, M. Y., Khoder, M. I.,Harrison, R. M., 2023. Pollution
characteristics and human health risk of potentially toxic elements associated with
deposited dust of sporting walkways during physical activity. Atmospheric Pollution
Research, 14(1), 101649. https://doi.org/10.1016/j.apr.2023.101649

Ali, M. U., Liu, G., Yousaf, B., Abbas, Q., Ullah, H., Munir, M. A. M., Fu, B. (2017). Pollution
characteristics and human health risks of potentially (eco)toxic elements (PTEs) in road
dust from metropolitan area of Hefei, China. Chemosphere, 181, 111–121.
https://doi.org/10.1016/j.chemosphere.2017.04.061

Aliewi, A., Hadi, K., Bhandary, H., Al-Qallaf, H., Rashed, T., Abdulhadi, A., Al-Salem, S. M.
(2021). Investigation of landfill leachate pollution impact on shallow aquifers using
numerical simulation. Arabian Journal of Geosciences, 14(20).
https://doi.org/10.1007/s12517-021-08536-8

Alleoni, L. R. F., Iglesias, C. S. M., Mello, S. D. C., Camargo, O. A. de, Casagrande, J. C.,
Lavorenti, N. A. (2005). Soil attributes related to cadmium and copper adsorption in
tropical soils. Acta Scientiarum. Agronomy, 27(4).
https://doi.org/10.4025/actasciagron.v27i4.1348

Aller, L.; Bennett, T.; Lehr, J.H.; Petty, R.J.; Hackett, G. (1987). DRASTIC: A Standardized
Method for Evaluating Ground Water Pollution Potential Using Hydrogeologic Settings.
NWWA/EPA-600/2-87-035

Alloway, B.J. (2012). Heavy Metals in Soils: Trace Metals and Metalloids in Soils and Their
Bioavailability, Environmental Pollution (Vol. 22; Springer, ed.). Dordrecht.

Altaf, R., Altaf, S., Hussain, M., Shah, R. U., Ullah, R., Ullah, M. I., … Datta, R. (2021). Heavy
metal accumulation by roadside vegetation and implications for pollution control. PLoS
ONE, 16, 1–15. https://doi.org/10.1371/journal.pone.0249147

Alves, C. A., Evtyugina, M., Vicente, A. M. P., Vicente, E. D., Nunes, T. V, Silva, P. M. A.,
Querol, X. (2018). Science of the Total Environment Chemical pro fi ling of PM 10 from
urban road dust. Science of the Total Environment, 634, 41–51.
https://doi.org/10.1016/j.scitotenv.2018.03.338
179

Amato, F., Cassee, F. R., Denier van der Gon, H. A. C., Gehrig, R., Gustafsson, M., Hafner, W.,
Querol, X. (2014). Urban air quality: The challenge of traffic non-exhaust emissions.
Journal of Hazardous Materials, 275, 31–36.
https://doi.org/10.1016/j.jhazmat.2014.04.053

Andaloussi, K., Achtak, H., Nakhcha, C., Haboubi, K., Stitou, M. (2021). Assessment of soil
trace metal contamination of an uncontrolled landfill and its vicinity: the case of the city of
‘Targuist’ (Northern Morocco). Moroccan Journal of Chemistry, 9(3), 513–529.
https://doi.org/10.48317/IMIST.PRSM/morjchem-v9i2.23680

Anshumala, K., Shukla, J. P., Patel, S. S., Singh, A. (2021). Assessment of Groundwater
Vulnerability Zone in Mandideep Industrial Area using DRASTIC Model. Journal of the
Geological Society of India, 97(9), 1080–1086. https://doi.org/10.1007/s12594-021-1823-
y

Asfaw, D., Mengistu, D. (2020). Modeling megech watershed aquifer vulnerability to pollution
using modified DRASTIC model for sustainable groundwater management, Northwestern
Ethiopia. Groundwater for Sustainable Development, 11.
https://doi.org/https://doi.org/10.1016/j.gsd.2020.100375

Asowata, I. T. (2021). Geophagic clay around Uteh-Uzalla near Benin : mineral and trace
elements compositions and possible health implications. SN Applied Sciences, 3(5), 1–
12. https://doi.org/10.1007/s42452-021-04565-w

Ashraf, M., Zeshan, M., Hafeez, S., Hussain, R., Qadir, A., Majid, M., Ahmad, S. R. (2022).
Temporal variation in leachate composition of a newly constructed landfill site in Lahore in
context to environmental pollution and risks. Environmental Science and Pollution
Research, (0123456789). https://doi.org/10.1007/s11356-022-18646-9

Ashraf, M., Zeshan, M., Hafeez, S., Hussain, R., Qadir, A., Majid, M., Ahmad, S. R. (2022).
Temporal variation in leachate composition of a newly constructed landfill site in Lahore in
context to environmental pollution and risks. Environmental Science and Pollution
Research, (0123456789). https://doi.org/10.1007/s11356-022-18646-9

Ashraf, M., Zeshan, M., Hafeez, S., Hussain, R., Qadir, A., Majid, M., Ahmad, S. R. (2022).
Temporal variation in leachate composition of a newly constructed landfill site in Lahore in
context to environmental pollution and risks. Environmental Science and Pollution
Research, (0123456789). https://doi.org/10.1007/s11356-022-18646-9

Astuti, R. D. P., Mallongi, A., Rauf, A. U. (2021). Natural enrichment of chromium and nickel in
the soil surrounds the karst watershed. 7(3), 383–400.
https://doi.org/10.22034/gjesm.2021.03.05

ATSDR (Agency for Toxic Substances and Disease Registry). (2002). Zinc.
https://doi.org/10.1201/9781420061888_ch97

Audu, P., Wuana, R. A. (2021). Evaluating the Levels and Human Health Risks of Heavy Metals
in Soils around Onne Landfill, Rivers State, Nigeria. International Research Journal of
Pure and Applied Chemistry, 22(3), 43–59.https://doi.org/10.9734/irjpac/2021/v22i330395

Aydi, A., Mhimdi, A., Hamdi, I., Touaylia, S., Sdiri, A. (2020). Application of electrical resistivity
tomography and hydro-chemical analysis for an integrated environmental assessment.
180

Environmental Nanotechnology, Monitoring and Management, 14, 100351.


https://doi.org/10.1016/j.enmm.2020.100351

Azeez, J. O., Hassan, O. A., Egunjobi, P. O. (2011). Soil contamination at dumpsites:


Implication of soil heavy metals distribution in municipal solid waste disposal system: A
case study of Abeokuta, southwestern Nigeria. Soil and Sediment Contamination, 20(4),
370–386. https://doi.org/10.1080/15320383.2011.571312

Baawain, M. S., Al-Futaisi, A. M., Ebrahimi, A., Omidvarborna, H. (2018). Characterizing


leachate contamination in a landfill site using Time Domain Electromagnetic (TDEM)
imaging. Journal of Applied Geophysics, 151, 73–81.
https://doi.org/10.1016/j.jappgeo.2018.02.002

Bai, B., Long, F., Rao, D., Xu, T. (2017). The effect of temperature on the seepage transport of
suspended particles in a porous medium. Hydrological Processes, 31(2), 382–393.
https://doi.org/10.1002/hyp.11034

Balasooriya, S., Diyabalanage, S., Yatigammana, S. K., Ileperuma, O. A., Chandrajith, R.


(2021). Major and trace elements in rice paddy soils in Sri Lanka with special emphasis
on regions with endemic chronic kidney disease of undetermined origin. Environmental
Geochemistry and Health, 4. https://doi.org/10.1007/s10653-021-01036-4

Bhardwaj, S., Soni, R., Gupta, S. K., Shukla, D. P. (2020). Mercury, arsenic, lead and cadmium
in waters of the Singrauli coal mining and power plants industrial zone, Central East India.
Environmental Monitoring and Assessment, 192(4). https://doi.org/10.1007/s10661-020-
8225-2

Bandeira, S. and Paula, J. (2014). The Maputo Bay Ecosystem. WIOMSA. isbn: 978-9987-
9559-3-0

Barry, A. A., Yameogo, S., Ayach, M., Jabrane, M., Tiouiouine, A., Nakolendousse, S., …
Mohsine, I. (2021). Basement Terrain in Ouagadougou, Burkina Faso , Using Self-
Potential Geophysical Technique.

Barry, A. A., Yameogo, S., Ayach, M., Jabrane, M., Tiouiouine, A., Nakolendousse, S., Mohsine,
I. (2021). Basement Terrain in Ouagadougou, Burkina Faso , Using Self-Potential
Geophysical Technique.

Barry, A. A., Yameogo, S., Ayach, M., Jabrane, M., Tiouiouine, A., Nakolendousse, S., …
Mohsine, I. (2021). Basement Terrain in Ouagadougou, Burkina Faso , Using Self-
Potential Geophysical Technique.

Baruya, P., Kessels, J. (2013). Coal prospects in Botswana, Zimbabwe and Namibia. IEA Clean
Coal Centre Abstract, (December). https://doi.org/10.13140/RG.2.2.36546.48325

Begum, W., Summi R., Soujanya B., Sudip B., Monohar H. M., Ajaya, B., Bidyut, S. (2022). A
Comprehensive Review on the Sources, Essentiality and Toxicological Profile of Nickel.
RSC Advances 12 (15): 9139–53. https://doi.org/10.1039/d2ra00378c.

Beji, A., Deboudt, K., Khardi, S., Muresan, B., Flament, P., Fourmentin, M., Lumière, L. (2020).
Non-exhaust particle emissions under various driving conditions: Implications for
sustainable mobility. Transportation Research Part D: Transport and Environment, 81,
102290. https://doi.org/10.1016/j.trd.2020.102290
181

Bernardo, B. J. (2021). A influência da dinâmica urbana e a ocupação de áreas inundáveis no


bairro Magoanine-A (Moçambique): uma reflexão sobre o zoneamento ambiental. Revista
Internacional Em Língua Portuguesa, (35), 61–68. https://doi.org/10.31492/2184-
2043.RILP2018.35/pp.

Bernardo, B., Candeias, C., Rocha, F. (2022a). Application of Geophysics in geo-environmental


diagnosis on the surroundings of the Hulene-B waste dump, Maputo, Mozambique.
Journal of African Earth Sciences, 185, 104415.
https://doi.org/10.1016/j.jafrearsci.2021.104415

Bernardo, B., Candeias, C., Rocha, F. (2022b). Characterization of the Dynamics of Leachate
Contamination Plumes in the Surroundings of the Hulene-B Waste Dump in Maputo,
Mozambique. Environments - MDPI, 9(2). https://doi.org/10.3390/environments9020019

Bernardo, B., Candeias, C., Rocha, F. (2022c). Soil properties and environmental risk
assessment of soils in the surrounding area of Hulene-B waste dump, Maputo
(Mozambique). Environ Earth Sci 81, 542 https://doi.org/10.1007/s12665-022-10672-7

Bernardo, B., Candeias, C., Rocha, F. (2022d). Soil Risk Assessment in the Surrounding Area
of Hulene-B Waste Dump, Maputo (Mozambique). Geosciences, 12(290).
https://doi.org/https://doi.org/10.3390/ geosciences12080290

Bernardo, B.; Candeias, C.; Rocha, F. (2022e). Integration of Electrical Resistivity and Modified
DRASTIC Model to Assess Groundwater Vulnerability in the Surrounding. Water,
14(1746). https://doi.org/https://doi.org/10.3390/ w14111746

Bernardo, B.; Candeias, C.; Rocha, F. (2023). The Contribution of the Hulene-B Waste Dump
(Maputo, Mozambique) to the Contamination of Rhizosphere Soils, Edible Plants, Stream
Waters, and Groundwaters. Environments , 10, 45.
https://doi.org/10.3390/environments10030045

Bichet, V., Grisey, E., Aleya, L. (2016). Spatial characterization of leachate plume using
electrical resistivity tomography in a landfill composed of old and new cells (Belfort,
France). Engineering Geology, 211, 61–73. https://doi.org/10.1016/j.enggeo.2016.06.026

Binh, N. T. L., Hoang, N. T., Truc, N. T. T., Khang, V. D., Le, H. A. (2021). Estimating the
Possibility of Lead Contamination in Soil Surface due to Lead Deposition in Atmosphere.
Journal of Nanomaterials, 2021. https://doi.org/10.1155/2021/5586951

Biosca, B., Arévalo-Lomas, L., Izquierdo-Díaz, M., Díaz-Curiel, J. (2021). Detection of


chlorinated contaminants coming from the manufacture of lindane in a surface detritic
aquifer by electrical resistivity tomography. Journal of Applied Geophysics, 191
https://doi.org/10.1016/j.jappgeo.2021.104358

Bisht, L., Gupta, V., Singh, A., Gautam, A. S., Gautam, S. (2022). Heavy metal concentration
and its distribution analysis in urban road dust: A case study from most populated city of
Indian state of Uttarakhand. Spatial and Spatio-Temporal Epidemiology, 40, 100470.
https://doi.org/10.1016/j.sste.2021.100470

Bisht, L., Gupta, V., Singh, A., Gautam, A. S., Gautam, S. (2022). Heavy metal concentration
and its distribution analysis in urban road dust: A case study from most populated city of
182

Indian state of Uttarakhand. Spatial and Spatio-Temporal Epidemiology, 40,


100470https://doi.org/10.1016/j.sste.2021.100470

Blarasin, M., Matiatos, I., Cabrera, A., Lutri, V., Giacobone, D., Quinodoz, F. B. (2021).
Characterization of groundwater dynamics and contamination in an unconfined aquifer
using isotope techniques to evaluate domestic supply in an urban area. Journal of South
American Earth Sciences, 110(103360). https://doi.org/10.1016/j.jsames.2021.103360

Boumaiza, L., Walter, J., Chesnaux, R., Brindha, K., Elango, L. (2021). An operational
methodology for determining relevant DRASTIC factors and their relative weights in the
assessment of aquifer vulnerability to contamination. Environmental Earth Sciences,
80(7), 1–19. https://doi.org/10.1007/s12665-021-09575-w

Bradl, H. B. (2004). Adsorption of heavy metal ions on soils and soils constituents. Journal of
Colloid and Interface Science, 277(1), 1–18. https://doi.org/10.1016/j.jcis.2004.04.005

Brahmi, S., Baali, F., Hadji, R., Brahmi, S., Hamad, A., Rahal, O., Hamed, Y. (2021).
Assessment of groundwater and soil pollution by leachate using electrical resistivity and
induced polarization imaging survey, case of Tebessa municipal landfill, NE Algeria.
Arabian Journal of Geosciences, 14(4). https://doi.org/10.1007/s12517-021-06571-z

Breus, D., Yevtushenko, O. (2022). Modeling of Trace Elements and Heavy Metals Content in
the Steppe Soils of Ukraine. Journal of Ecological Engineering Journal, 23(2), 159–165

Buque, L. I. B., Ribeiro, H. (2015). Panorama da coleta seletiva com catadores no município de
Maputo, Moçambique: Desafios e perspectivas. Saude e Sociedade, 24(1), 298–307.
https://doi.org/10.1590/S0104-12902015000100023

Campos, C. (2010). Soil attributes and risk of leaching of heavy metals in tropical soils.
Ambiência, 6(3), 547–565.
https://revistas.unicentro.br/index.php/ambiencia/article/view/591

Campos, C. (2010). Soil attributes and risk of leaching of heavy metals in tropical soils.
Ambiência, 6(3), 547–565.
https://revistas.unicentro.br/index.php/ambiencia/article/view/591

Canadian Council of Minister of the Environment. (1999). Development of a preliminary


database of digestate chemistry, heavy metal and pathogen content to assist in Alberta
regulation compliance. Canadian Environmnetal Quality Guidelines, 10.
http://energy.alberta.ca/BioEnergy/pdfs/HeavyMetalReport.pdf%5Cnhttp://ceqg-
rcqe.ccme.ca/download/en/269

Candeias, C., da Silva, E. F., Ávila, P. F., Teixeira, J. P. (2014). Identifying sources and
assessing potential risk of exposure to heavy metals and hazardous materials in mining
areas: The case study of panasqueira mine (Central Portugal) as an example.
Geosciences (Switzerland), 4(4), 240–268. https://doi.org/10.3390/geosciences404024

Candeias, C.; Vicente, E.; Tomé, M.; Rocha, F.; Ávila, P.; Célia, A. (2020). Geochemical,
Mineralogical and Morphological Characterization of Road Dust and Associated Health
Risks. Int. J. Environ. Res. Public Health. 17, 1563.
https://doi.org/10.3390/ijerph17051563.
183

Candeias, C., Paula, F. Á., Ferreira, E. (2021a). Metal (Loids) Bioaccessibility in Road Dust
from the Surrounding Villages of an Active Mine. Atmosphere, 12, 685.
https://doi.org/10.3390/atmos12060685

Candeias, C., Ávila, P. F., Sequeira, C., Manuel, A., Rocha, F. (2022). Potentially toxic elements
dynamics in the soil rhizospheric-plant system in the active volcano of Fogo (Cape Verde)
and interactions with human health. Catena, 209(2021).
https://doi.org/10.1016/j.catena.2021.105843
Catuneanu, O., Wopfner, H., Eriksson, P. G., Cairncross, B., Rubidge, B. S., Smith, R. M. H.,
Hancox, P. J. (2005). The Karoo basins of south-central Africa. Journal of African Earth
Sciences, 43(1–3), 211–253. https://doi.org/10.1016/j.jafrearsci.2005.07.007
CCME (Canadian Council of Ministers of the Environment). (2015). Canadian Soil Quality
Guidelines for the Protection of Environmental and Human Health - Nickel.
https://ccme.ca/en/res/nickel-canadian-soil-quality-guidelines-for-the-protection-of-
environmental-and-human-health-en.pdf

CCME (Canadian Council of Ministers of the Environment). (1997). Canadian Soil Quality
Guidelines for the Protection of Environmental and Human Health-Chromium.
https://ccme.ca/en/res/chromium-canadian-soil-quality-guidelines-for-the-protection-of-
environmental-and-human-health-en.pdf

CCME (Canadian Council of Ministers of the Environment). (2018a). Canadian Soil Quality
Guidelines for the Protection of Environmental and Human Health. Excerpt from
Publication No. 1299; ISBN 1-896997-34-1

CCME (Canadian Council of Ministers of the Environment). (2018b). Scientific criteria document
for the development of the canadian soil quality guideline for Zinc Protection of
Environmental and Human Health. https://www.ccme.ca/fr/res/2018-zinc-csqg-scd-1577-
en.pdf

Ceballos, E., Dubny, S., Othax, N., Zabala, M. E., Peluso, F. (2021). Assessment of Human
Health Risk of Chromium and Nitrate Pollution in Groundwater and Soil of the Matanza-
Riachuelo River Basin, Argentina. Exposure and Health, 13(3), 323–336.
https://doi.org/10.1007/s12403-021-00386-9

Cendón, D. I., Haldorsen, S., Chen, J., Hankin, S., Nogueira, G. E. H., Momade, F., Stigter, T.
Y. (2020). Hydrogeochemical aquifer characterization and its implication for groundwater
development in the Maputo district, Mozambique. Quaternary International, 547, 113–
126. https://doi.org/10.1016/j.quaint.2019.06.024

Chaaou, A., Chikhaoui, M., Naimi, M., El Miad, A. K., Achemrk, A., Seif-Ennasr, M., El Harche,
S. (2022). Mapping soil salinity risk using the approach of soil salinity index and land
cover: a case study from Tadla plain, Morocco. Arabian Journal of Geosciences, 15(8).
https://doi.org/10.1007/s12517-022-10009-5

Chen, H., Wang, L., Hu, B., Xu, J., Liu, X. (2022). Potential driving forces and probabilistic
health risks of heavy metal accumulation in the soils from an e-waste area, southeast
China. Chemosphere, 289, 133182. https://doi.org/10.1016/j.chemosphere.2021.133182

Chen, Y., Qin, X., Huang, Q., Gan, F., Han, K., Zheng, Z., Meng, Y. (2018). Anomalous
spontaneous electrical potential characteristics of epi-karst in the Longrui Depression,
184

Southern Guangxi Province, China. Environmental Earth Sciences, 77(19), 1–9.


https://doi.org/10.1007/s12665-018-7839-y

Chetri, J. K., Reddy, K. R. (2021). Advancements in Municipal Solid Waste Landfill Cover
System: A Review. Journal of the Indian Institute of Science, 101(4), 557–588.
https://doi.org/10.1007/s41745-021-00229-1

CIAT(Center for Tropical Agriculture). (2017). Climate-Smart Agriculture in Mozambique.


Climate-Smart Agriculture in Mozambique.
https://climateknowledgeportal.worldbank.org/sites/default/files/2019-06/CSA-in-
Mozambique.pdf

CMCM (Conselho Municipal da Cidade de Maputo). (2020). Quadro da Política de


Reassentamento - QPR Novembro de 2020.
https://documents1.worldbank.org/curated/en/964181607106676324/pdf/Revised-
Resettlement-Framework-Maputo-Urban-Transformation-Project-P171449.pdf

Coker, E., Kizito, S. (2018). A Narrative Review on the Human Health Effects of Ambient Air
Pollution in Sub-Saharan Africa : An Urgent Need for Health Effects Studies. Int. J.
Environ. Res. Public Health 2018, 15(427). https://doi.org/10.3390/ijerph15030427

Corradini, E., Dreibrodt, S., Erkul, E., Groß, D., Lübke, H., Panning, D., … Rabbel, W. (2020).
Understanding wetlands stratigraphy: Geophysics and soil parameters for investigating
ancient basin development at lake duvensee. Geosciences (Switzerland), 10(8), 1–35.
https://doi.org/10.3390/geosciences10080314

Dakheel Almaliki, A. J., Bashir, M. J. K., Llamas Borrajo, J. F. (2022). Appraisal of groundwater
contamination from surface spills of fluids associated with hydraulic fracturing operations.
Science of the Total Environment, 815, 152949.
https://doi.org/10.1016/j.scitotenv.2022.152949

Davoudi, M., Esmaili-Sari, A., Bahramifar, N. (2022). Spatio-temporal variation and risk
assessment of polycyclic aromatic hydrocarbons (PAHs) in surface dust of Qom
metropolis, Iran. Environ Sci Pollut Res 28, 9276-9289. https://doi.org/10.1007/s11356-
020-08863-5
Dehghani, S., Moore, F., Keshavarzi, B., Hale, A. (2017). Health risk implications of potentially
toxic metals in street dust and surface soil of Tehran, Iran. Ecotoxicology and
Environmental Safety, 136, 92–103. https://doi.org/10.1016/j.ecoenv.2016.10.037

Demayo, A., Taylor, M. C., Taylor, K. W., Hodson, P. V. (1982). Toxic Effects of Lead and
Lead Compounds on Human Health, Aquatic Life, Wildlife Plants, and Livestock. Critical
Reviews in Environmental Control, 12(4), 257–305.
https://doi.org/10.1080/10643388209381698

Diniz, M. A., Bandeira, S., Martins, E. S. (2013). Flora e Vegetação Da Província De Maputo :
Sua Apropriação Pelas Populações. Atas Do Congresso Internacional Saber Tropical Em
Moçambique: História, Memória E Ciência, 24–26.

Dirks, P. H. G. M., Blenkinsop, Tom G, Jelsma, H. A. (2003). The Geological evolution of Africa.
Geology, IV. http://eolss.net/Sample-Chapters/C01/E6-15-07-02.pdf
185

Dregulo, A. M., Bobylev, N. G. (2020). Heavy metals and arsenic soil contamination resulting
from wastewater sludge urban landfill disposal. Polish Journal of Environmental Studies,
30(1), 81–89. https://doi.org/10.15244/pjoes/121989

El Fadili, H., Ben Ali, M., Touach, N., El Mahi, M., MostaphaLotfi, E. (2022). Ecotoxicological
and Pre-remedial risk assessment of heavy metals in municipal solid wastes dumpsite
impacted soil in Morocco. Environmental Nanotechnology, Monitoring Management, 17,
100640. https://doi.org/10.1016/j.enmm.2021.100640

El Mouine, Y., El Hamdi, A., Morarech, M., Kacimi, I., Touzani, M., Mohsine, I., Dakak, H.
(2021). Article landfill pollution plume survey in the moroccantadla using spontaneous
potential. Water (Switzerland), 13(7), 1–11. https://doi.org/10.3390/w13070910

Ellegard, A. (1997). Household Energy, Air Pollution and Health in Maputo (Stockholm
Environment Institute). isbn:918871411X

El Naqa, A. (2004). Aquifer vulnerability assessment using the DRASTIC model at Russeifa
landfill, northeast Jordan. Environmental Geology, 47(1), 51–62.
https://doi.org/10.1007/s00254-004-1126-9

Ersoy, A., Gültekin, F. (2013). DRASTIC-based methodology for assessing groundwater


vulnerability in the Gümüşhacıköy and Merzifon basin (Amasya, Turkey). Res. SJ, 17(1),
33–40.

Essien, J. P., Ikpe, D. I., Inam, E. D., Okon, A. O., Ebong, G. A., Benson, N. U. (2022).
Occurrence and spatial distribution of heavy metals in landfill leachates and impacted
freshwater ecosystem: An environmental and human health threat. PLoS ONE, 17, 1–18.
https://doi.org/10.1371/journal.pone.0263279

EU (European Union). (2004). Heavy metals and organic compounds from wastes used as
organic fertilisers. Final report for ENV. A. 2./ETU/2001/0024. Final Report for ENV. A.
2./ETU/2001/0024.
https://ec.europa.eu/environment/pdf/waste/compost/hm_finalreport.pdf

European Environment Agency. (2020). Air quality in Europe — 2020 report.


https://doi.org/doi:10.2800/78665

FAO. 2020. Fruit and vegetables – your dietary essentials. The International Year of Fruits and
Vegetables, 2021, background paper. Rome. https://doi.org/10.4060/cb2395enFatoba, J.
O., Eluwole, A. B., Sanuade, O. A., Hammed, O. S., Igboama, W. N., Amosun, J. O.
(2021). Geophysical and geochemical assessments of the environmental impact of
Abule-Egba landfill, southwestern Nigeria. Modeling Earth Systems and Environment,
7(2), 695–701. https://doi.org/10.1007/s40808-020-00991-8

Fayiga, A. O., Ipinmoroti, M. O., Chirenje, T. (2018). Environmental pollution in Africa. In


Environment, Development and Sustainability (Vol. 20). https://doi.org/10.1007/s10668-
016-9894-4

Feng, S. J., Wu, S. J., Fu, W. D., Zheng, Q. T., Zhang, X. L. (2021). Slope stability analysis of a
landfill subjected to leachate recirculation and aeration considering bio-hydro coupled
processes. Geoenvironmental Disasters, 8(1). https://doi.org/10.1186/s40677-021-00201-
2
186

Ferrão, D. A. G. (2006). Evaluation of removal and disposal of solid waste in Maputo City,
Mozambique. Master’s thesis, University of Cape Town.
https://open.uct.ac.za/bitstream/handle/11427/4851/thesis_sci_2006_ferrao_d_a_g.pdf?s
equence=1

Fiala, M., Hwang, H. M. (2021). Influence of Highway Pavement on Metals in Road Dust: a
Case Study in Houston, Texas. Water, Air, and Soil Pollution, 232(5).
https://doi.org/10.1007/s11270-021-05139-7

Friberg, J., Abera, A., Friberg, J., Isaxon, C., Jerrett, M., Malmqvist, E., … Vargas, A. M. (2021).
Air Quality in Africa: Public Health Implications. Annual Review Of Public
Health.https://doi.org/10.1146/annurev-publhealth100119-113802

Gascon, M., Rojas-rueda, D., Torrico, S. (2016). Urban Policies and Health in Developing
Countries: The Case of Maputo (Mozambique) and Cochabamba (Bolivia). Public Health
Open J. 2016;, 1(2). https://doi.org/10.17140/PHOJ-1-106

Gautam, P., Kumar, S. (2021). Characterisation of hazardous waste landfill leachate and its
reliance on landfill age and seasonal variation: A statistical approach. Journal of
Environmental Chemical Engineering, 9(4), 105496.
https://doi.org/10.1016/j.jece.2021.105496

Gemail, K. S., Alfy, M. El, Ghoneim, M. F., Shishtawy, A. M., El-Bary, M. A. (2017). Comparison
of DRASTIC and DC resistivity modeling for assessing aquifer vulnerability in the central
Nile Delta, Egypt. Environmental Earth Sciences, 76(9), 1–17.
https://doi.org/10.1007/s12665-017-6688-4

Gebeyehu, H. R., Bayissa, L. D. (2020). Levels of heavy metals in soil and vegetables and
associated health risks in Mojo area, Ethiopia. PLoS ONE, 15(1), 1–22.
https://doi.org/10.1371/journal.pone.0227883

George, N. J. (2021). Geo-electrically and hydrogeologically derived vulnerability assessments


of aquifer resources in the hinterland of parts of Akwa Ibom State, Nigeria. Solid Earth
Sciences, 6(2), 70–79. https://doi.org/10.1016/j.sesci.2021.04.002

Ghosh, R., Sutradhar, S., Mondal, P., Das, N. (2021). Application of DRASTIC model for
assessing groundwater vulnerability: a study on Birbhum district, West Bengal, India.
Modeling Earth Systems and Environment, 7(2), 1225–1239.
https://doi.org/10.1007/s40808-020-01047-7

Ghosh, S., Sharma, A., Talukder, G. (1992). Zirconium An Abnormal Trace Element in Biology.
35.

Giang, N. V., Kochanek, K., Vu, N. T., Duan, N. B. (2018). Landfill leachate assessment by
hydrological and geophysical data: case study NamSon, Hanoi, Vietnam. Journal of
Material Cycles and Waste Management, 20(3), 1648–1662.
https://doi.org/10.1007/s10163-018-0732-7

Grigoratos, T., Martini, G. (2015). Brake wear particle emissions: a review. Environmental
Science and Pollution Research, 22(4), 2491–2504. https://doi.org/10.1007/s11356-014-
3696-8

Githaiga, K. B., Njuguna, S. M., Gituru, R. W., Yan, X. (2021). Water quality assessment,
multivariate analysis and human health risks of heavy metals in eight major lakes in
187

Kenya. Journal of Environmental Management, 297.


https://doi.org/10.1016/j.jenvman.2021.113410

Guadie, A., Yesigat, A., Gatew, S., Worku, A., Liu, W., Ajibade, F. O., Wang, A. (2021).
Evaluating the health risks of heavy metals from vegetables grown on soil irrigated with
untreated and treated wastewater in Arba Minch, Ethiopia. In Science of the Total
Environment, 761. https://doi.org/10.1016/j.scitotenv.2020.143302

Guerra-García, J. M., Navarro-Barranco, C., Martínez-Laiz, G., Moreira, J., Giráldez, I., Morales,
E., … Ros, M. (2021). Assessing environmental pollution levels in marinas. Science of the
Total Environment, 762, 144169. https://doi.org/10.1016/j.scitotenv.2020.144169

Gujre, N., Mitra, S., Soni, A., Agnihotri, R., Rangan, L., Rene, E. R., Sharma, M. P. (2021).
Speciation, contamination, ecological and human health risks assessment of heavy
metals in soils dumped with municipal solid wastes. Chemosphere, 262, 128013.
https://doi.org/10.1016/j.chemosphere.2020.128013

Gunawardana, C., Goonetilleke, A., Egodawatta, P., Dawes, L., Kokot, S. (2012). Chemosphere
Source characterisation of road dust based on chemical and mineralogical composition.
Chemosphere, 87(2), 163–170. https://doi.org/10.1016/j.chemosphere.2011.12.012

Guo, W., Liu, X., Liu, Z., Li, G. (2010). Pollution and potential ecological risk evaluation of
heavy metals in the sediments around Dongjiang Harbor, Tianjin. Procedia Environmental
Sciences, 2(5), 729–736. https://doi.org/10.1016/j.proenv.2010.10.084

Gupta, N., Yadav, K. K., Kumar, V., Krishnan, S., Kumar, S., Nejad, Z. D., … Alam, J. (2021).
Evaluating heavy metals contamination in soil and vegetables in the region of North India:
Levels, transfer and potential human health risk analysis. Environmental Toxicology and
Pharmacology, 82, 103563. https://doi.org/10.1016/j.etap.2020.103563

Halder, S., Agüero, J., Dolle, P., Fernández, E., Schmidt, C., Yang, M. (2018). Perspectives of
Urban Agriculture in Maputo and Cape Town. https://www.sle-
berlin.de/files/sle/publikationen/S%20275-Maputo-Internet-Klein.pdf

Hamanaka, R. B., Mutlu, G. M. (2018). Particulate Matter Air Pollution: Effects on the
Cardiovascular System. Frontiers in Endocrinology, 9, 1–15.
https://doi.org/10.3389/fendo.2018.00680

Han, N. M. im M., Latif, M. T., Othman, M., Dominick, D., Mohamad, N., Juahir, H., Tahir, N. M.
(2014). Composition of selected heavy metals in road dust from Kuala Lumpur city centre.
Environmental Earth Sciences, 72(3), 849–859. https://doi.org/10.1007/s12665-013-3008-
5

Harjito, Suntoro, Gunawan, T., Maskuri, M. (2018). Underground leachate distribution based on
electrical resistivity tomography in Piyungan landfill, Bantul. Indonesian Journal of
Geography, 50(1), 34–40. https://doi.org/10.22146/ijg.18315

Hasan, M. A., Ahmad, S., Mohammed, T. (2021). Groundwater Contamination by Hazardous


Wastes. Arabian Journal for Science and Engineering, 46(5), 4191–4212.
https://doi.org/10.1007/s13369-021-05452-7

Helene, L. P. I., Moreira, C. A. (2021). Analysis of Leachate Generation Dynamics in a Closed


Municipal Solid Waste Landfill by Means of Geophysical Data (DC Resistivity and Self-
188

Potential Methods). Pure and Applied Geophysics, 178(4), 1355–1367.


https://doi.org/10.1007/s00024-021-02700-

Heo, S., Kim, D. Y., Kwoun, Y., Lee, T. J., Jo, Y. M. (2021). Characterization and source
identification of fine dust in Seoul elementary school classrooms. Journal of Hazardous
Materials, 414. https://doi.org/10.1016/j.jhazmat.2021.125531

Hoai, S. T., Lan, H. N., Viet, N. T. T., Hoang, G. N., Kawamoto, K. (2021). Characterizing
seasonal variation in landfill leachate using leachate pollution index (LPI) at nam son solid
waste landfill in Hanoi, Vietnam. Environments - MDPI, 8(3), 1–11.
https://doi.org/10.3390/environments8030017

Hoang, H. G., Chiang, C. F., Lin, C., Wu, C. Y., Lee, C. W., Cheruiyot, N. K., … Bui, X. T.
(2021). Human health risk simulation and assessment of heavy metal contamination in a
river affected by industrial activities. Environmental Pollution, 285, 117414.
https://doi.org/10.1016/j.envpol.2021.117414

Hodson, M. E. (2002). Experimental evidence for mobility of Zr and other trace elements in
soils. Geochimica et Cosmochimica Acta, 66(5), 819–828. https://doi.org/10.1016/S0016-
7037(01)00803-1

Hosseini, M., Saremi, A. (2018). Assessment and Estimating Groundwater Vulnerability to


Pollution Using a Modified DRASTIC and GODS Models (Case Study: Malayer Plain of
Iran). Civil Engineering Journal, 4(2), 433. https://doi.org/10.28991/cej-0309103

Horwell, C.J., Williamson, B.J., Donaldson, Blond, J.SL., Damby, D.E., Bowen, L. (2012). The
structure of volcanic cristobalite in relation to its toxicity; relevance for the variable
crystalline silica hazard. Part Fibre Toxicol 9, 44. https://doi.org/10.1186/1743-8977-9-44.

Hussein, M., Yoneda, K., Mohd-Zaki, Z., Amir, A., Othman, N. (2021). Heavy metals in leachate,
impacted soils and natural soils of different landfills in Malaysia: An alarming threat.
Chemosphere, 267. https://doi.org/10.1016/j.chemosphere.2020.128874

Iddrisu, T. I., Debrah, K. D. (2021). Consequences of Poor Landfill Management on the People
of Gbalahi in the Sagnarigu Municipality of Northern Ghana. Journal of Geoscience and
Environment Protection, 09(08), 211–224. https://doi.org/10.4236/gep.2021.98014

Igboama, W. N., Hammed, O. S., Fatoba, J. O., Aroyehun, M. T., Ehiabhili, J. C. (2022). Review
article on impact of groundwater contamination due to dumpsites using geophysical and
physiochemical methods. Applied Water Science, 12(6), 1–14.
https://doi.org/10.1007/s13201-022-01653-z

INE(Instituto Nacional de Estatistica). (2020). Boletim de Estatísticas Demográficas e Sociais,


Maputo Cidade 2019. http://www.ine.gov.mz/estatisticas/estatisticas-demograficas-e-
indicadores-sociais/boletim-de-indicadores-demograficos-22-de-julho-de-
2020.pdf/at_download/file

NIA (Instituto Nacional de Investigação Agronómica). (1993). Os solos das Provincias de


Maputo e Gaza- Explicação dos mapas dos solos (Escala 1/50.000).
https://docplayer.com.br/44532401-Os-solos-das-provincias-de-maputo-e-gaza-
explicagoes-dos-mapas-de-solos-i-escala-1-50-000-inia-departamento-terra-e-agua-
versao-preliminair.html
189

Ihedioha, J. N., Ukoha, P. O., Ekere, N. R. (2017). Ecological and human health risk
assessment of heavy metal contamination in soil of a municipal solid waste dump in Uyo,
Nigeria. Environmental Geochemistry and Health, 39(3), 497–515.
https://doi.org/10.1007/s10653-016-9830-4

IQAir. (2020). World Air Quality Report Region City PM2.5 Ranking. Goldach, Switz.

Isa, B. K., Amina, S. B., Aminu, U., Sabo, Y. (2015). Health risk assessment of heavy metals in
water, air, soil and fish. African Journal of Pure and Applied Chemistry, 9(11), 204–210.
https://doi.org/10.5897/ajpac2015.0654

Isiuku, B. O., Enyoh, C. E. (2020). Pollution and health risks assessment of nitrate and
phosphate concentrations in water bodies in South Eastern , Nigeria. Environmental
Advances, 2, 100018. https://doi.org/10.1016/j.envadv.2020.100018

Islamd, M. S., Idris, A. M., Islam, A. R. M. T., Phoungthong, K., Ali, M. M., Kabir, M. H. (2021b).
Geochemical variation and contamination level of potentially toxic elements in land-uses
urban soils. International Journal of Environmental Analytical Chemistry, 1–18.
https://doi.org/10.1080/03067319.2021.1977286

Islami, N., Irianti, M., Fakhruddin, F., Azhar, A., Nor, M. (2020). Application of geoelectrical
resistivity method for the assessment of shallow aquifer quality in landfill areas.
Environmental Monitoring and Assessment, 192(9). https://doi.org/10.1007/s10661-020-
08564-z

Itai, T., Otsuka, M., Asante, K. A., Muto, M., Opoku-Ankomah, Y., Ansa-Asare, O. D., Tanabe,
S. (2014). Variation and distribution of metals and metalloids in soil/ash mixtures from
Agbogbloshie e-waste recycling site in Accra, Ghana. Science of the Total Environment,
470–471, https://doi.org/10.1016/j.scitotenv.2013.10.037

Jackson, M. J. (1989). Physiology of Zinc: General Aspects. 1–14. https://doi.org/10.1007/978-


1-4471-3879-2_1

Jayawickreme, D. H., Jobbágy, E. G., Jackson, R. B. (2014). Geophysical subsurface imaging


for ecological applications. New Phytologist, 201(4), 1170–1175.
https://doi.org/10.1111/nph.12619

Johansson, C., Norman, M., Burman, L. (2009). Road traffic emission factors for heavy metals.
Atmospheric Environment, 43(31), 4681–4688.
https://doi.org/10.1016/j.atmosenv.2008.10.024

Jones, J.V., Piatak, N.M., Bedinger, G. M. (2017). Zirconium and Hafnium. In U.S. Geological
Survey Professional Paper 1802 (Vol. V1–V26, p. Reston, Virginia).
https://doi.org/https://doi.org/ 10.3133/pp1802V

Joubert, B. R., Mantooth, S. N., Mcallister, K. A. (2020). Environmental Health Research in


Africa : Important Progress and Promising Opportunities. 10, 1–29.
https://doi.org/10.3389/fgene.2019.01166

Jovanović, B., Jovanović, N., Cvetković, V. J., Matić, S., Stanić, S., Whitley, E. M., Mitrović, T.
L. (2018). The effects of a human food additive , titanium dioxide nanoparticles E171 , on
Drosophila melanogaster - a 20 generation dietary exposure experiment. 1–12.
https://doi.org/10.1038/s41598-018-36174-w
190

Kalisa, E., Archer, S., Nagato, E., Bizuru, E., Lee, K., Tang, N., Lacap-bugler, D. (2019).
Chemical and Biological Components of Urban Aerosols in Africa : Current Status and
Knowledge Gaps. Int. J. Environ. Res. Public Health, 9(16), 941.
https://doi.org/10.3390/ijerph16060941

Kampunzu, A. B., Popoff, M. (1991). Distribution of the Main Phanerozoic African Rifts and
Associated Magmatism: Introductory Notes. Magmatism in Extensional Structural
Settings, (McCurry 1975), 2–10. https://doi.org/10.1007/978-3-642-73966-8_1

Karuppasamy, M. B., Natesan, U., Karuppannan, S., Chandrasekaran, L. N., Hussain, S.,
Almohamad, H. (2022). Multivariate Urban Air Quality Assessment of Indoor and Outdoor
Environments at Chennai Metropolis in South India. Atmosphere, 13(1627).
https://doi.org/https://doi.org/10.3390/ atmos13101627

Kaufman, A. A., Anderson, B. I. (2010). Electrical Methods of Borehole Geophysics. In Methods


in Geochemistry and Geophysics (1st ed., Vol. 44). https://doi.org/10.1016/S0076-
6895(10)44007-X

Kayode, J. S., Arifin, M. H., Nawawi, M. (2019). Characterization of a Proposed Quarry Site
using Multi-Electrode Electrical Resistivity Tomography. Sains Malaysiana, 48(5), 945–
963. https://doi.org/10.17576/jsm-2019-4805-03

Kazemi, Z., Hesami Arani, M., Panahande, M., Kermani, M., Kazemi, Z. (2021). Chemical
quality assessment and health risk of heavy metals in groundwater sources around
Saravan landfill, the northernmost province of Iran. International Journal of Environmental
Analytical Chemistry, 1–19. https://doi.org/10.1080/03067319.2021.1958800

Kebonye, N. M., Eze, P. N. (2019). Zirconium as a suitable reference element for estimating
potentially toxic element enrichment in treated wastewater discharge vicinity. Environ
Monit Assess, 191: 705. https://doi.org/10.1007/s10661-019-7812-6

Keller, G. V. (1978). Principles of induced polarization for geophysical exploration. Earth-


Science Reviews. Earth-Science Reviews, 14(1), 73–74.

Kennou, B., El Meray, M., Romane, A., Arjouni, Y. (2015). Assessment of heavy metal
availability (Pb, Cu, Cr, Cd, Zn) and speciation in contaminated soils and sediment of
discharge by sequential extraction. Environmental Earth Sciences, 74(7), 5849–5858.
https://doi.org/10.1007/s12665-015-4609-y

Kessouri, P., Furman, A., Huisman, J. A., Martin, T., Mellage, A., Ntarlagiannis, D., Placencia-
Gomez, E. (2019). Induced polarization applied to biogeophysics: recent advances and
future prospects. Near Surface Geophysics, 17(6), 595–621.
https://doi.org/10.1002/nsg.12072

Khan, R. K., Strand, M. A. (2018). Road dust and its effect on human health : a literature review.
Epidemiology and Healthy, 40(e2018013), 1–11.
https://doi.org/https://doi.org/10.4178/epih.e2018013

Khan, S., Anjum, R., Turab, S., Ahmed, N., Ihtisham, M. (2022). Chemosphere Technologies for
municipal solid waste management : Current status , challenges , and future
perspectives. Chemosphere, 288(P1), 132403.
https://doi.org/10.1016/j.chemosphere.2021.132403
191

Khattak, S. A., Rashid, A., Tariq, M., Ali, L., Gao, X., Ayub, M., Javed, A. (2021). Potential risk
and source distribution of groundwater contamination by mercury in district Swabi,
Pakistan: Application of multivariate study. Environment, Development and Sustainability,
23(2), 2279–2297. https://doi.org/10.1007/s10668-020-00674-5

Kimani, N. (2012). Environmental Pollution and Impacts on Public Health :Implications of the
Dandora Municipal Dumping Site in Nairobi, Kenya. Environmental Pollution and Impacts
on Public Health, 1, 14.

Koda, E., Tkaczyk, A., Lech, M., Osinski, P., 2017. Application of electrical resistivity
data sets for the evaluation of the pollution concentration level within landfill
subsoil. Appl. Sci. 7 (3), 262. https://doi.org/10.3390/app7030262.

Koliyabandara, S. M. P. A., Asitha, T. C., Sudantha, L., Siriwardana, C. (2020). Assessment of


the impact of an open dumpsite on the surface water quality deterioration in Karadiyana,
Sri Lanka. Environmental Nanotechnology, Monitoring and Management, 14(September
2019), 100371. https://doi.org/10.1016/j.enmm.2020.100371

Kosheleva, N.E., Vlasov, D.V. Ksenia S. N. (2019). Physicochemical Properties of Road Dust in
Moscow. Geography, Environment, Sustainability 12 (4): 95–113.
https://doi.org/10.24057/2071-9388-2019-55.

Kozłowski, M., Sojka, M. (2019). Applying a Modified DRASTIC Model to Assess Groundwater
Vulnerability to Pollution : A Case Study in Central Poland. Pol. J. Environ. Stud, 28(3),
1223–1231. https://doi.org/10.15244/pjoes/84772

Kozłowski, M., Sojka, M. (2019). Applying a Modified DRASTIC Model to Assess Groundwater
Vulnerability to Pollution : A Case Study in Central Poland. Pol. J. Environ. Stud, 28(3),
1223–1231. https://doi.org/10.15244/pjoes/84772

Kozłowski, M., Sojka, M. (2019). Applying a Modified DRASTIC Model to Assess Groundwater
Vulnerability to Pollution : A Case Study in Central Poland. Pol. J. Environ. Stud, 28(3),
1223–1231. https://doi.org/10.15244/pjoes/84772

Kraidy, N., Armel, B., Bi, B., Emile, B., Daniel, A. K. (2022). Distribution and Characterization of
Heavy Metal and Pollution Indices in Landfill Soil for Its Rehabilitation by
Phytoremediation. 151–172. https://doi.org/10.4236/gep.2022.101011

Kuang, W., Chen, Z., Shi, K., Sun, H., Li, H., Huang, L., Bi, J. (2020). Adverse health effects of
lead exposure on physical growth, erythrocyte parameters and school performances for
school-aged children in eastern China. Environment International, 145, 106130.
https://doi.org/10.1016/j.envint.2020.106130

Kukier, U., Peters, C. A., Chaney, R. L., Angle, J. S., Roseberg, R. J. (2004). The Effect of pH
on Metal Accumulation in Two Alyssum Species . Journal of Environmental Quality, 33(6),
2090–2102. https://doi.org/10.2134/jeq2004.2090

Kumar, P. G., Lekhana, P., Tejaswi, M., Chandrakala, S. (2020). Effects of vehicular emissions
on the urban environment- a state of the art. Materials Today: Proceedings, 45, 6314–
6320. https://doi.org/10.1016/j.matpr.2020.10.739
192

Kumpiene, J., Lagerkvist, A., Maurice, C. (2008). Stabilization of As, Cr, Cu, Pb and Zn in soil
using amendments - A review. Waste Management, 28(1), 215–225.
https://doi.org/10.1016/j.wasman.2006.12.012

Langa, C., Hara, J., Wang, J., Nakamura, K., Watanabe, N., Komai, T. (2021). Dynamic
evaluation method for planning sustainable landfills using GIS and multi- criteria in areas
of urban sprawl with land-use conflicts. PLoS ONE, 16((8): e0254441).
https://doi.org/10.1371/journal.pone.0254441

Latosińska, J. (2020). Risk assessment of soil contamination with heavy metals from sewage
sludge and ash after its incineration. Desalination and Water Treatment, 199(September
2019), 297–306. https://doi.org/10.5004/dwt.2020.25837

Lashen, Z. M., Shams, M. S., El-sheshtawy, H. S., Slaný, M., Elmahdy, S. M. (2022).
Remediation of Cd and Cu contaminated water and soil using novel nanomaterials
derived from sugar beet processing- and clay brick factory-solid wastes. 428, 0–2.
https://doi.org/10.1016/j.jhazmat.2021.128205

Lau, A. M. P., Ferreira, F. J. F., Stevanato, R., da Rosa Filho, E. F. (2019). Geophysical and
physicochemical investigations of an area contaminated by tannery waste: a case study
from southern Brazil. Environmental Earth Sciences, 78(16), 1–16.
https://doi.org/10.1007/s12665-019-8536-1

Lelieveld, J., Evans, J. S., Fnais, M., Giannadaki, D., Pozzer, A. (2015). The contribution of
outdoor air pollution sources to premature mortality on a global scale. Nature, 525(17).
https://doi.org/10.1038/nature15371

Leung, C. C., Yu, I. T. S., Chen, W. (2012). Silicosis. The Lancet, 379(9830), 2008–2018.
https://doi.org/10.1016/S0140-6736(12)60235-9

Li, D., Chen, J., Zhang, Y., Gao, Z., Ying, N., Gao, J., Zhang, K. (2021). Dust emissions from
urban roads using the AP-42 and TRAKER methods: A case study. Atmospheric Pollution
Research, 12(5), 101051. https://doi.org/10.1016/j.apr.2021.03.014

Liao, N. S., Sidney, S., Deosaransingh, K., Eeden, S. K. Van Den, Schwartz, J., Alexeeff, S. E.
(2021). Particulate Air Pollution and Risk of Cardiovascular Events Among Adults With
Infarction. Journal of the American Heart Association Downloaded, 10:e019758.
https://doi.org/10.1161/JAHA.120.019758

Liao, Q., Deng, Y., Shi, X., Sun, Y., Duan, W., Wu, J. (2018). Delineation of contaminant plume
for an inorganic contaminated site using electrical resistivity tomography: comparison with
direct-push technique. Environmental Monitoring and Assessment, 190(4).
https://doi.org/10.1007/s10661-018-6560-3

Loke, M. H. (2010). RES2DINV ver. 3.59 - Rapid 2-D Resistivity IP inversion using the least-
squares method Wenner (α,β,γ), dipole-dipole, inline pole-pole, pole- dipole, equatorial
dipole-dipole, offset pole-dipole, Wenner-Schlumberger, gradient and non-conventional
arrays. 1–148.

Loke, M. H., Barker, R. D. (1996). Rapid least-squares inversion of apparent resistivity


pseudosections by a quasi-Newton method. Geophysical Prospecting, 44(1), 131–152.
https://doi.org/10.1111/j.1365-2478.1996.tb00142.x
193

Loska, K., Wiechulła, D., Korus, I. (2004). Metal contamination of farming soils affected by
industry. Environment International, 30(2), 159–165. https://doi.org/10.1016/S0160-
4120(03)00157-0

Lothe, A. G.,Sinha, A. (2017). Development of model for prediction of Leachate Pollution Index
(LPI) in absence of leachate parameters. Waste Management, 63, 327–336.
https://doi.org/10.1016/j.wasman.2016.07.026

Lucchini, R. G., Albini, E., Benedetti, L., Borghesi, S., Coccaglio, R., Malara, E. C., … Alessio,
L. (2007). High prevalence of Parkinsonian disorders associated to manganese exposure
in the vicinities of ferroalloy industries. American Journal of Industrial Medicine, 50(11),
788–800. https://doi.org/10.1002/ajim.20494

Lund, E. D. (2008). Soil electrical conductivity. In T. T. Sally Logsdon, Dave Clay, Demie Moore
(Ed.), Soil Science: Step-by-Step Field Analysis (pp. 137–146).
https://doi.org/10.2136/2008.soilsciencestepbystep.c11

Madani, R. A., Kermani, S., Sami, M., Esfandiari, Z., Karamian, E. (2020). Risk assessment of
heavy metals (chromium, nickel, lead, copper, and iron) in fast foods consumed in
Isfahan, Iran. Journal of Bioenergy and Food Science, 7(4), 3032020.
https://doi.org/10.18067/jbfs.v7i4.303

Mama, C. N., Nnaji, C. C., Nnam, J. P., Opata, O. C. (2021). Environmental burden of
unprocessed solid waste handling in Enugu State, Nigeria. Environmental Science and
Pollution Research, 28(15), 19439–19457. https://doi.org/10.1007/s11356-020-12265-y

Mandal, S., Bhattacharya, S., Paul, S. (2022). Assessing the level of contamination of metals in
surface soils at thermal power area: Evidence from developing country (India).
Environmental Chemistry and Ecotoxicology, 4, 37–49.
https://doi.org/10.1016/j.enceco.2021.11.003

Maria Csuros, C. C. (2002). Environmental Sampling and Analysis for Metals. (1st ed.). CRC
Press.2002. https://doi.org/10.1201/9781420032345

Marino, A., Pariso, P. (2020). Comparing European countries’ performances in the transition
towards the Circular Economy. The Science of the Total Environment, 729, 138142.
https://doi.org/10.1016/j.scitotenv.2020.138142

Marques, T., Matias, M. S., da Silva, E. F., Durães, N., Patinha, C. (2021). Temporal and spatial
groundwater contamination assessment using geophysical and hydrochemical methods:
The industrial chemical complex of estarreja (portugal) case study. Applied Sciences
(Switzerland), 11(15). https://doi.org/10.3390/app11156732

Matsinhe, F. O., Paulo, M. (2020). Estudo Etnográfico sobre os catadores de Lixo da Lixeira de
Hulene (Maputo). Cadernos de África Contemporânea, Vol. 03 |. issn: 2595-5713

Mboera, L. E. G., Mfinanga, S. G., Karimuribo, E. D., Rumisha, S. F., Sindato, C., Karimuribo,
E., Kikuu, C. (2014). The changing landscape of public health in sub-Saharan Africa :
Control and prevention of communicable diseases needs rethinking. Onderstepoort
Journal of Veterinary Research, 81(2), 1–6. https://doi.org/10.4102/ojvr.v81i2.734

Meloni, F., Montegrossi, G., Lazzaroni, M., Rappuoli, D., Nisi, B., Vaselli, O. (2021). Total and
leached arsenic, mercury and antimony in the mining waste dumping area of
194

abbadiasansalvatore (Mt. amiata, central Italy). Applied Sciences (Switzerland), 11(17).


https://doi.org/10.3390/app11177893

Men, C., Liu, R., Wang, Y. Cao, L. Jiao, L., Li, L., Wang, Y. (2022) Impact of particle sizes on health
risks and source-specific health risks for heavy metals in road dust. Environ Sci Pollut Res
29, 75471-75486. https://doi.org/10.1007/s11356-022-21060-w
Mertanen, S. T., Langa, J. J., Ferrari, K. (2013). Catadores de Lixo de Maputo Quem são e
como trabalham? 7768/RLINLD/2013

Mertz, W. (1993). Chromium in human nutrition: A review. Journal of Nutrition, 123(4), 626–633.
https://doi.org/10.1093/jn/123.4.626

Miao, D., Young, S. L., Golden, C. D. (2015). A meta-analysis of pica and micronutrient status.
American Journal of Human Biology, 27(1), 84–93. https://doi.org/10.1002/ajhb.22598

Mitra, D., Banerji, S. (2020). Urban hydrodynamics in the planned township of New Town, West
Bengal, India. Applied Geography, 123(July 2019), 102277.
https://doi.org/10.1016/j.apgeog.2020.102277

Momade, F.J., Ferrara, M., Oliveira, J. T. (1996). Notícia explicativa da carta geológica 2532
Maputo (Escala 1:50 000). Maputo.In Portuguese.

Mor, S., Ravindra, K. (2023). Municipal solid waste landfills in lower- and middle-income
countries: Environmental impacts, challenges and sustainable management practices.
Process Safety and Environmental Protection, 174, 510–530.
https://doi.org/10.1016/j.psep.2023.04.014

Morita, A. K. M., Pelinson, N. de S., Wendland, E. (2020). Persistent impacts of an abandoned


non-sanitary landfill in its surroundings. Environmental Monitoring and Assessment,
192(7). https://doi.org/10.1007/s10661-020-08451-7

Morita, A. K. M., Ibelli-Bianco, C., Anache, J. A. A., Coutinho, J. V., Pelinson, N. S., Nobrega, J.,
Wendland, E. (2021). Pollution threat to water and soil quality by dumpsites and non-
sanitary landfills in Brazil: A review. Waste Management, 131, 163–176.
https://doi.org/10.1016/j.wasman.2021.06.004

Muchangos, A. dos. (1999). Paisagens e Regiões Naturais de Moçambique. 5–163.


https://docplayer.com.br/47220681-Mocambique-paisagens-e-regioes-naturais.html

Muchimbane, A. B. D. (2010). Estudo dos indicadores da contaminação das aguas


subterrâneas por sistemas de saneamento in Situ – Distrito Urbano 4, Cidade de Maputo,
Moçambique. Dissertação de Mestrado. Doi 10.11606/D.44.2010.tde-06052010-153107

Munsell Color. Munsell Soil Color Book; Color Charts; Munsell Colour Company. Inc.:
Newburgh, NY, USA, 2009

Murray, H. H. (2000). Traditional and new applications for kaolin, smectite, and palygorskite: A
general overview. Applied Clay Science, 17(5–6), 207–221.
https://doi.org/10.1016/S0169-1317(00)00016-8
195

Mussa, T. (2020). EBioMedicine Contributing to a better understanding of infectious respiratory


diseases in Mozambique. EBioMedicine, 62, 103128.
https://doi.org/10.1016/j.ebiom.2020.103128

Mwaanga, P., Silondwa, M., Kasali, G., Banda, P. M. (2019). Preliminary review of mine air
pollution in Zambia. Heliyon, 5(9), e02485. https://doi.org/10.1016/j.heliyon.2019.e02485

Nai, J., Lu, Y., Yu, L., Wang, X., Lou, X. W. D. (2017). Formation of Ni–Fe Mixed Diselenide
Nanocages as a Superior Oxygen Evolution Electrocatalyst. Advanced Materials, 29(41),
1–8. https://doi.org/10.1002/adma.201703870

Naidja, L., Ali-khodja, H., Khardi, S. (2018). Sources and levels of particulate matter in North
African and Sub-Saharan cities : a literature review. Environmental Science and Pollution
Research, 25, 12303–12328.https://doi.org/10.1007/s11356-018-1715-x

Nasri, G., Hajji, S., Aydi, W., Boughariou, E., Allouche, N., Bouri, S. (2021). Water vulnerability
of coastal aquifers using AHP and parametric models: methodological overview and a
case study assessment. Arabian Journal of Geosciences, 14(1).
https://doi.org/10.1007/s12517-020-06390-8

Naveed, M., Bukhari, S. S., Mustafa, A., Ditta, A., Alamri, S., El-Esawi, M. A., Siddiqui, M. H.
(2020). Mitigation of nickel toxicity and growth promotion in sesame through the
application of a bacterial endophyte and zeolite in nickel contaminated soil. International
Journal of Environmental Research and Public Health, 17(23), 1–27.
https://doi.org/10.3390/ijerph17238859

Naveed, M., Bukhari, S. S., Mustafa, A., Ditta, A., Alamri, S., El-Esawi, M. A., … Siddiqui, M. H.
(2020). Mitigation of nickel toxicity and growth promotion in sesame through the
application of a bacterial endophyte and zeolite in nickel contaminated soil. International
Journal of Environmental Research and Public Health, 17(23), 1–27.
https://doi.org/10.3390/ijerph17238859

Nematollahi, M. J., Keshavarzi, B., Zaremoaiedi, F., Rajabzadeh, M. A., Moore, F. (2020).
Ecological-health risk assessment and bioavailability of potentially toxic elements (PTEs)
in soil and plant around a copper smelter. Environmental Monitoring and Assessment,
192(10). https://doi.org/10.1007/s10661-020-08589-4

Netto, L. G., Filho, W. M., Moreira, C. A., di Donato, F. T., Helene, L. P. I. (2021). Delineation of
necroleachate pathways using electrical resistivity tomography (ERT): Case study on a
cemetery in Brazil. Environmental Challenges, 5, 100344.
https://doi.org/10.1016/j.envc.2021.100344

NIH (National institutes of Health). (2012). Manganese. In Manganese.


http://www.ncbi.nlm.nih.gov/books/nbk158872 (accessed on 10 June 2022).

NIH (National institutes of Health). (2021). Zinc.


https://doi.org/https://ods.od.nih.gov/factsheets/Zinc-HealthProfessional/#h4(accessed on
08July 2022)

Nogueira, G., Stigter, T. Y., Zhou, Y., Mussa, F., Juizo, D. (2019). Understanding groundwater
salinization mechanisms to secure freshwater resources in the water-scarce city of
Maputo, Mozambique. Science of the Total Environment, 661, 723–736.
https://doi.org/10.1016/j.scitotenv.2018.12.343
196

Nta, S. A., Ayotamuno, M. J., Igoni, A. H., Okparanma, R. N. (2020). Leachate Characterization
from Municipal Solid Waste Dump Site and Its Adverse Impacts on Surface Water Quality
Downstream - Uyo Village Road, Akwa Ibom State - Nigeria. Journal of Engineering
Research and Reports, 13(2), 11–19. https://doi.org/10.9734/jerr/2020/v13i217096

Obiri-Nyarko, F., Duah, A. A., Karikari, A. Y., Agyekum, W. A., Manu, E., Tagoe, R. (2021).
Assessment of heavy metal contamination in soils at the Kpone landfill site, Ghana:
Implication for ecological and health risk assessment. Chemosphere, 282, 131007.
https://doi.org/10.1016/j.chemosphere.2021.131007

Odom, F., Gikunoo, E., Arthur, E. K., Agyemang, F. O., Mensah-Darkwa, K. (2021).
Stabilization of heavy metals in soil and leachate at Dompoase landfill site in Ghana.
Environmental Challenges, 5(June), 100308. https://doi.org/10.1016/j.envc.2021.100308

Odukoya, K., Olalemi, A. A. (2023). Mineralogy and Geochemical Characterization of


Geophagic Clays Consumed in Parts of Southern Nigeria. Journal of Trace Elements and
Minerals, 100063. https://doi.org/10.1016/j.jtemin.2023.100063

Oke, S. A. (2020). Regional Aquifer Vulnerability and Pollution Sensitivity Analysis of Drastic
Application to Dahomey Basin of Nigeria. Res. Public Health, 17(7), 2609.
https://doi.org/https://doi.org/10.3390/ijerph17072609

Olayinka, A. I., Yaramanci, U. (2000). Use of block inversion in the 2-D interpretation of
apparent resistivity data and its comparison with smooth inversion. Journal of Applied
Geophysics, 45(2), 63–81. https://doi.org/10.1016/S0926-9851(00)00019-7

Oliveira, L. J. P. (2009). Caracterização Da Pluma De Contaminação Numa Antiga Lixeira Com


O Método De Resistividade Eléctrica (Universidade Nova de Lisboa).
https://run.unl.pt/bitstream/10362/2364/1/Oliveira_2009.pdf

Olla, A., Akinlalu, A., Olayanju, M., Adelusi, O., Adiat, A. (2015). Geophysical and
Hydrochemical Investigation of a Municipal Dumpsite in Ibadan, Southwest Nigeria.
Journal of Environment and Earth Science, 5(14)

Ololade, O. O., Mavimbela, S., Oke, S. A., Makhadi, R. (2019). Impact of leachate from northern
landfill site in Bloemfontein on water and soil quality: Implications for water and food
security. Sustainability (Switzerland), 11(15). https://doi.org/10.3390/su11154238

Palalane, J., Segala, I. O. (2008). Urbanização e desenvolvimento municipal em Moçambique:


gestão de resíduos sólidos. https://limpezapublica.com.br/urbanizacao-e-
desenvolvimento-municipal-em-mocambique-capitulo-gestao-de-residuos-
solidos/(accessed on 01June 2018).

Pan, S., Dixon, K. L., Nawaz, T., Rahman, A., Selvaratnam, T. (2021). Evaluation of
Galdieriasulphuraria for nitrogen removal and biomass production from raw landfill
leachate. Algal Research, 54, 102183. https://doi.org/10.1016/j.algal.2021.102183

Parvin, F., Tareq, S. M. (2021). Impact of landfill leachate contamination on surface and
groundwater of Bangladesh: a systematic review and possible public health risks
assessment. Applied Water Science, 11(6). https://doi.org/10.1007/s13201-021-01431-3

Paul, S., Surabhi, C. (2021). An investigation of groundwater vulnerability in the North 24


parganas district using DRASTIC and hybrid-DRASTIC models : A case study.
Environmental Advances, 5, 100093. https://doi.org/10.1016/j.envadv.2021.100093
197

Perera, F. (2018). Pollution from fossil-fuel combustion is the leading environmental threat to
global pediatric health and equity: Solutions exist. International Journal of Environmental
Research and Public Health, 15(1). https://doi.org/10.3390/ijerph15010016

Perrone, A., Lapenna, V., Piscitelli, S. (2014). Electrical resistivity tomography technique for
landslide investigation: A review. Earth-Science Reviews, 135, 65–82.
https://doi.org/10.1016/j.earscirev.2014.04.002

Perrone, A., Lapenna, V., Piscitelli, S. (2014). Electrical resistivity tomography technique for
landslide investigation: A review. Earth-Science Reviews, 135, 65–82.
https://doi.org/10.1016/j.earscirev.2014.04.002

Piedrahita, R., Anenberg, S. C., Henze, D. K., Lacey, F., Irfan, A., Kinney, P., Kleiman, G.
(2017). Air pollution-related health and climate benefits of clean cookstove programs in
Mozambique Air pollution-related health and climate benefits of clean cookstove
programs in Mozambique. Environmental Research Letters, 12.
https://doi.org/10.1088/1748-9326/aa5557

Zhang, WJ., Qiu, QW (2010). Analysis on contaminant migration through vertical barrier walls in
a landfill in China. Environ Earth Sci 61, 847–852. https://doi.org/10.1007/s12665-009-
0399-4

Pikuła, D., Stępień, W. (2021). Effect of the degree of soil contamination with heavy metals on
their mobility in the soil profile in a microplot experiment. Agronomy, 11(5), 1–11.
https://doi.org/10.3390/agronomy11050878

Pio, C., Alves, C., Nunes, T., Cerqueira, M., Lucarelli, F., Nava, S., Querol, X. (2020). Source
apportionment of PM 2 . 5 and PM 10 by Ionic and Mass Balance (IMB ) in a traffic-
influenced urban atmosphere , in Portugal. Atmospheric Environment Journal, 223.
https://doi.org/10.1016/j.atmosenv.2019.117217

Plum, L. M., Rink, L., Hajo, H. (2010). The essential toxin: Impact of zinc on human health.
International Journal of Environmental Research and Public Health, 7(4), 1342–1365.
https://doi.org/10.3390/ijerph7041342

RAIS. The Risk Assessment Information System (RAIS); U.S. Department of Energy’s Oak
Ridge Operations Office (ORO): Oak Ridge, TN, USA, 2022. Available
online: https://rais.ornl.gov/ (accessed on 10 May 2022).

Rajagopalan, Sadeer G. Al-Kindi, R. D. B. (2018). Air Pollution and Cardiovascular Disease


Jacc State-of-the-art Review. Journal of the American College of Cardiology, 72(17),
2054–2070. https://doi.org/10.1016/j.jacc.2018.07.099

Rapti, D., Masi, S., Sdao, F. (2021). SIVRAD: An integrated system for the assessment of the
environmental risk from solid waste landfills - Guidelines. Italian Journal of Groundwater,
10(2), 49–62. https://doi.org/10.7343/as-2021-507

Rapti-Caputo, D., Sdao, F., Masi, S. (2006). Pollution risk assessment based on hydrogeological
data and management of solid waste landfills. Engineering Geology, 85, 122–131.
https://doi.org/10.1016/j.enggeo.2005.09.033

Reynolds, J. M. (1997). An Introduction to Applied and Environmental Geophysics (Reynolds


G). UK.
198

Ruengruehan, K., Junggoth, R., Suttibak, S., Sirikoon, C., Sanphoti, N. (2021). Contamination of
cadmium, lead, mercury and manganese in leachate from open dump, controlled dump
and sanitary landfill sites in Rural Thailand: A case study in Sakon Nakhon Province. Nature
Environment and Pollution Technology, 20(3), 1257–1261.
https://doi.org/10.46488/NEPT.2021.V20I03.036

Rwakarehe, E. (2022). Review of Strategies for Curbing Traffic Congestion in Sub-Saharan


Africa Cities: Technical and Policy Perspectives. Tanzania Journal of Engineering and
Technology, 40(2), 24–32. https://doi.org/10.52339/tjet.v40i2.730

Saentho, A., Wisawapipat, W., Lawongsa, P., Aramrak, S., Prakongkep, N., Klysubun, W.,
Christl, I. (2022). Speciation and pH- and particle size-dependent solubility of phosphorus
in tropical sandy soils. Geoderma, 408, 115590.
https://doi.org/10.1016/j.geoderma.2021.115590

Sahu, D., Ramteke, S., Dahariya, N. S., Sahu, B. L., Patel, K. S., Matini, L., … Hoinkis, J.
(2016). Assessment of Road Dust Contamination in. (January), 77–88.

Sarmento, L., Tokai, A., Hanashima, A. (2015). Analyzing the structure of barriers to municipal
solid waste management policy planning in Maputo city , Mozambique. Environmental
Development, 16, 76–89. https://doi.org/10.1016/j.envdev.2015.07.002

Sato, M., Mooney, H. M. (1960). MOTOAKI SATOt. Geophysics, XXV(1), 226–249.


https://doi.org/https://doi.org/10.1190/1.1438689

Schaumlöffel, D. (2012). Nickel Species: Analysis and Toxic Effects. Journal of Trace Elements
in Medicine and Biology 26 (1): 1–6. https://doi.org/10.1016/j.jtemb.2012.01.002

Schlüter, T. (2008). Geological Atlas of Africa -With Notes on Stratigraphy, Tectonics, Economic
Geology, Geohazards and Geosites of Each Country.
http://www.geokniga.org/bookfiles/geokniga-geological-atlas-africa.pdf

Seidl, M., Le Roux, J., Mazerolles, R., Bousserrhine, N. (2021). Assessment of leaching risk of
trace metals, PAHs and PCBs from a brownfield located in a flooding zone.
Environmental Science and Pollution Research, (Ademe 2014).
https://doi.org/10.1007/s11356-021-15491-0

Sentenac, P., Benes, V., Keenan, H. (2018). Reservoir assessment using non-invasive
geophysical techniques. Environmental Earth Sciences, 77(7), 1–14.
https://doi.org/10.1007/s12665-018-7463-x

Serra, C. (2012). Da problemática Ambiental à mudança: rumo à um mundo melhor.


ISBN%0A9789896700300.

Shah, S. H. I. A., Yan, J., Ullah, I., Aslam, B., Tariq, A., Zhang, L., Mumtaz, F. (2021).
Classification of aquifer vulnerability by using the drastic index and geo-electrical
techniques. Water (Switzerland), 13(16). https://doi.org/10.3390/w13162144

Shahab, A., Hui, Z., Rad, S., Xiao, H., Siddique, J., Liang, L. (2022). A comprehensive review
on pollution status and associated health risk assessment of human exposure to selected
heavy metals in road dust across different cities of the world. Environmental
Geochemistry and Health, (0123456789). https://doi.org/10.1007/s10653-022-01255-3
199

Sharma, N., Chaudhry, K. K., Chalapati Rao, C. V. (2004). Vehicular pollution prediction
modelling: A review of highway dispersion models. Transport Reviews, 24(4), 409–435.
https://doi.org/10.1080/0144164042000196071

Sharma, S., Kaur, I., Nagpal, A. K. (2021). Contamination of rice crop with potentially toxic
elements and associated human health risks—a review. Environmental Science and
Pollution Research, 28(10), 12282–12299. https://doi.org/10.1007/s11356-020-11696-x

Sharma, M., Satyam, N., Reddy, K. R., Chrysochoou, M. (2022). Multiple heavy metal
immobilization and strength improvement of contaminated soil using bio-mediated calcite
precipitation technique. Environmental Science and Pollution Research, 29(34), 51827–
51846. https://doi.org/10.1007/s11356-022-19551-x

Shaylinda, M. Z. N. (2020). Metals contamination on soil and surface water ( earth drainage )
due to leachate migration from Piyungan land Metals contamination on soil and surface
water ( earth drainage ) due to leachate migration from Piyungan land. Materials Science
and Engineering PAPER, 012063. https://doi.org/10.1088/1757-899X/1144/1/012063

Silva, H. F. Silva, N. F. Oliveira,C. M. Matos, M. J. (2021b). Heavy Metals Contamination of


Urban Soils — A Decade Study in the City of Lisbon, Portugal. Soils Systems, 5(27).
https://doi.org/https://doi.org/10.3390/soilsystems 5020027

Skrbic, B., Novakovic, J., Miljevic, N. (2002). Mobility of heavy metals originating from bombing
of industrial sites. Journal of Environmental Science and Health - Part A Toxic/Hazardous
Substances and Environmental Engineering, 37(1), 7–16. https://doi.org/10.1081/ESE-
100108478

Soubra, G., Massoud, M. A., Alameddine, I., Al Hindi, M., Sukhn, C. (2021). Assessing the
environmental risk and pollution status of soil and water resources in the vicinity of
municipal solid waste dumpsites. Environmental Monitoring and Assessment, 193(12).
https://doi.org/10.1007/s10661-021-09640-8

Sparling, G. P. (2020). Soil Quality: Indicators. Managing Soils and Terrestrial Systems, 357–
360. https://doi.org/10.1201/9780429346255-44

Srivastava, P., Singh, B., Angove, M. (2006). Competitive adsorption behavior of heavy metals
on kaolinite. 290(2005), 28–38. https://doi.org/10.1016/j.jcis.2005.04.036

Stuckey, J. W., Neaman, A., Verdejo, J., Navarro-Villarroel, C., Peñaloza, P., Dovletyarova, E.
A. (2021). Zinc Alleviates Copper Toxicity to Lettuce and Oat in Copper-Contaminated
Soils. Journal of Soil Science and Plant Nutrition, 21(2), 1229–1235.
https://doi.org/10.1007/s42729-021-00435-x

Sumalatha, J., Sivapullaiah, P. V., Prabhakara, R. (2022). Immobilization Remediation of a


Heavy Metals Contaminated Soil: A Case Study of Dump Site at Bangalore, India.
Journal of The Institution of Engineers (India): Series A. https://doi.org/10.1007/s40030-
021-00590-5

Sumbana, J., Sacarlal, J., Rubino, S. (2020). Air pollution and other risk factors might buffer
COVID-19 severity in Mozambique. J Infect Dev Ctries 2020, 14(9), 994–1000.
https://doi.org/10.3855/jidc.13057
200

Tan, M., Wang, K., Xu, Z., Li, H., Qu, J. (2020). Study on Heavy Metal Contamination in High
Water Table Coal Mining Subsidence Ponds That Use Different Resource Reutilization
Methods. Water, 12(3348). https://doi.org/10.3390/w12123348

Tao, Z., Deng, H., Li, M., Chai, X. (2020). Mercury transport and fate in municipal solid waste
landfills and its implications. Biogeochemistry, 148(1), 19–29.
https://doi.org/10.1007/s10533-020-00642-1

Telford, W. M., Geldart, L. P., Sheriff, R. E. (2012). Magnetic Methods. In Applied Geophysics.
https://doi.org/10.1017/cbo9781139167932.007

Tezkan, B. (1999). A review of environmental applications of quasi-stationary electromagnetic


techniques. Surveys in Geophysics, 20(3–4), 279–308.
https://doi.org/10.1023/a:1006669218545

Touzani, M., Mohsine, I., Ouardi, J., Kacimi, I., Morarech, M., Habib, M., Mahrad, B. El. (2021).
Mapping the Pollution Plume Using the Self-Potential. Water, 13(961).
https://doi.org/https://doi.org/ 10.3390/w13070961

Tundermann. J. (2000). Kirk-Othmer Encyclopedia of Chemical Technology || Nickel and Nickel


Alloys. In Uhlig’s Corrosion Handbook: Third Edition.
https://doi.org/10.1002/9780470872864.ch59

Tvedten, I., Candiracci, S. (2018). Flooding our eyes with rubbish: urban waste management in
Maputo, Mozambique. Environment and Urbanization, 30(2), 631–646.
https://doi.org/10.1177/0956247818780090

Ubechu, B.O., Israel, H.O., Amadi, C.C. Ofor, I. J. (2021). Seasonal Variations of Cyanide and
Heavy Metals in Groundwater around Enyimba Dumpsite, Aba Southeastern Nigeria.
Journal of Research in Environmental and Earth Sciences, (10), 16–23.

Udosen, N. I. (2021). Geo-electrical modeling of leachate contamination at a major waste


disposal site in south-eastern Nigeria. Modeling Earth Systems and Environment,
(0123456789). https://doi.org/10.1007/s40808-021-01120-9

Ugbor, C. C., Ikwuagwu, I. E., Ogboke, O. J. (2021). 2D inversion of electrical resistivity


investigation of contaminant plume around a dumpsite near Onitsha expressway in
southeastern Nigeria. Scientific Reports, 11(1), 1–14. https://doi.org/10.1038/s41598-021-
91019-3

Urme, S. A., Radia, M. A., Alam, R., Chowdhury, M. U., Ahmed, S. (2021). Europe PMC
Funders Group Dhaka landfill waste practices : addressing urban pollution and health
hazards. 2(2019), 700–716. https://doi.org/10.5334/bc.108.Dhaka

USDA (United States Department of Agriculture). (1998). Soil Quality Indicators: pH. Managing
Soils and Terrestrial Systems. https://doi.org/http://soils.usda.gov(accessed on 08 April
2019).

USDA (United States Department of Agriculture). (2011). Soil Health Quality Indicators:
chemical Properties, soil electrical conductivity.
3.https://www.nrcs.usda.gov/wps/portal/nrcs/detail/soils/health/assessment/?cid=stelprdb
1237387(accessed on 09 April 2019).
201

USEPA. (U.S. Environmental Protection Agency). Screening Levels (RSL) for Chemical
Contaminants at Superfund Sites. U.S.; U.S. Environmental Protection Agency:
Washington, DC, USA, 2013(accessed on 08 July 2019).

USEPA (U.S. Environmental Protection Agency). (2016). Manganese Compounds Hazard


Summary. In Health Effects Notebook for Hazardous Air Pollutants.
https://www.epa.gov/sites/production/files/2016-10/documents/manganese.pdf(accessed
on 12 July 2019).

USEPA. (1989). Risk Assessment Guidance for Superfund, Volume I: Human Health Evaluation
Manual; EPA 540-1-89-002; U.S. Environmental Protection Agency: Washington, DC,
USA.https://semspub.epa.gov/work/HQ/191.pdf (accessed on 18 July 2020).

Valotto, G., Rampazzo, G., Visin, F., Gonella, F., Cattaruzza, E., Glisenti, A., Tieppo, P. (2015).
Environmental and traffic-related parameters affecting road dust composition: A multi-
technique approach applied to Venice area (Italy). Atmospheric Environment, 122, 596–
608. https://doi.org/10.1016/j.atmosenv.2015.10.006

Vasconcelos, L. (2014). Breve apresentação sobre os recursos geológicos de Moçambique.


ComunicacoesGeologicas, 101, 869–874.
http://www.lneg.pt/iedt/unidades/16/paginas/26/30/185

Vicente, E. M. (2011). Aspects of the enginering geologic of Maputo City. Doctoral Thesis,
School of Geological Sciences University of KwaZulu-Natal Durban, 22–38.
http://hdl.handle.net/10413/8078

Vicente, E. M., Jermy, C. A., Schreiner, H. D. (2006a). Urban geology of Maputo ,Mozambique.
The Geological Society of London, (338), 1–
13.https://citeseerx.ist.psu.edu/viewdoc/download?doi=10.1.1.606.7220
rep=rep1&type=pdf (accessed on 15 July 2019).

VOA (Voice of America News). (2018). Desabamento de lixeira deixa 17 mortos em Maputo.
https://www.voaportugues.com/a/desabamento-lixeira-17-mortos-maputo/4260624.html
(accessed on 04July 2020).

Wani, A. L., Ara, A., Usmani, J. A. (2015). Lead toxicity: A review. Interdisciplinary Toxicology,
8(2), 55–64. https://doi.org/10.1515/intox-2015-0009

Wang, X. X., Chang, Y. bang, Deng, J. Z., Chen, J. song. (2021). 3D spatial distribution of old
landfills and groundwater pollution from electrical resistivity tomography with fuzzy set
theory. Exploration Geophysics. https://doi.org/10.1080/08123985.2021.1917292

Wang, W., Han, Z., Wang, J., He, X., Zhou, Z., Hu, X. (2022). Environmental risk assessment
and factors influencing heavy metal concentrations in the soil of municipal solid waste
landfills. Waste Management, 139, 330–340.
https://doi.org/10.1016/j.wasman.2021.11.036

Weibel, G., Eggenberger, U., Schlumberger, S., Mäder, U. K. (2017). Chemical associations
and mobilization of heavy metals in fly ash from municipal solid waste incineration. Waste
Management, 62, 147–159. https://doi.org/10.1016/j.wasman.2016.12.004

Weir, A., Westerhoff, P., Fabricius, L., Hristovski, K., von Goetz, N. (2012). Titanium Dioxide
Nanoparticles in Food and Personal Care Products. Environmental Science &
Technology, 46(4), 2242–2250. doi:10.1021/es204168d
202

Westerhof, A. B. P., Lehtonen, M. I., Mäkitie, H., Manninen, T., Pekkala, Y., Gustafsson, B.,
Tahon, A. (2008). The Tete-Chipata belt : A New Multiple Terrane Element From Western
Mozambique and Southern Zambia. Geological Survey of Finland, 48, 145–166.
https://tupa.gtk.fi/julkaisu/specialpaper/sp_048_pages_145_166.pdf

WHO (World Health Organization). (2020). Chromium in drinking-water.


https://apps.who.int/iris/bitstream/handle/10665/338062/WHO-HEP-ECH-WSH-2020.3-
eng.pdf (accessed on 23 July 2019).

WHO (World Health Organization). (2020a). global air quality guidelines Particulate matter
(PM2.5 and PM10), ozone, nitrogen dioxide, sulfur dioxide and carbon monoxide.
https://doi.org/https://www.who.int/publications/i/item/9789240034228 (accessed on 23
July 2019).

WHO (World Health Organization). (2021). WHO guideline for clinical management of exposure
to lead Executive summary.
https://www.who.int/publications/i/item/9789240036888.(accessed on 08 July 2021).

WHO (World Health Organization). 2021. WHO Guideline for Clinical Management of Exposure
to Lead Executive Summary. https://www.who.int/publications/i/item/9789240036888.
.(accessed on 12 December 2021).

WHO (World Health Organization). (2021). Compendium of WHO and other UN guidance on
health and environment Chapter 2. Air pollution. 2021, 0–25.
https://cdn.who.int/media/docs/default-source/who-compendium-on-health-and-
environment/who_compendium_chapter2_01092021.pdf?sfvrsn=14f84896_5.(accessed
on 10 December 2021).

WHO. (2017). Guidelines for drinking-water quality: fourth edition incorporating the first
addendum. ISBN 978-92-4-154995-0.(accessed on 08 December 2021).

Wieczorek, J., Baran, A., Urbański, K., Mazurek, R., Klimowicz-Pawlas, A. (2018). Assessment
of the pollution and ecological risk of lead and cadmium in soils. Environmental
Geochemistry and Health, 40(6), 2325–2342. https://doi.org/10.1007/s10653-018-0100-5

Wijekoon, P., Koliyabandara, P. A., Cooray, A. T., Lam, S. S., Athapattu, B. C. L., & Vithanage,
M. (2022). Progress and prospects in mitigation of landfill leachate pollution: Risk,
pollution potential, treatment and challenges. Journal of Hazardous Materials, 421,
126627. https://doi.org/10.1016/j.jhazmat.2021.126627

Wu, L., Zhan, L., Lan, J., Chen, Y., Zhang, S., Li, J., Liao, G. (2021). Leachate migration
investigation at an unlined landfill located in granite region using borehole groundwater
TDS profiles. Engineering Geology, 292, 106259.
https://doi.org/10.1016/j.enggeo.2021.106259

Wu, L., Zhan, L., Lan, J., Chen, Y., Zhang, S., Li, J., Liao, G. (2021). Leachate migration
investigation at an unlined landfill located in granite region using borehole groundwater
TDS profiles. Engineering Geology, 292, 106259.
https://doi.org/10.1016/j.enggeo.2021.106259

Wu, Q., Hu, W., Wang, H., Liu, P., Wang, X., Huang, B. (2021). Spatial distribution, ecological
risk and sources of heavy metals in soils from a typical economic development area,
203

Southeastern China. Science of the Total Environment, 780, 146557.


https://doi.org/10.1016/j.scitotenv.2021.146557

Wysocka, M. E., Zabielska-Adamska, K. (2017). Impact of protective barriers on groundwater


quality. 10th International Conference on Environmental Engineering, ICEE 2017.
https://doi.org/10.3846/enviro.2017.063

Yahaya, S. M., Abubakar, F., Abdu, N. (2021). Ecological risk assessment of heavy
metal - contaminated soils of selected villages in Zamfara State , Nigeria. SN Applied
Sciences, 3(2), 1–13. https://doi.org/10.1007/s42452-021-04175-6

Yap, C. K., Chew, W., Al-mutairi, K. A., Nulit, R., Ibrahim, M. H., Wong, K. W., … Al-shami, S. A.
(2022). Assessments of the Ecological and Health Risks of Potentially Toxic Metals in the
Topsoils of Different Land Uses: A Case Study in Peninsular Malaysia. Biology, 11(2).
https://doi.org/https:// doi.org/10.3390/biology11010002

Yan, X., An, J., Yin, Y., Gao, C., Wang, B., Wei, S. (2022). Heavy metals uptake and
translocation of typical wetland plants and their ecological effects on the coastal soil of a
contaminated bay in Northeast China. Science of the Total Environment, 803, 149871.
https://doi.org/10.1016/j.scitotenv.2021.149871

Yesil, H., Molaey, R., Calli, B., Tugtas, A. E. (2021). Extent of bioleaching and bioavailability
reduction of potentially toxic heavy metals from sewage sludge through pH-controlled
fermentation. Water Research, 201, 117303.
https://doi.org/10.1016/j.watres.2021.117303

Yin, Z., Duan, R., Li, P., Li, W. (2021). Water quality characteristics and health risk assessment
of main water supply reservoirs in Taizhou City, East China. Human and Ecological Risk
Assessment, 27(8), 2142–2160. https://doi.org/10.1080/10807039.2021.1958670

Zaki, K., Karhat, Y., Falaki, K. E. L. (2022). Temporal Monitoring and Effect of Precipitation on
the Quality of Leachate from the Greater Casablanca Landfill in Morocco. Pollution, i(2),
407–433. https://doi.org/10.22059/POLL.2021.328358.1158

Zeitoun, R., Vandergeest, M., Vasava, H. B., Machado, P. V. F., Jordan, S., Parkin, G., Biswas,
A. (2021). In-situ estimation of soil water retention curve in silt loam and loamy sand soils
at different soil depths. Sensors (Switzerland), 21(2), 1–15.
https://doi.org/10.3390/s21020447

Zhang, M., Li, X., Yang, R., Wang, J., Ai, Y., Gao, Y., Yu, H. (2019). Multipotential Toxic Metals
Accumulated in Urban Soil and Street Dust from Xining City, NW China: Spatial
Occurrences, Sources, and Health Risks. Archives of Environmental Contamination and
Toxicology, 76(2), 308–330. https://doi.org/10.1007/s00244-018-00592-8

Zhao, G., Zhang, R., Han, Y., Meng, J., Qiao, Q., Li, H. (2021). Pollution characteristics, spatial
distribution, and source identification of heavy metals in road dust in a central eastern city
in China: a comprehensive survey. Environmental Monitoring and Assessment, 193(12).
https://doi.org/10.1007/s10661-021-09584-z

Zimik, H. V., Farooq, S. H., Prusty, P. (2021). Source characterization of trace elements and
assessment of heavy metal contamination in the soil around Tarabalo geothermal field,
Odisha, India. Arabian Journal of Geosciences, 14(11). https://doi.org/10.1007/s12517-
021-07366-y
204
205

13 Appendixes
206
207

13.1 Appendix 1. Characterization of some Potentially Toxic Elements

Chromium (Cr)
Chromium is a trace element of the earth's crust, being a transition metal. It has atomic
number 24 and atomic mass 51,996. It exhibits average concentrations between 0.5 and
250 mg/ kg-1 in soils (Bradl, 2004). It is a bluish white metal, being hard, brittle and
corrosion resistant (Alloway, 2012)

Although Cr is found in soils, this metal is not found in its elemental state in the
environment (WHO 2020). Cr is combined with other elements such as O, Fe and Pb in
the form of oxides (Kumpiene et al., 2008). Although Cr can exist in nine different
oxidation states, Cr3+ and Cr6+ are the two most common forms (Kumpiene et al., 2008).

Trivalent Cr, Cr(III), is the most stable form of Cr in the environment (Kumpiene et al.,
2008). While trivalent Cr exists naturally in the environment, hexavalent Cr, Cr(VI),
comes mainly from anthropogenic pollution sources such as atmospheric deposition of
particulate matter from the metal industry (CCME, 1997).

Cr (VI) is considered the most toxic species and the most easily mobilised in the soil,
and in soils with pH> 6 (Madani et al., 2020). The main source of Cr (VI) in the
environment is anthropogenic pollution as it rarely occurs naturally due to its affinity to
react with organic matter (CCME, 1997; Kumpiene et al., 2008). Cr (III) is much less
mobile and adsorbs more strongly to particles, its solubility decreases at pH> 4 and
above 5.5 complete precipitation occurs (Kumpiene et al., 2008).

The depth movement of Cr (VI) in soils is greater in alkaline soils compared to acid soils,
thus indicating a lower adsorption of these Cr species when the pH is higher, while for
Cr (III) its adsorption increases with soil pH (CCME et al., 1997; Kumpiene et al., 2008).

The solubility of trivalent chromium increases with decreasing soil pH (CCME, 1997). In
addition to pH, immobilization of Cr in soils depends on several other factors such as
oxidation state, clay minerals, competing ions and complexing agents. Immobilization of
Cr III in soils can be explained by adsorption to organic matter and iron oxides, since
humic acids have a strong affinity for this species (CCME, 1997)

Chromium, when added to soil, may have the following possible fates: be oxidised or
reduced, remain in solution, be adsorbed on minerals and organic exchange complexes
or on hydrated oxides of Fe and Mn covering soil particles, form chelates with an organic
208

ligand or precipitate as poorly soluble or highly insoluble compounds (Sumalatha et al.,


2022).

Chromium is an essential nutrient that works by increasing blood glucose tolerance and
thereby potentiating the action of insulin (Mertz, 1993). Cr3+ is an essential nutrient that
assists the body in the use of sugar, protein and fat(Mertz, 1993).

Chromium is one of the most important pollutants in industry and in unplanned solid
waste disposal areas (Andaloussi et al., 2021). All its compounds are considered to be
toxic. Inhalation of high levels of Cr can cause nasal irritation, such as bleeding,
ulceration and even holes in the nasal septum (Ceballos et al., 2021)). Ingestion of large
amounts of Cr can cause stomach disorders and ulcers, convulsions, kidney and liver
damage and even death(Ceballos et al., 2021). Contact with the skin of a certain amount
of Cr can cause ulcerations (CCME, 1997)

Copper (Cu)
Copper is a trace element of the earth's crust, being a transition metal. It has atomic
number 29, atomic mass 63,546 (Csuros et al., 2002). It has an average concentration
in soil between 2 and 50 mg/kg-1. Alloway, (2012) states that it is the first element of the
transition metal group in the periodic table and although it can be found in elemental form
in the environment, it is usually present in minerals containing sulphides, chlorides and
carbonates. It is a malleable metal, ductile and very resistant to corrosion, being the
second metal with the highest thermal and electrical conductivity after silver. The colour
of native copper is reddish brown, and it has a green appearance when altered due to
contact with air (Csuros et al., 2002).

The concentration of Cu in soils varies according to some factors, namely the type of soil
and the changes it undergoes over time, the distance from anthropogenic sources, the
distance to ores and the composition of the parent rock from which the soil comes,
among others( (Sharma et al., 2021). In general, soils from urban centres, industrialised
areas, mines and waste disposal sites have higher concentrations of Cu (Mwaanga et
al., 2019; Obiri-Nyarko et al., 2021; Chen et al., 2022) .

Cu is strongly absorbed by soil particles and has a low mobility when compared to other
metals (H. Chen et al., 2022). As a result of the poor mobility, Cu tends to accumulate in
the soil, however, leaching phenomena of this metal may occur when the retention
capacity of the soil particles is exceeded (Alleoni et al., 2005; Campos, 2010).
209

There are several factors that influence the availability of Cu in the soil namely pH, the
presence of Fe, Mn and Al oxides, cation exchange capacity and oxidation-reduction
potential (Guo et al., 2010).

The capacity of the soil to absorb Cu increases as the pH increases, where maximum
absorption capacities are found at neutral to alkaline pH values (between 6.7 and 7.8).
In the presence of alkaline soil pH conditions, Cu undergoes precipitation. Thus, being
this metal has higher mobility in acidic pH conditions when compared to alkaline pH
conditions (Alleoni et al., 2005; Campos, 2010)

The cation exchange capacity is influenced by the type and quantity of clays present in
the soil, the quantity of organic matter and the pH (Campos, 2010). Cu shows an extreme
affinity for organic matter and binds to it (Campos, 2010). This strong affinity is due to
the high cation exchange potential. Cu found in soil solution is often associated with
dissolved organic matter and is only released in ionic form when there are oxidation
conditions or conditions favourable to the degradation of organic matter by
microorganisms (Latosińska, 2020). Since Cu tends to bind to organic matter, the
mobility of this metal in the soil is conditioned (Guo et al., 2010).

Cu is particularly adsorbed by Fe, Al and Mn oxides. The water content of soils influences
the capacity of the soil to retain Cu in oxidation-reduction reactions (Srivastava et al.,
2006). Despite being essential for some vital human processes, both in food and
enzymatic processes for example, large quantities of Cu produce a toxic effect. The toxic
effects of Cu include anaemia and disorders in the central nervous system and the
cardiovascular system (Hoang et al., 2021).

Lead (Pb)
The average concentration of Pb in the lithosphere is about 14 μg/g and in soil varies
between 2 and 200 mg/ kg-1 (CCME 1999). The most abundant sources of the metal are
the galena minerals (PbS), anglesite (PbSO4) e cerussite (PbCO3) (WHO, 2021).

Natural sources such as volcanoes, forest fires and sea salt account for the natural part
of atmospheric emissions of this metal (CCME, 1999). The most important anthropogenic
sources of this metal in the environment are gasoline combustion (currently a secondary
source, but for several decades a major contributor to Pb pollution), Cu-Zn-Pb smelting,
battery factories, sewage sludge, coal combustion and solid waste incineration (Obiri-
Nyarko et al., 2021; Kennou et al., 2015;Loska et al., 2004).
210

Pb nitrate, Pb chloride and Pb acetate compounds are quite soluble in soil and are easily
leached from the soil. Metallic Pb is relatively insoluble and undergoes oxidation
processes in the soil giving rise to Pb oxides that can later be dissolved in the soil or be
transformed into more stable compounds (CCME, 2006).

The transport, dispersion and enrichment of Pb into the environment from stationary and
mobile sources occur mainly through the atmosphere, but significant discharges may
also occur directly into water or soil (Bai et al., 2017). In soil, Pb tends to localise near
the discharge point due to the low solubility of lead compounds that form on contact with
soil and water (Morita et al., 2021). Generally, lead accumulates at the soil surface in the
first centimetres (2–5 cm) decreasing in depth (Binh et al., 2021). Organic matter in soils
tends to decrease the bioavailability of Pb (Wieczorek et al., 2018; CCME, 2006).

Campos, (2010); Saentho et al. 2022) report that pH and cation exchange capacity
(CEC) are the main factors involved in Pb immobilization, while organic matter and clay
content are the key constituents in Pb adsorption, as well as the role of Mn and Fe oxides
and possibly Al.

In solid waste disposal areas, Pb contamination has been associated with waste
electronic equipment, refillables, paint cans, varnishes, organic solvents, and waste
glass. For elm from cars and roads that are significant for lead (Azeez et al. 2011;
Kennou et al., 2015)

The effects of Pb on human health depend on the intensity, duration and, depending on
the level of exposure, may result in a series of toxic effects, such as haematological,
neurological, psychological, renal, gene mutation and on the central nervous system and
reproduction are examples of the cumulative effect on the organism (Binh et al. 2021;
CCME, 1999). It usually enters the body through ingestion or inhalation and can affect
almost all organs and the nervous system (WHO 2021). High exposures in children can
cause kidney damage, brain damage, neurological problems including difficulty
concentrating, aggressive learning and behaviour (WHO 2021; Kuang et al., 2020;
Demayo et al., 1982)

Manganese (Mn)
Manganese occurs naturally in over 100 minerals with background levels in soil ranging
from 40 to 900 mg/kg, with an estimated average background concentration of 330 mg/kg
(USEPA, 2016). It is a transition metal of group 7 of the periodic table and one of the
most abundant metals in the earth's crust (NIH, 2012).
211

Manganese is naturally released into the environment by volcanic eruptions. However,


the largest emissions are due to industry, the combustion of fossil fuels, and the erosion
of manganese-containing soils (Lucchini et al., 2007). Almost 80% of industrial
manganese emissions are attributable to iron and steel production facilities (USEPA,
2016). Emissions from power plants and coke ovens contribute about 20% (USEPA,
2016). Manganese can also be released into the environment through the use of gasoline
with manganese additive (Lucchini et al., 2007). Thus, all humans are exposed to
manganese, and manganese is a normal component of the human body (USEPA, 2016).

The cycles of manganese in soil involve the bivalent (Mn2+) and trivalent (Mn3+)forms of
the metal. There is a dynamic equilibrium between all forms. The divalent form is
transformed, through biological oxidation, into the trivalent form which is subsequently
reduced to Mn2+. In very acid soils, oxidation is considered low, on the other hand, in
alkaline soils, the divalent form practically disappears (Lucchini et al., 2007).

In areas around waste dumps, Mn could be associated with waste paints, cosmetics,
pharmaceuticals, textiles, electrical waste, traffic-related emissions and electronic
equipment (Islamd et al., 2021a).

Manganese is an important element for animal and plant life, but its high consumption or
exposure can cause serious problems (NIH, 2012). In the human body, the metal is
absorbed in the small intestine. Prolonged exposures to manganese compounds, by
inhalation or orally, can cause adverse effects on the nervous and respiratory
system(USEPA, 2016; Gujre et al., 2021). Problems are reported to cause dermatitis,
decreased serum cholesterol, increased blood calcium levels, infertility, decreased
glucose metabolism, decreased protein metabolism, decreased growth and skeletal
disorders, pancreatic dysfunction, increased blood pressure, decreased immune
function, depression of mammary gland activity (Lucchini et al., 2007).

Nickel (Ni)
Nickel is a trace element in the earth's crust and is a transition metal. Its average
concentration in soils varies between 2 and 100 mg/kg-1 (Bradl, 2005).

It has a silvery white colour, is hard, bright, malleable, ductile, corrosion resistant, with
magnetic properties and good thermal and electrical conductivity (Maria Csuros, 2002).
Is strongly reactive with ambient air, can even spontaneously ignite and easily forms
metallic alloys with other metals such as Fe, Cu, Zn and Cr (CCME, 2015).
212

Ni shows a strong affinity for Fe and S forming Fe sulphide compounds such as


pentlandite [(Ni,Fe)9S8] in igneous rocks, millerite (NiS) and ulmanite (NiSbS) in
mineralised areas (Nai et al., 2017). There are also Fe and Ni rich minerals, oxides and
silicates that can often be found in soils due to prolonged weathering of the parent rock
(Bisht et al., 2022).

The main uses of Ni are the production of alloys, including Ni stainless steel, the
manufacture of batteries and welding electrodes, and the production of chemicals such
as nickel sulphate, nickel chloride and some catalysts (Tundermann et al., 2000).

In topsoils, especially in soils with high organic matter, Ni combines with it to form soluble
chelates. Ni adsorption in soils decreases with increasing organic matter (Kukier et al.,
2004). The solubility of Ni in soil is influenced by pH, the presence of organic matter and
clays and manganese and iron ions in soil, and the mobility of nickel in soil increases
with decreasing pH(Kukier et al., 2004). Most Ni compounds are soluble for pH values
below 6.5. In comparison with other heavy metals such as Cd and Zn, Ni is more mobile
in soils (Sumalatha et al., 2022)

In solid waste disposal areas, soil contamination by Ni has been associated with
enrichment by the bad dumping of waste electronic equipment, refillables, paint cans,
varnishes, organic solvents and waste glass (Alghamdi et al., 2021).

Due to its potential toxicity, inhalation effects of nickel may include chronic bronchitis,
asthma, reduced vital capacity, pulmonary emphysema, gastrointestinal and
neurological effects, including lung cancer (Naveed et al., 2020). Dermal exposure can
trigger allergic problems (Astuti et al., 2021).

Zinc (Zn)
Zinc is a trace element in the earth's crust. Its atomic number is 30 and its atomic mass
is 65.39. It has an average concentration in soil between 10 and 300 mg/kg-1 (Jackson,
1989). It is a bluish-white, shiny metal that becomes dull in contact with air due to the
formation of a film (Jackson, 1989)(NIH, 2021)

Zn is one of the elements most commonly found in the earth's crust. It is an element that
is not found in its elemental form in the environment, and is thus extracted from the
mineral sphalerite [(ZnFe)S] (CCME, 2018) Sphalerite is the most important Zn mineral
(CCME, 2018). Zn metallurgy is the main emitter of Zn to the atmosphere, as well as Cd.
Cd replaces Zn in sphalerite and other minerals (CCME, 2018)
213

Zn is considered an essential element in human nutrition and is also an important


component of enzymes (NIH, 2021)

The main industrial use of Zn is summarized in the application of this metal as corrosion
protection for Fe and steel, application in batteries and production of metal alloys, cans
and bronze. Zn oxide is one of the most used compounds in industry, in the production
of paints, plastics, cosmetic products, pharmaceuticals, textiles and electrical and
electronic equipment(Plum et al., 2010; CCME, 2018b).

Zn concentrations in air, water and soil have been increasing due to anthropogenic
activities such as mining, combustion of fossil fuels, steel production, uncontrolled waste
disposal and incineration (Stuckey et al., 2021). Once present in the environment, Zn
remains in the soil, forming insoluble compounds (Stuckey et al., 2021). However, it is
one of the most mobile metals in soil due to its high solubility in the presence of neutral
or acidic pH soil solutions (Guo et al., 2010).

Zn is very reactive in soil and is present both in soluble and insoluble solutions (ATSDR,
2002). Zn is often associated with clay minerals or metal oxides (Srivastava et al., 2006).
The pH has been pointed out in several studies as one of the main factors influencing
the mobility of Zn in the soil, since as the pH decreases the solubility of Zn increases
(Campos, 2010; Kennou et al., 2015).

Zn is extremely mobile and becomes biologically available to organisms in low pH


environments, especially at pH values below 5. For pH values below 7.7 Zinc occurs in
the soil solution in the second oxidation state (Zn2+) and for pH values above 7.7 the
dominant form of soil Zn is Zn hydroxide. The leaching process of Zn in soils occurs more
easily when they have acidic pH values (CCME, 2018)

Zinc is an essential trace element in all living systems, playing an important role in
enzyme activities, nucleic acid metabolism, protein synthesis, maintenance of membrane
structure and function, hormone activity, reproduction and sexual maturity (Plum et al.,
2010; Miao et al., 2015).

In solid waste disposal areas, Zinc enrichment may be associated with batteries, metal
alloys, paint cans, cosmetics, pharmaceuticals, textiles, and electrical and electronic
equipment, accumulated over time (Fiala et al., 2021;Islamd et al., 2021a). Zinc toxicity
leads to loss of appetite, dehydration, weakness, weight loss or gain, diarrhea, and
jaundice (Stuckey et al., 2021).

Zirconium (Zr)
214

Zirconium is considered as an as an essential trace metal in the biosphere. It is an


abundant element in the earth's crust (Ghosh et al., 1992). Its average concentration in
rock is 170 mg/kg, in soil 300 mg/kg, in marine sediments 132 mg/kg, and in seawater 4
ppb. Since carbonates, hydroxides, oxides, and silicates of Zr are insoluble, the
dissolved salts can be rapidly precipitated from sea water, producing high concentrations
in marine sediments(Kebonye et al., 2019). The principal minerals are baddeleyite, a
form of ZrO2, and zircon or zirconium orthosilicate (ZrSiO4) (Ghosh et al., 1992).

In its metallic form, Zr is hard and resistant to corrosion, heat, and acid. Cationic salts
are hydrolysed to form insoluble but stable zirconyl or zirconium oxysalts. ZrO2+ ions
polymerise readily with increasing pH, allowing extensive olation and oxolation (Hodson,
2002). Zr forms unstable simple cation salts such as halides and sulfates, and boiling
stable anionic zirconates (Kebonye et al., 2019). It has a great affinity for free PO3+ ions,
forming insoluble precipitates, which hinders the coprecipitation of other metals such as
Cd, Cu, Mn, and Pb (Hodson et al., 2002). In a laboratory study, (Hodson, 2002)
described the low mobility of Zr in soils, which leads to a concentration at the soil surface.

In areas surrounding landfills, Zr can be enriched by residues from paints, cosmetics,


pharmaceuticals, textiles, electrical waste, traffic-related emissions, and electronic
equipment (Gautam et al., 2021)

The main exposure risks are associated with industrial inhalation and dermal exposure.
Due to the low solubility of zirconium, ecological health concerns in the aquatic
environment and in soils are minimal. Zr is classified with low toxicity, but chronic
exposure can cause respiratory tract irritation, dermatitis, and pulmonary fibrosis, with a
few cases reported in non-industrial settings (Jones et al., 2017). The limits of its
concentration in soils are still in debate (Jones et al., 2017).
215

13.2 Paper in their published format and submitted


216
217

13.2.1 Appendix 2. Paper 1


218
Article
Characterization of the Dynamics of Leachate Contamination
Plumes in the Surroundings of the Hulene‐B Waste Dump
in Maputo, Mozambique
Bernardino Bernardo 1,2, Carla Candeias 1 and Fernando Rocha 1,*

1 GeoBioTec Research Centre, Department of Geosciences, University of Aveiro, 3810‐193 Aveiro, Portugal;
bernardino.bernardo@ua.pt (B.B.); candeias@ua.pt (C.C.)
2 Faculty of Earth Sciences and Environment, Pedagogic University of Maputo, Av. do Trabalho,

Maputo 2482, Mozambique


* Correspondence: tavares.rocha@ua.pt

Abstract: The contamination of areas around solid urban waste dumps is a global challenge for the
maintenance of environmental quality in large urban centres in developing countries. This study
applied a geophysical method (electrical resistivity) to identify leachate contamination plumes in
the subsoil and groundwater, as well as to describe their temporal dynamics (2020 and 2021) in the
surroundings of the Hulene‐B waste dump, Maputo, Mozambique. Eight 400 m electrical resistivity
profiles were performed, four profiles in January 2020 and four profiles in May 2021 overlapped,
and the data were inverted with RES2D software. The electrical resistivity models predominantly
indicate an E‐W movement of large contamination plumes that are successively diluted with
saturated media and groundwater, creating zones of less resistive anomalies (< 4.2–< 8.5 Ω∙m)
possibly contaminated at the two analysed seasons, between 2020–2021. The thickness of the
Citation: Bernardo, B.; Candeias, C.; contamination plumes was higher in summer (2020) for profiles 1 and 2, and we associate it with
Rocha, F. Characterization of the the production and migration mechanisms of leachate that are intense in the hot and rainy season.
Dynamics of Leachate Southwest of the dump, profile 4b showed the propagation of anomalous areas on the surface and
Contamination Plumes in the at depth, which are associated with the production of leachate resulting from the continuous
Surroundings of the Hulene‐B Waste decomposition of waste that is continuously deposited in a new area southwest of the dump, thus
Dump in Maputo, Mozambique. generating a slow and continuous migration of leachate at depth, mainly in winter (2021). The
Environments 2022, 9, 19. spatial distribution of contamination plumes during both seasons was reduced significantly farther
https://doi.org/10.3390/
away from the waste deposit, revealing the attenuating effect of groundwater and lithological
environments9020019
substrate (Profile 3).
Academic Editor: Manuel Soto
Keywords: plumes; dynamics; resistivity; contamination; groundwater
Received: 9 December 2021
Accepted: 20 January 2022
Published: 26 January 2022

Publisher’s Note: MDPI stays


1. Introduction
neutral with regard to jurisdictional
claims in published maps and
In recent decades, waste generation has increased in quantity and diversity
institutional affiliations.
worldwide, resulting from population explosion and economic growth [1,2]. The global
annual production of municipal solid waste (MSW) in 2025 is expected to reach about 2.2
billion metric tons [3]. This fact implies many challenges in defining and managing the
final disposal sites for municipal solid waste [4,5]. It is estimated that about 33% of
Copyright: © 2022 by the authors. municipal solid waste produced worldwide is disposed of improperly [3]. Thus, several
Licensee MDPI, Basel, Switzerland. studies have reported environmental problems associated with poor solid waste disposal
This article is an open access article [6]. Morita et al. [7] estimate that open dumpsites will account for 10% of global
distributed under the terms and greenhouse gas emissions by 2025. Municipal solid waste, when disposed of, produces
conditions of the Creative Commons leachate, which is a highly contaminated liquid containing high amounts of inorganic
Attribution (CC BY) license ions, organic compounds, and other toxic elements such as heavy metals and ammonia
(https://creativecommons.org/license [8,9]. In unplanned landfills, leachate is commonly mobilised to the surrounding
s/by/4.0/).

Environments 2022, 9, 19. https://doi.org/10.3390/environments9020019 www.mdpi.com/journal/environments


Environments 2022, 9, 19 2 of 13

environment (soils, surface, and groundwater), causing contamination [10]. Kumar et al.
[11] and Khattak et al. [12] have shown that in waste dump in areas with high
temperatures and precipitation, the process of waste decomposition and leachate
production is higher. The greatest environmental impacts resulting from environmental
contamination by waste are described in developing countries, where the planning of
waste disposal sites is deficient [13]. In Mozambique, namely in Maputo city, the
production of municipal solid waste has been increasing, and daily waste production is
estimated to be around 1250 tonnes [14–16]. Of these, about 1000 tons are deposited in the
largest open‐air dump in Maputo city, the Hulene‐B dump [15]. In this dump, all types of
waste, food, electronic, construction, health and industrial are deposited without any
treatment [14,17]. Studies developed by Nogueira et al. [16] and Vicente et al. [17]
denounced a possible contamination of soils and groundwater in the environment around
the Hulene‐B dump, which suggests the need for its continuous study and monitoring of
its geo‐environmental context [18].
Many methods are currently used to study contamination problems in solid waste
disposal areas, and geophysical methods are pointed out as effective in identifying waste‐
contaminated areas [19,20]. Among these methods, electrical resistivity is widely applied
due to its non‐invasive nature in data acquisition and processing [21–23]. Arifin et al. [20]
and Lau et al. [21] demonstrated that electrical resistivity is effective in locating buried
hazardous waste and identifying contamination plumes resulting from leachate flow.
In general, the typical resistivity of a waste mass is between 15 and 30 Ω∙m in a
saturated medium and between 30–70 Ω∙m in an unsaturated medium [22]. However,
certain materials present in a waste mass can increase or further decrease the resistivity
values [24,25]. Ashes from incinerating plants, tree cuttings and textiles (when they can
retain moisture), previously treated waste and chemicals give rise to low resistivities;
however, plastics, rubber, certain types of building demolition rubble, rubbish preserved
inside plastic bags, heavily compacted newspapers and organic chemicals have high
resistivities [23,25]. The reference resistivity for sandy soils is 50–100 Ω∙m [26,27].
This study applied the geophysical method (electrical resistivity) to identify the
leachate contamination plumes and to describe their temporal dynamics on the
surroundings of the Hulene‐B Waste dump in Maputo, Mozambique (2020 and 2021).

2. Materials and Methods


2.1. Study Area
The Hulene‐B dump is in the Hulene‐B neighbourhood, in the northern direction of
Maputo city, Mozambique (Figure 1) [17,28], a residential area with approximately 48, 717
inhabitants. The dump receives all types of waste produced in Maputo City [15,29–31].
The height of the waste is estimated to be about 6 to 15 m and occupies an area of 17
hectares [25]. The dump is in a former quarry with no previous preparation for waste
reception [26]. These characteristics are described as conducive to contamination of the
local hydrogeological system [27]. Sallwey et al. [30], Serra [31] and Vicente et al. [17] have
associated the Hulene‐B dump with heavy metal (Hg, As, Pb, Cu and Zn) contamination
of surrounding groundwater and soils. The Hulene‐B hydrogeological system is part of
the Tertiary‐Quaternary aquifer system [16]. The aquifer substrate is formed by the layer
of clayey marl to grey clay [18,32]. In the surroundings of the Hulene‐B dump, the
localised presence of the semi‐impermeable layer (clayey sands) between the fine to coarse
sand and the sandstones causes the water circulation of these two sectors to continuously
connect [33]. There are places where the coarse sands lie directly on top of the clay layer,
developing semi‐confined conditions[18,34]. Nogueira et al. [35] showed that the aquifer
system in Maputo city is prone to contamination. Concerning regional and local geology,
the Hulene‐B dump is inserted into the Mesocenozoic sedimentary basin of southern
Mozambique [17], and is situated in a contact zone of two lithologies, Ponta Vermelha
Formation and Malhazine (Figure 1b) on a gentle dune slope with east–west orientation
Environments 2022, 9, 19 3 of 13

[36]. The Ponta Vermelha Formation (TPv) dates from the upper Pliocene to the lower
Plistocene and is composed, in the upper part, of ferruginous sandstones and red silty
sands, which gradually change to yellow and whitish sands [18,37]. At the surface, this
unit presents a red colour, and poorly consolidated sands may appear [16]. The Malhazine
Formation (QMa) dates to the upper Pleistocene and consists of coarse to fine, poorly
consolidated, sands with whitish to reddish colours, fixed by vegetation because of
successive consolidation processes [18,37]. The soils in the surrounding of the Hulene‐B
dump have been classified as sandy dune [36]. The predominant climate is of subtropical
type, with two seasons: (a) hot and rainy period from December to March with more than
60% of the annual precipitation, with the highest concentration of precipitation in January
(with an average of 125 mm) and (b) dry and cold season from April to September with
lower temperatures in June and July, as well as a weak and irregular precipitation, whose
minimum values are recorded in August (12 mm). The average annual precipitation is
789.2 mm [38,39]. The prevailing winds are SE [40].

Figure 1. (a) Location, geophysical survey lines and environment context (b) geology and
topography. Adapted by Oliveira et al. [36].

2.2. Geophysical Studies (Electrical Resistivity)


In recent years, the awareness of the increasingly complex environmental issues has
boosted the use of geophysical prospection methods, in a quite successful way, for the
study of complex environments, such as urban areas [41,42]. Among the geophysical
methods most employed in environmental studies, electrical resistivity has been pointed
out as relevant [19,20], and it has been prominent in the study of environmental problems
in soil and groundwater, mainly in the location of buried hazardous waste, contamination
from different sources, and planning of safe sites for deposit of industrial and domestic
waste [24,43], being widely used to identify areas of heavy metal contamination,
contamination plumes [10], groundwater [20,44], and lithological variations [20].
Leachates cause, when they meet geological materials, natural electrical resistivity of the
material to decrease due to the high concentration of dissolved metal ions, creating
anomalous resistive zones [45,46].
The resistivity method is based on the electric current injected into the ground
through a pair of electrodes (A and B‐current electrodes) and the resulting potential
difference between another pair of electrodes (M and N potential electrodes) [19,45]. The
Environments 2022, 9, 19 4 of 13

ground resistivity is calculated from the distances between the electrodes, applied current
and measured potential difference, based on the Law of Ohm [47]. The apparent resistivity
of the soil can be determined based on the known differences between the electric field
potential (∆V) and the current (I), and the distance between the electrodes [21].
The resistivity is given by the equation:
∆𝑉
𝜌𝑎 𝑘 (1)
𝐼
where: ρa—resistivity of a bedrock, I—intensity of current applied to the soil by electrodes
AB (mA), ∆V—differential potential between electrodes MN (mV), k—geometrical
coefficient of electrode positioning (m).
The geometrical factor k is dependent on the distribution geometry of the electrodes,
as follows:
2𝜋
𝐾 (2)
1 1 1 1
𝐴𝑀 𝐵𝑀 𝐴𝑁 𝐵𝑁
where AM, BM, AN and BN represent the geometrical distance between the electrodes A
and M, B and M, A and N, and B and N, respectively [1,48].
The electrical resistivity of the terrains is a characteristic closely linked to the type,
nature, and state of alteration of the geological formations [22,49]. Thus, the method
allows for: (1) the identification of the lithology of landfill subsoil; (2) the determination
of the groundwater table depth; (3) the determination of the distribution of the
contamination zones and the direction of the pollutant migration; (4) the evaluation of
waste thickness disposed at a landfill site [21,28,50].
In this research, 8 electrical resistivity profiles were performed in two different
seasons (Figure 1a): 4 profiles in January 2020, corresponding to the hot and rainy period,
and in May 2021, 4 more profiles were executed overlapping those of the first campaign
on the western edge of the dump.
For the data acquisition, a resistor ABEM SAS 4000 was used, including 4 rollers of
100 m cables with 21 outlets that connect to the same number of electrodes. The layout
produced by this sequence of cables (100 m and 21 outputs) corresponds to the standard
of the reading program hosted by the resistivimeter LUND Imaging System. For data
acquisition, it employed a 50 Hz current frequency, using a multigradient protocol
(GRAD4LX8 and GRAD4SX8); the GRAD4LX8 was selected because it provides dense
coverage on the nearby surface and adopts the Wenner–Schlumberger protocol [51,52].
The electrode spacing for data acquisition was 5 m. All the electrode take‐outs were
connected in the GRAD4S8 protocol (Figure S1). The resistivimeter automatically switches
the electrodes to serve as current or potential pairs. After the readings, the data were
transferred to the resistivimeter, after storing a minimum of 3 and a maximum of 6
readings, to obtain the lowest average error between readings. The inversion of the
electrical resistivity data obtained in the 8 lines was performed based on the standards
defined in the software RES2DINV3.59.106, namely, application of the smoothness
constraint method in the resistivity values of the final model, calculation of the Jacobian
matrix in each iteration, and Gauss–Newton optimization method [19,42,49]. The
interpretation of the profiles was based on the direction of each profile and the length of
the profiles (400 m). The analysis zones were delimited as Zone 1 from 0–200 m (Z1) and
Zone 2 from 200–400 m (Z2).

3. Results and Discussion


3.1. Geophysical Studies—Electrical Resistivity
The interpretation of the electrical resistivity models allowed for the understanding
of the leachate formation areas, dynamics, and dispersion of contamination plumes in the
groundwater as well as for comparing the variations of the resistive anomalies (2020–
Environments 2022, 9, 19 5 of 13

2021). The following anomalous areas were distinguished: (I) areas of possible leachate
formation and enrichment; (II) contamination plumes in subsurface and groundwater.

3.1.1. Profile 1: 2020 (a) and 2021 (b)


The two profiles (a) and (b) extend in the S–N direction and are parallel to the western
boundary of the dump (Figure 2).

Figure 2. Electrical resistivity model of profile 1, 2020, (a) and 2021 (b).

Zone 1—Along the first 200 m of both profiles, no significant changes are noted near
the surface, and in general, the resistivity is higher and associated with the rubble and
debris of old houses that were built in this space until 2018 [49]. At depth, slight
differences in resistive layers are noted, which may be associated with the effect of
variation in rainfall and moisture, more abundant in (a) and less in (b). The zone of
resistivity < 19.2 Ω∙m in (a) and < 15.4 Ω∙m (b), which we consider as the leachate
concentration zone, resulting from the vertical migration of leachate from the surface
washing of the dump by precipitation being channelled in the unprotected ditch and
parallel to these profiles (Figure 3b). Wu et al. [50] and Ololade et al. [51] showed that
leachate flow in non‐isolated areas can cause migration to great depths and groundwater
contamination. The leachate accumulation zones, < 19.2 Ω∙m (a) and < 15.4 Ω∙m (b), show
variable thicknesses, which may be associated with the effect of the relatively saturated
lower layers, < 16.1 to 11.7 Ω∙m (a), being more extensive in summer, which causes slow
vertical migration at depth and the production of thick plumes at great depth, < 8.46 Ω∙m
in profile (a) and < 8.56 Ω∙m in profile (b). The leachate migration layers can be considered
as contaminated, < 16.1–11.7 Ω∙m (a) and < 13.2–10.7 Ω∙m (b). At great depth, at the
southern end of both profiles, there is a localised anomalous zone that we interpret as
being influenced by groundwater contamination by plumes resulting from vertical
leachate migration.
Zone 2—In both profiles, anomalous zones are evident near the surface at 280 m
onwards, which are more extensive in profile (b), representing residues humidified by
surface water (leachate producer) < 15.4 Ω∙m, which are more visible in the north of both
profiles and more extensive in profile (b). The clearly visible plumes in both profiles, in
the surface water, correspond to the contamination in the natural receiving basin by
surface leachates and plumes mobilizing at depth in the E–W direction, described in
profile 2.
Environments 2022, 9, 19 6 of 13

Figure 3. Geophysical surveys: (a) profile 1 in 2020—the arrow indicates the surface
leachate concentration ditch parallel to the profile 1; (b) drainage ditch with uninsulated
surface leachate 2021; (c) southern section of profile 1 in 2021.

3.1.2. Profile 2 2020 (a) and 2021(b)


Profile 2, with a west–east orientation, and northern boundary of the dump (Figure
4).

Figure 4. Electrical resistivity model of profile 2, 2020 (a) and 2021 (b).

Zone 1—From 0 to 80 m surface, in profile (a) we can notice a marked variation of


resistivities. The zones of low resistivity (< 24.5 Ω∙m), we interpret as a wet and leached
waste production zone, given the new solid waste depositions recorded at this point to
the west of the basin. The same profile section (b) show high resistivities, given the scarcity
of precipitation and low ambient temperature. The high temperatures, humidity and age
of the dump are primary factors in decomposition and leachate production [17,44,53].
From 8.35 m depth downwards, in profile (a), we note an extensive saturated zone that
extends in the northern direction and connects to an extensive subsurface flow system. In
turn, in profile (b), this saturated zone appears more confined and with an expressive
concentration of anomalous values < 4.96 Ω∙m (b), which we interpret as a plume
migrating horizontally in summer in the east–west direction and being confined at this
point. Cendón et al. [16] described the aquifer system of this region as semi‐confined that
seasonally binds continuously, a fact that is visualized in these bands of the two profiles
and that are associated with the transfer of possible plumes at depth. From 80 to 200 m,
superficial in profile (a), we verify the alternation of high and low resistivities. The high
resistivities, we interpret as compact material on the surface, which alternates between
debris, rubble, and old house debris. From 160 to 200 m (a), we observe surface waters of
the natural leachate reception basin, which are constantly enriched by the surface
Environments 2022, 9, 19 7 of 13

leachates. In this section at a depth of 8.35 to 47.7 m (a), there is an extensive plume of
contamination arising from the large mass of wet waste and producing leachates < 24.5
Ω∙m (a) that migrate horizontally in an east–west direction. Complex mechanisms of
leachate movement from the surface and at depth are also observed. The vertical and
horizontal movement of the leachate produces an extensive plume migrating E–W, < 6.21
Ω∙m (a). In profile (b), from 80 to 140 m, the resistivity is higher along a larger area than
in (a), given the significant reduction in the extent of surface water responsible for the
decomposition of the waste mass and consequent decrease in resistivity. At the depth of
the same section, confined groundwater receives leachate, which moves horizontally from
E–W < 16.8 Ω∙m (b) and a vertical migration between 142.5 to 147.5 m. The contamination
plume < 7.9 Ω∙m (b) at this point is quite pronounced and may indicate a high level of
contamination, given the reduced dilution and migration environment.
Zone 2—In (a) between 200 m to 315 m, it shows alternating resistivities between the
less moist waste mass < 20.7 Ω∙m to saturated zones < 10.6 Ω∙m (b) which represent pits
and small surface depressions enriched by surface water, and the same characteristics are
noted in (b), but with lower moisture extent. From 320 m onwards, in profile (a), an
increase in resistivity > 30.8 Ω∙m is noted which corresponds to waste mass mixed with
less moist soils and resistivity > 52.8 Ω∙m which represents compacted dry waste. In profile
(b), to the same extent, the increase in resistivity is much more noticeable, which indicates
dry waste and soil (< 33.4 Ω∙m) and compacted waste (> 33.4 Ω∙m). At depth, in both
profiles, the large mass of waste, < 30.94 Ω∙m (a) and < 20.74 Ω∙m (b), gains successive
moisture in (a), establishing an extensive area of leachate production < 24.5 Ω∙m (a),
whereas in (b), it is confined (< 16.8 Ω∙m). These differences are the result of the variation
in precipitation and temperature in the two seasons studied, which are responsible for the
variation in waste mass decomposition, production, and migration of leachate [54,55].

3.1.3. Profile 3 2020 (a) and 2021 (b)


Profile 3, with an S–N orientation northwest of the dump (Figure 5).

Figure 5. Electrical resistivity model of profile 3 in 2020 (a) and 2021 (b).

The execution area of this profile corresponds to the strip that temporarily floods
with run‐off water from the dump. It was executed to understand the spatial dynamics of
possible contamination plumes in the northern direction of the surroundings of the dump
(reception basin). At 40 m, in both profiles, the influence of moisture in a localized band,
at depth, is noted, evidenced by resistivity < 23.5 Ω∙m in profile (a) and 21.6 Ω∙m (b). In
the first profile, the strip occupies a relatively larger area due to the abundant precipitation
in this period that infiltrates to the deeper layers. At 140 m depth, there is an anomalous
zone in both profiles, which we interpret as a plume of contamination, < 7.18 (a) and < 7.34
Environments 2022, 9, 19 8 of 13

Ω∙m (b). We consider that this anomaly corresponds to the northern limit of the large
plume described in profile 2. From 160 m onwards, in both profiles at all depths, the
resistivity tends to be equal, and we consider as typical of local strata and anomalous
(positively) zones, > 38.2 Ω∙m (a) and > 33.7 (b), corresponding to compacted soils or
rubble, given the strong movement of cars at these points. The similarity of the data in
both profiles and the absence of low depth anomalous zones show the decreasing effect
of the contamination plumes on the subsoil and the subsurface environment, as one move
away from the dump to the North, due to the natural process of attenuation by dilution
and dispersion of the plumes in the natural receiving basin. Similar situations have been
described in areas around several dumpsites in Africa, Morocco by El Mouine et al. [56]
and Touzani et al. [39], Nigeria by Fatoba et al. [53] and Burkina Faso by Barry et al. [54].

3.1.4. Profile 4 in 2020 (a) and 2021 (b)


Profile 4 has SW–NE orientation (a), and the NE–SW (b) profile both overlap (Figure
6).

Figure 6. Electrical resistivity model of profile 4 in 2020 (a) and 2021 (b).

Zone 1—Profile 4 (a) from the beginning to 120 m shallow shows zones with
anomalous resistivities. The resistivity < 13.7 Ω∙m was interpreted as a zone of leachate
production and dispersion, resulting from new waste deposits in the southwest of the
natural basin. These leachates migrate horizontally in the NE–SW direction and vertically
until they mix with groundwater. In profile 4 (b), these anomalies occupy a large area in
the dry period and extend over an area of about 240 m surface as well as at depth, < 16.3
Ω∙m. The extensive anomalous area at depth in profile 4 (b) (< 16.3 Ω∙m) may be associated
with a vertical and continuous migration of leachate accumulated at the end of the rainy
season in the southeast of natural reception basin.
Zone 2—In profile 4 (a) from 200 to 400 m, the surface resistivities are generally
higher and we interpret them as plastic waste, rubble, and debris from old, buried houses.
The same occurs in profile 4b, from 140 m to the end of the profile. The resistivity
corresponding to possible contamination plumes were interpreted as < 6.42 Ω∙m in (a),
<8.9 Ω∙m in (b), which at great depth, did not show great changes in the two seasons of
the years studied. This reality can be associated with the local aquifer system, which is
described as semi‐confined in clay layers [16].

3.2. Spatial Distribution of Possible Leachate Plumes (2020–2021)


The spatial distribution of the plumes in the study period (2020–2021) (Figure 7),
shows predominantly an east–west movement of the deep contamination plumes and
surface leachates in profiles 1 and 2, which mix with the surface waters of the natural basin
Environments 2022, 9, 19 9 of 13

to the west and migrate vertically to the groundwater, causing resistive anomalies at a
larger scale in winter (1). The same flow direction was described by Bernardo et al. [18].
The flow is conditioned by the arrangement of the local topography (soft dune E–W),
which collects all the leachate from the subsurface compression of the waste, since it is at
a higher altitude than the local surface. Another relevant factor in this direction of
contaminant flow is the recent construction of leachate drainage channels, which create
natural leachate flows in the E–W direction (Figure 4b). However, the dilution process of
leachate plumes shows a varied dispersion in the natural reception basin and
groundwater (Figure 7. < 4.26 – < 8.5 Ω∙m), suggesting that contamination migrates in
several directions. Similar anomalous values were interpreted as plumes in groundwater
by Harjito et al. [55] 3–9 Ω∙m near a Bantul waste dump in Indonesia, Bichet et al. [22] refer
between 5–12 Ω∙m at a landfill in Belfort (France), and Ugbor et al. [43] registered values
of 3.12–8.7 Ω∙m in the surroundings of the Onitsha waste dump in southeast Nigeria. To
the southwest of the natural basin reception, there is a subplume of contamination
resulting from new depositions producing leachate that migrates to the southwest during
the wet season and in all directions during the dry season (profile 4). The spatialization of
the plumes shows that as one moves away from the dump, the anomalies tend to
disappear (profile 3), suggesting the attenuation of the contamination by the local
lithology. The role of lithology and groundwater in the attenuation of contamination by
leachate plumes in the surrounding of waste dumps has been described in many studies,
for example by Fatoba et al. [53] and Biosca et al. [57].

Figure 7. Possible flow directions of contamination plumes 2020–2021, modified from Bernardo et
al. [18].
Environments 2022, 9, 19 10 of 13

4. Conclusions
The application of the electrical resistivity method to the study of the dynamics of
contamination plumes around the Hulene‐B waste dump has proven to be efficient in
identifying anomalous areas of low resistivity that we consider as leachate production and
contamination plume migration zones in lithological substrates and surface and
groundwater.
The analysis of the dynamics of the possible contamination plumes showed that
summer (2020) occupies an extensive area and with strong horizontal and vertical
migration, which may be associated with the greater production of leachates by the greater
decomposition of waste in the hot and rainy period (Profiles 1a, 2a). Conversely, in winter
(2021), the contamination plumes were thick with a predominantly vertical movement,
which allows for a large vertical migration to great depths, causing extensive subsurface
anomalies in the lithologies above the groundwater (Profiles 1b and 4b). Unusually,
Profile 3 showed that the southern part is partially affected by the large plume described
in profile 2, without significant variations in both analysed seasons.
The contamination plumes in the two study seasons (2020–2021) show
predominantly an E–W movement (profiles 1 and 2). In profile 3, the movement was in
the S–N direction and was influenced by the diffusion of the large plume described in
profile 2. In profile 4, the plume assumes a NE–SW movement, and its enrichment is
associated with the new deposition at the SW limit of the natural leachate reception basin.
The four overlapping profiles show that leachate enrichment zones play a relevant role in
the direction of flow and migration of contamination plumes in lithologies, groundwater
and surface water.
The spatialized electrical resistivity models at the studied stations show that the
anomalies in the subsurface environment decrease as one moves away from the dump to
the north (profile 3), which reveals the attenuating effect of groundwater and lithological
substrates. Thus, the resistivity models have proven to be efficient in assisting in the
adoption of structural measures for monitoring the possible leachate contamination flows
in the lithologies, surface and groundwater around the Hulene‐B waste dump, Maputo,
Mozambique. Studies are ongoing to quantify the chemical contamination of groundwater
and surface water.

Supplementary Materials: The following are available online


www.mdpi.com/article/10.3390/environments9020019/s1. Figure S1 (a) Resistivimeter applied in
data acquisition, (b) data acquisition in January 2020 (profile 2), (c) data acquisition in profile 2 in
May 2021, (d) southern section of profile 1 in 2021, (e) central section of profile 1 in 2020, (f) western
section of profile 2 in 2020, (g) northern section of profile 3, (h) electrode connection in the northwest
section of profile 4 in 2021.
Author Contributions: Conceptualization, B.B. and F.R.; methodology, B.B.; software, B.B.;
validation, B.B., C.C., and F.R.; formal analysis, F.R.; investigation, B.B.; resources, B.B.; data
curation, B.B.; writing—original draft preparation, B.B.; writing—review and editing, B.B. and F.R.;
visualization, B.B.; supervision, F.R. and C.C.; project administration, F.R.; funding acquisition, B.B.
All authors have read and agreed to the published version of the manuscript.
Funding: This work was partially supported by GeoBioTec (UIDB/04035/2020) Research Centre,
funded by FEDER funds through the Operational Program Competitiveness Factors COMPETE and
by National funds through FCT. The first author acknowledges grants from the Portuguese Institute
Camões and FNI (Investigation National Fund—Mozambique).
Institutional Review Board Statement: Not applicable.
Informed Consent Statement: Not applicable.
Data Availability Statement: Not applicable.
Acknowledgments: Not applicable.
Conflicts of Interest: The authors declare no conflict of interest
Environments 2022, 9, 19 11 of 13

References
1. Brahmi, S.; Baali, F.; Hadji, R.; Brahmi, S.; Hamad, A.; Rahal, O.; Zerrouki, H.; Saadali, B.; Hamed, Y. Assessment of groundwater
and soil pollution by leachate using electrical resistivity and induced polarization imaging survey, case of Tebessa municipal
landfill, NE Algeria. Arab. J. Geosci. 2021, 14, 1–13, https://doi.org/10.1007/s12517‐021‐06571‐z.
2. Ooi, J.; Woon, K.S. Simultaneous Greenhouse Gas Reduction and Cost Optimization of Municipal Solid Waste Management
System in Malaysia., Chem. Eng. Trans. 2021, 83, 487–492, https://doi.org/10.3303/CET2183082.
3. Ferronato, N.; Torretta, V. Waste Mismanagement in Developing Countries: A Review of Global Issues. Int. J. Environ. Res. Public
Health 2019, 16, 1060, https://doi.org/10.3390/ijerph16061060.
4. Nanda, S.; Berruti, F. Municipal solid waste management and landfilling technologies: a review. Environ. Chem. Lett. 2020, 19,
1433–1456, https://doi.org/10.1007/s10311‐020‐01100‐y.
5. Marino, A.; Pariso, P. Comparing European countriesʹ performances in the transition towards the Circular Economy. Sci. Total.
Environ. 2020, 729, 138142, https://doi.org/10.1016/j.scitotenv.2020.138142.
6. Koliyabandara, S.; Asitha, T.C.; Sudantha, L.; Siriwardana, C. Assessment of the impact of an open dumpsite on the surface
water quality deterioration in Karadiyana, Sri Lanka. Environ. Nanotechnology, Monit. Manag. 2020, 14, 100371,
https://doi.org/10.1016/j.enmm.2020.100371.
7. Morita, A.K.; Ibelli‐Bianco, C.; Anache, J.A.; Coutinho, J.V.; Pelinson, N.S.; Nobrega, J.; Rosalem, L.M.; Leite, C.M.; Niviadonski,
L.M.; Manastella, C.; et al. Pollution threat to water and soil quality by dumpsites and non‐sanitary landfills in Brazil: A review.
Waste Manag. 2021, 131, 163–176, https://doi.org/10.1016/j.wasman.2021.06.004.
8. Wijekoon, P.; Koliyabandara, P.A.; Cooray, A.T.; Lam, S.S.; Athapattu, B.C.; Vithanage, M. Progress and prospects in mitigation
of landfill leachate pollution: Risk, pollution potential, treatment and challenges. J. Hazard. Mater. 2021, 421, 126627,
https://doi.org/10.1016/j.jhazmat.2021.126627.
9. Jayawardhana, Y.; Kumarathilaka, P.; Herath, I.; Vithanage, M. Municipal Solid Waste Biochar for Prevention of Pollution from
Landfill Leachate. 2016, 117–148, https://doi.org/10.1016/B978‐0‐12‐803837‐6.00006‐8.
10. Wang, F.; Song, K.; He, X.; Peng, Y.; Liu, D.; Liu, J. Identification of Groundwater Pollution Characteristics and Health Risk
Assessment of a Landfill in a Low Permeability Area. Int. J. Environ. Res. Public Heal. 2021, 18, 7690,
https://doi.org/10.3390/ijerph18147690.
11. Kumar, G.; Reddy, K.R.; McDougall, J. Numerical modeling of coupled biochemical and thermal behavior of municipal solid
waste in landfills. Comput. Geotech. 2020, 128, 103836, https://doi.org/10.1016/j.compgeo.2020.103836.
12. Khattak, S.A.; Rashid, A.; Tariq, M.; Ali, L.; Gao, X.; Ayub, M.; Javed, A. Potential risk and source distribution of groundwater
contamination by mercury in district Swabi, Pakistan: Application of multivariate study. Environ. Dev. Sustain. 2020, 23, 2279–
2297, https://doi.org/10.1007/s10668‐020‐00674‐5.
13. Iddrisu, T.I.; Debrah, K.D. Consequences of Poor Landfill Management on the People of Gbalahi in the Sagnarigu Municipality
of Northern Ghana. J. Geosci. Environ. Prot. 2021, 9, 211–224, https://doi.org/10.4236/gep.2021.98014.
14. Tvedten, I.; Candiracci, S. Flooding our eyes with rubbish: urban waste management in Maputo, Mozambique. Environ. Urban.
2018, 30, 631–646, https://doi.org/10.1177/0956247818780090.
15. Serra, C. Da Problemática Ambiental à Mudança: Rumo à um Mundo Melhor; Editora Escolar: Lisboa, Portugal, 2012.
16. Cendón, D.I.; Haldorsen, S.; Chen, J.; Hankin, S.; Nogueira, G.; Momade, F.; Achimo, M.; Muiuane, E.; Mugabe, J.; Stigter, T.Y.
Hydrogeochemical aquifer characterization and its implication for groundwater development in the Maputo district,
Mozambique. Quat. Int. 2019, 547, 113–126, https://doi.org/10.1016/j.quaint.2019.06.024.
17. Vicente, E.M.; Jermy, C.A.; Schreiner, H.D. Urban geology of Maputo, Mocambique. Geol. Soc. 2006, 338, 1–13. Available online:
https://citeseerx.ist.psu.edu/viewdoc/download?doi=10.1.1.606.7220&rep=rep1&type=pdf. (Accessed on 28 January 2020).
18. Bernardo, B.; Candeias, C.; Rocha, F. Application of Geophysics in geo‐environmental diagnosis on the surroundings of the
Hulene‐B waste dump, Maputo, Mozambique. J. Afr. Earth Sci. 2021, 185, 104415, https://doi.org/10.1016/j.jafrearsci.2021.104415.
19. Helene, L.P.I.; Moreira, C.A. Analysis of Leachate Generation Dynamics in a Closed Municipal Solid Waste Landfill by Means
of Geophysical Data (DC Resistivity and Self‐Potential Methods). Pure Appl. Geophys. PAGEOPH 2021, 178, 1355–1367,
https://doi.org/10.1007/s00024‐021‐02700‐7.
20. Kayode, J.S.; Arifin, M.H.; Nawawi, M.N.M. Characterization of a Proposed Quarry Site using Multi‐Electrode Electrical
Resistivity Tomography. Sains Malays. 2019, 48, 945–963, https://doi.org/10.17576/jsm‐2019‐4805‐03.
21. Lau, A.M.P.; Ferreira, F.J.F.; Stevanato, R.; Filho, E.F.D.R. Geophysical and physicochemical investigations of an area
contaminated by tannery waste: a case study from southern Brazil. Environ. Earth Sci. 2019, 78, 1–16,
https://doi.org/10.1007/s12665‐019‐8536‐1.
22. Bichet, V.; Grisey, E.; Aleya, L. Spatial characterization of leachate plume using electrical resistivity tomography in a landfill
composed of old and new cells (Belfort, France). Eng. Geol. 2016, 211, 61–73, https://doi.org/10.1016/j.enggeo.2016.06.026.
23. Bernstone, C.; Dahlin, T.; Ohlsson, T.; Hogland, H. DC‐resistivity mapping of internal landfill structures: two pre‐excavation
surveys. Environ. Earth Sci. 2000, 39, 360–371, https://doi.org/10.1007/s002540050015.
24. Nassereddine, M.; Rizk, J.; Nasserddine, G. Soil Resistivity Structure and Its Implication on the Pole Grid Resistance for
Transmission Lines. Int. J. Electr. Comput. Eng. 2013, 7, 19–23.
25. Palalane, j.; Segala, I.; Opressa, I. Urbanização e desenvolvimento municipal em Moçambique: gestão de resíduos sólidos.
Instituto Brasileiro de Administração Municipal, Área de desenvolvimento Urbano e Meio Ambiente, RJ, 2008, 12. Available
Environments 2022, 9, 19 12 of 13

online: https://www.scribd.com/document/419123335/Gestao‐de‐Residuos‐Solidos‐Em‐Mocambique. (Accessed on 05 August


2021)
26. Ferrão. D.A.G. Evaluation of removal and disposal of solid waste in Maputo City, Mozambique, Master’s thesis, University of
Cape Town, Cape Town, South Africa, 2006. Available online: http://hdl.handle.net/11427/4851 (accessed on 30 August 2021.)
27. Wang, X.‐X.; Chang, Y.‐B.; Deng, J.‐Z.; Chen, J.‐S. 3D spatial distribution of old landfills and groundwater pollution from
electrical resistivity tomography with fuzzy set theory. Explor. Geophys. 2021, 1–9,
https://doi.org/10.1080/08123985.2021.1917292.
28. Dhakate, R.; Mogali, N.J.; Modi, D. Characterization of proposed waste disposal site of granite quarry pits near Hyderabad
using hydro‐geophysical and groundwater modeling studies. Environ. Earth Sci. 2021, 80, 1–20, https://doi.org/10.1007/s12665‐
021‐09821‐1.
29. Adamo, N.; Al‐Ansari, N.; Sissakian, V.; Laue, J.; Knutsson, S. Geophysical Methods and their Applications in Dam Safety
Monitoring. J. Earth Sci. Geotech. Eng. 2020, 291–345, https://doi.org/10.47260/jesge/1118.
30. Sallwey, J.; Hettiarachchi, H.; Hülsmann, S. Challenges and opportunities in municipal solid waste management in
Mozambique: a review in the light of nexus thinking. AIMS Environ. Sci. 2017, 4, 621–639,
https://doi.org/10.3934/environsci.2017.5.621.
31. Bandeira, J.; Paula, S. The Maputo Bay Ecosystem. WIOMSA. Maputo: Western Indian Ocean Marine Science Association
(WIOMSA), Zanzibar Town, Tanzania, 2014, 427, Available online: https://www.biofund.org.mz/wp‐
content/uploads/2019/02/1550049666‐Je02‐036.pdf. (Accessed on 30 July 2021).
32. INAM [Mozambique National Meteorological Institute]. Available online: https://www.inam.gov.mz/index.php/pt/ (Accessed
on 28 January 2020)
33. Muchimbane, A. B. D. Estudo dos indicadores da contaminação das aguas subterrâneas por sistemas de saneamento in Situ –
Distrito Urbano 4, Cidade de Maputo, Moçambique. Dissertação de Mestrado. USP Instituto de Geociências, 2010.
https://doi.org/10.11606/D.44.2010.tde‐06052010‐153107.
34. CIAT World Bank. Climate‐Smart Agriculture in Mozambique. Clim. Agric. Mozambique, 2017, 1–25.
35. Nogueira, G.; Stigter, T.; Zhou, Y.; Mussa, F.; Juizo, D. Understanding groundwater salinization mechanisms to secure
freshwater resources in the water‐scarce city of Maputo, Mozambique. Sci. Total Environ. 2018, 661, 723–736,
https://doi.org/10.1016/j.scitotenv.2018.12.343.
36. Oliveira, J.T.; Dias R.P.; Pereira, A. Cooperaça o entre Portugal e Moçambique na Á rea das Geocie ncias 1986‐2012. Maputo, 23‐
26, ISBN: 978‐989‐675‐026‐8. Available online: https://repositorio.lneg.pt/bitstream/10400.9/3336/1/35652.pdf. (Accessed on 28
January 2020)
37. Mama, C.N.; Nnaji, C.C.; Nnam, J.P.; Opata, O.C. Environmental burden of unprocessed solid waste handling in Enugu State,
Nigeria. Environ. Sci. Pollut. Res. 2021, 28, 19439–19457, https://doi.org/10.1007/s11356‐020‐12265‐y.
38. Udosen, N.I. Geo‐electrical modeling of leachate contamination at a major waste disposal site in south‐eastern Nigeria. Model.
Earth Syst. Environ. 2021, 1–10, https://doi.org/10.1007/s40808‐021‐01120‐9.
39. Touzani, M.; Mohsine, I.; Ouardi, J.; Kacimi, I.; Morarech, M.; El Bahajji, M.; Bouramtane, T.; Tiouiouine, A.; Yameogo, S.; El
Mahrad, B. Mapping the Pollution Plume Using the Self‐Potential Geophysical Method: Case of Oum Azza Landfill, Rabat,
Morocco. Water 2021, 13, 961, https://doi.org/10.3390/w13070961.
40. Dos Muchangos, A. Paisagens e Regiões Naturais, Maputo. Editora Escolar: Lisboa, Portugal, 1999, 5–163.
41. Arifin, M. H.; Kayode, J. S.; Khairel, M., Abdul, I.; Abdullah, M.; Shahidah M.N.; Azmi, A. Environmental hazard assessment of
industrial and municipal waste materials with the applications of RES2‐D method and 3‐D Oasis Montaj modeling: A case study
at Kepong, Kuala Lumpur, Peninsula Malaysia, J. Hazard. Mater., 2020, 406, 124282,
https://doi.org/10.1016/j.jhazmat.2020.124282.
42. Geotomo. RES2DINV ver. 3.59 ‐ Rapid 2‐D Resistivity & IP Inversion using the Least‐Squares Method Wenner (α,β,γ), Dipole‐
Dipole, Inline Pole‐Pole, Pole‐ Dipole, Equatorial Dipole‐Dipole, Offset Pole‐Dipole, Wenner‐Schlumberger, Gradient and Non‐
Conventional Arrays, 2010, 1–148. Available online: http://epsc.wustl.edu/~epsc454/instruction‐sheets/Res2dinv03.59.pdf.
(Accessed on 28 January 2020)
43. Ugbor, C.C.; Ikwuagwu, I.E.; Ogboke, O.J. 2D inversion of electrical resistivity investigation of contaminant plume around a
dumpsite near Onitsha expressway in southeastern Nigeria. Sci. Rep. 2021, 11, 1–14, https://doi.org/10.1038/s41598‐021‐91019‐3.
44. Nta, S. A.;. Ayotamuno, M. J.; Igoni, A. H.; Okparanma, R. N. Leachate Characterization from Municipal Solid Waste Dump
Site and Its Adverse Impacts on Surface Water Quality Downstream ‐ Uyo Village Road, Akwa Ibom State ‐ Nigeria,” J. Eng.
Res. Reports, 2020, 13, 11–19, https://doi.org/10.9734/jerr/2020/v13i217096
45. Qiu, L.; Yang, Y.; Ma, L.; Qiao, J. Research on the Electrical Resistivity Characteristics of Statue Remolded Soil. IOP Conf. Series:
Earth Environ. Sci. 2021, 692, 042076, https://doi.org/10.1088/1755‐1315/692/4/042076.
46. Ejiogu, B.C.; Opara, A.; Nwosu, E.I.; Nwofor, O.K.; Onyema, J.C.; Chinaka, J.C. Estimates of aquifer geo‐hydraulic and
vulnerability characteristics of Imo State and environs, Southeastern Nigeria, using electrical conductivity data. Environ. Monit.
Assess. 2019, 191, 238, https://doi.org/10.1007/s10661‐019‐7335‐1.
47. Koda, E.; Tkaczyk, A.; Lech, M.; Osiński, P. Application of Electrical Resistivity Data Sets for the Evaluation of the Pollution
Concentration Level within Landfill Subsoil. Appl. Sci. 2017, 7, 262, https://doi.org/10.3390/app7030262.
Environments 2022, 9, 19 13 of 13

48. Vasantrao, B.M.; Bhaskarrao, P.J.; Mukund, B.A.; Baburao, G.R.; Narayan, P.S. Comparative study of Wenner and Schlumberger
electrical resistivity method for groundwater investigation: a case study from Dhule district (M.S.), India. Appl. Water Sci. 2017,
7, 4321–4340, https://doi.org/10.1007/s13201‐017‐0576‐7.
49. VOA. Desabamento de Lixeira Deixa 17 Mortos em Maputo. Voice of America News, 2018. Available online:
https://www.voaportugues.com/a/desabamento‐lixeira‐17‐mortos‐maputo/4260624.html (accessed on 30 September 2021).
50. Wu, Q.; Hu, W.; Wang, H.; Liu, P.; Wang, X.; Huang, B. Spatial distribution, ecological risk and sources of heavy metals in soils
from a typical economic development area, Southeastern China, Sci. Total Environ., 2021, 780, 146557,
https://doi.org/10.1016/j.scitotenv.2021.146557.
51. Ololade, O.O.; Mavimbela, S.; Oke, S.A.; Makhadi, R. Impact of Leachate from Northern Landfill Site in Bloemfontein on Water
and Soil Quality: Implications for Water and Food Security. Sustainability 2019, 11, 4238, https://doi.org/10.3390/su11154238.
52. Chaudhary, R.; Nain, P.; Kumar, A. Temporal variation of leachate pollution index of Indian landfill sites and associated human
health risk. Environ. Sci. Pollut. Res. 2021, 28, 28391–28406, https://doi.org/10.1007/s11356‐021‐12383‐1.
53. Fatoba, J.O.; Eluwole, A.B.; Sanuade, O.A.; Hammed, O.S.; Igboama, W.N.; Amosun, J.O. Geophysical and geochemical
assessments of the environmental impact of Abule‐Egba landfill, southwestern Nigeria. Model. Earth Syst. Environ. 2020, 7, 695–
701, https://doi.org/10.1007/s40808‐020‐00991‐8.
54. Barry, A.; Yameogo, S.; Ayach, M.; Jabrane, M.; Tiouiouine, A.; Nakolendousse, S.; Lazar, H.; Filki, A.; Touzani, M.; Mohsine, I.
Mapping Contaminant Plume at a Landfill in a Crystalline Basement Terrain in Ouagadougou, Burkina Faso, Using Self‐
Potential Geophysical Technique. Water 2021, 13, 1212, https://doi.org/10.3390/w13091212.
55. Harjito, S..; Gunawan, T.; Maskuri, M. Underground leachate distribution based on electrical resistivity tomography in
Piyungan landfill, Bantul, Indones. J. Geogr., 2018, 50, 34–40, https://doi.org/10.22146/ijg.18315.
56. El Mouine, Y.; El Hamdi, A.; Morarech, M.; Kacimi, I.; Touzani, M.; Mohsine, I.; Tiouiouine, A.; Ouardi, J.; Zouahri, A.; Yachou,
H.; Dakak, H. Landfill Pollution Plume Survey in the Moroccan Tadla Using Spontaneous Potential. Water 2021, 13, 910.
https://doi.org/10.3390/w13070910.
57. Biosca, B.; Arévalo‐Lomas, L.; Izquierdo‐Díaz, M.; Díaz‐Curiel, J. Detection of chlorinated contaminants coming from the
manufacture of lindane in a surface detritic aquifer by electrical resistivity tomography. J. Appl. Geophys. 2021, 191, 104358,
https://doi.org/10.1016/j.jappgeo.2021.104358.
232
233

13.2.2 Appendix 3. Paper 2


234
water
Article
Integration of Electrical Resistivity and Modified DRASTIC
Model to Assess Groundwater Vulnerability in the Surrounding
Area of Hulene-B Waste Dump, Maputo, Mozambique
Bernardino Bernardo 1,2 , Carla Candeias 1 and Fernando Rocha 1, *

1 GeoBioTec Research Centre, Department of Geosciences, University of Aveiro, 3810-193 Aveiro, Portugal;
bernardino.bernardo@ua.pt (B.B.); candeias@ua.pt (C.C.)
2 Faculty of Earth Sciences and Environment, Pedagogic University of Maputo, Av. do Trabalho,
Maputo 2482, Mozambique
* Correspondence: tavares.rocha@ua.pt

Abstract: In this study, electrical resistivity was applied in six 400 m profiles around the Hulene-B
waste dump (Mozambique). Afterwards, an inversion was performed by RES2Dinv. The use of
the electrical resistivity method allowed us to characterize in detail some underlying aspects of the
DRASTIC index by identifying anomalous zones considered to be permeable and prone to leachate
migration. The modified DRASTIC index revealed high values in areas near contaminated surface
groundwater and surface layers of the vadose zone, characterized by low resistivities. Areas with
lower index results were characterized by high resistivity on surface layers and high depth at which
groundwater was detected. The overall modified DRASTIC index result revealed medium vulnera-
bility. However, high vulnerability index values were detected in areas with higher surface elevation,
suggesting groundwater contamination by horizontal dilution of leachates from the surrounding area
of the Hulene-B waste dump.
Citation: Bernardo, B.; Candeias, C.;
Rocha, F. Integration of Electrical
Resistivity and Modified DRASTIC Keywords: resistivity; anomalous zones; modified DRASTIC model; groundwater vulnerability
Model to Assess Groundwater
Vulnerability in the Surrounding
Area of Hulene-B Waste Dump,
Maputo, Mozambique. Water 2022, 1. Introduction
14, 1746. https://doi.org/10.3390/ Urban areas are characterized by excessive production of solid waste [1], which is
w14111746 often deposited in areas not prepared for disposal or treatment, thus posing a risk of envi-
Academic Editor: Lluís Rivero ronmental contamination [2]. Soils and groundwater are described as extremely vulnerable
to pollution [3]. The concept of groundwater vulnerability was first described in the early
Received: 24 April 2022
1960s, aiming to identify areas prone to contamination [4]. Groundwater vulnerability
Accepted: 26 May 2022
depends not only on its flow system properties but also on contaminant sources’ proximity,
Published: 29 May 2022
and contaminant characteristics, among other factors. These can promote potential contam-
Publisher’s Note: MDPI stays neutral inants to reach groundwater resources [5]. In urban areas, one of the main groundwater
with regard to jurisdictional claims in contamination sources is leachates, resulting from the decomposition of solid urban waste
published maps and institutional affil- deposited without treatment in unplanned locations [6].
iations. Several non-invasive models have been developed to assess groundwater vulnerability,
of which geophysics, in particular electrical resistivity, and the DRASTIC hydrogeological
model are pointed as the most relevant [7]. The electrical resistivity method has been used
for locating hazardous waste in depth, and to identify different sources of contamination
Copyright: © 2022 by the authors.
in subsurface environments [8,9]. Has been widely used to detect areas with heavy metal
Licensee MDPI, Basel, Switzerland.
contamination plumes [10], groundwater [7] and lithological variations [11].
This article is an open access article
The DRASTIC model was defined by Aller, [12]. Seven hydrogeological parameters are
distributed under the terms and
conditions of the Creative Commons
included, being acronyms of the term “DRASTIC”, Depth of water table, net area Recharge,
Attribution (CC BY) license (https://
Aquifer media, Soil media, Topography vadose zone impact, and hydraulic Conductivity.
creativecommons.org/licenses/by/
This model has been applied to assess groundwater vulnerability in relatively large urban
4.0/). areas (>40 ha) [13,14]. Shah et al. [15], Arowoogun et al. [16], Dhakate et al. [17] and

Water 2022, 14, 1746. https://doi.org/10.3390/w14111746 https://www.mdpi.com/journal/water


Water 2022, 14, 1746 2 of 18

George [18] showed its effectiveness when combined with electrical resistivity to study
contamination plumes migration, with different sources (e.g., dumps, mines, cemeteries)
and estimate groundwater vulnerability.
Voudouris’s [14] recent study on the application of the DRASTIC model to assess
groundwater vulnerability suggested the use of the DRASTIC method in areas >40 ha,
combined with GIS. Shah et al. [15] applied the DRASTIC model and electrical conduc-
tivity to evaluate groundwater vulnerability in Pakistan, revealing lithology significance
on different resistivity and contamination flow. Other studies by George [18] demon-
strated that groundwater vulnerability analysis, using a combination of electrical re-
sistivity and DRASTIC, was successfully achieved. Islami et al. [19] successfully com-
bined resistivity and DRASTIC methods around the Pekanbaru dumpsite in Pekanbaru,
Indonesia. Additionally, the original DRASTIC model was extensively modified to en-
hance groundwater vulnerability studies around landfills and dumpsites, e.g., El Naqa [20];
Vosoogh et al. [21]; Santhosh et al. [22]; Mohammadi et al. [23]. The modified DRASTIC
model allows the incorporation of other variables that influence groundwater vulnerabil-
ity [22]. Vosoogoh et al. [21] applied the modified DRASTIC (land use “L” effect) model to
study older and recent landfills in Iran. The same method was used to study groundwater
vulnerability around a landfill in Nigeria [24]. However, there are few studies that applied
electrical resistivity and modified DRASTIC methods in smaller areas without using the
layer interpolation method (DRASTIC factors), which is often pointed to create greater
spatial bias, being the less obvious choice of specific areas for structural intervention [24].
Previous studies on groundwater in Mozambique, particularly in Maputo city, con-
sidered it vulnerable to contamination due to the hydrogeological context characterized
by its shallow level and high urban growth without adequate planning and sanitation sys-
tems [25,26]. Hulene-B, the largest open-air dump in Mozambique (~17 ha) and its influence
on soil and groundwater contamination due to horizontal and vertical leachate migration
has been studied [27,28]. This study aims to integrate electrical resistivity and the modified
DRASTIC model to identify anomalous leachate migration and to estimate groundwater
vulnerability in the surrounding area of the Hulene-B waste dump, Maputo, Mozambique.

2. Materials and Methods


2.1. Study Area
The Hulene-B waste dump is considered the largest in Mozambique [29], and is
located in Maputo city (Figure 1), surrounded by Hulene-B and Laulane residential areas,
with approximately 49,000 inhabitants [30]. The immediate area of the dump was densely
populated until February 2018, when the fall of a large mass of wastes caused the collapse of
32 houses and the death of 18 inhabitants, which led to the forced removal of the population
within a range of 50 m to the dump [31]. The Hulene-B dump, an abandoned quarry with
no previous preparation for waste deposition receives all types of wastes produced in
Maputo City, e.g., domestic, industrial, medical, and construction [32,33]. The height of the
waste is estimated to be between 6 and 15 m in depth, in an area of ~17 ha [34,35].
Geologically, the Hulene-B waste dump is in the Mesocenozoic sedimentary basin, in
southern Mozambique [36] in a contact zone of two lithologies (Ponta Vermelha TPv, and
Malhazine QMa Formations) [27] (Figure 2). The Ponta Vermelha Formation dates from the
upper Pliocene to the lower Pleistocene, being composed in the upper part of ferruginous
sandstones and red silty sands, which gradually change to yellow and whitish sands [37].
On the surface this unit presents a red color, being poorly consolidated, and loose sands
may appear [26]. The Malhazine Formation, from the upper Pleistocene, consists of fine,
poorly consolidated sands with whitish to reddish colors, fixed by vegetation on successive
consolidation processes [26]. Waste deposition in the Hulene-B dump is mostly located in
QMa, spreading to the East (TPv).
Water 2022, 14, 1746 3 of 18

Figure 1. Study Area (a) intra ‐


‐ dune depression; and (b) Hulene-B waste dump.
The eastern dump boundary ‐ corresponds to small slopes ranging from 52 to 54 m,
while the western boundary presents smaller slopes, ranging 32 to 34 m (Figure 2b).
Momade et al. [38] performed drillings L16 and L125 (Figure 2a) in two geological forma-
tions around the dump,
‐ showing that TPv formation underlain QMa, and TSa (Miocene-
Pliocene) is composed of clayey, calcareous sandstone with Ostrea cullata in its upper part.



‐ ‐

Figure 2. (a) Geological features; (b) study area topography; and (c) Hulene-B waste dump (adapt.
Momade et al. [38]).

The Hulene-B dump hydrogeological system is in the Tertiary-Quaternary aquifer


system [26]. The aquifer substrate is formed by a layer of clayey marl to grey clay [38].
The localized presence of the semi-impermeable layer (clayey sands), between fine and
coarse sand and sandstones, in the surroundings of the Hulene-B dump, causes water
circulation in these two sectors [28]. Coarse sands lie directly on top of the clay layer, in
some sections, promoting semi-confined conditions [25]. The water level on local wells
varies between 1.5 and 9.3 m in depth, with an average of 3.8 m [38]. Bernardo et al. [27]
used electrical resistivity profiles in 2020 and 2021, suggesting that groundwater in the
western boundary of the Hulene-B dump was at variable depths and with a potential
Water 2022, 14, 1746 4 of 18

risk of being contaminated by leachate plumes resulting from vertical and horizontal
migration, which, in the subsurface environment, were assigned d values 4.26 to 8.5 Ω.m.
The hydraulic conductivity was estimated to be 1 to 5 m/d on the surroundings of the
Hulene-B dump [28].
The predominant local climate is subtropical, with mean annual precipitation of
~789 mm, with two climatic seasons: (a) hot (mean 25 ◦ C) rainy period from December to
March, representing >60% of the annual precipitation, with its peak in January (~125 mm);
and (b) dry and cold season, from April to September, with lower temperatures in June and
July (mean 21 ◦ C), and scarce precipitation, whose minimum values recorded in August
(~12 mm) [39]. The prevailing winds are SE [40].

2.2. Electrical Resistivity


Electrical resistivity studies are based on electric current injected into the ground
through a pair of electrodes (A and B—current electrodes), and the resulting potential
difference between another pair of electrodes (M and N—potential electrodes) [41]. Ground
resistivity is calculated by distances between electrodes, applied current and measured
potential difference, based on the Law of Ohm [42].
The soil’s electrical resistivity is a characteristic closely linked to the type, nature,
and state of alteration of geological formations [8]. In areas with potential groundwater
contamination, it has been used for determining the depth of the groundwater table [43]
(Akhtar et al., 2021), determining the distribution of contamination areas and the direction
of migration of pollutants, assessing the thickness of wastes deposited in a landfill, and
identifying possible leachate plumes [41,44,45]. Soil apparent resistivity (ρa) can be deter-
mined based on the known difference between electric field potential (∆V), the current (I),
and the distance between electrodes [41], given by the equation:

ρa = k ∆V/I (1)

where ρa is apparent resistivity, I the intensity of current applied to the soil by electrodes
A and B (mA), ∆V the differential potential between electrodes M and N (mV), and k the
geometrical coefficient of electrode positioning (m). The geometrical factor k is dependent
on the distribution geometry of the electrodes, as follows:


k=   (2)
1 1 1 1
AM
− BM
− AN
+ BN

where AM, BM, AN and BN represent the geometrical distance between electrodes A and
M, B and M, A and N, and B and N, respectively.
In this study, 6 electrical resistivity profiles were performed in May 2021, of which,
4 were on the western border of the dump and 2 profiles were on the southern and northern
borders (Figure 3). Profile 3, on the north of the dump, was applied to understand the
possible migration of contaminants to areas further away from the dump (reference profile).
ABEM SAS 4000 was used for resistor data acquisition, including 4 rollers of 100 m
cables with 21 outlets connected to the same number of electrodes. The layout produced
by this sequence of cables corresponds to the standard of the reading program hosted by
the resistivimeter LUND Imaging System. Data acquisition employed a 50 Hz current
frequency, using GRAD4LX8 multigradient protocol, once provides dense coverage nearby
surface and adopts the Wenner–Schlumberger protocol (ABEM, 2018). The electrode spac-
ing for data acquisition was 5 m. All electrode take-outs were connected in the GRAD4S8
protocol. The resistivimeter automatically switches electrodes to serve as current or poten-
tial pairs. After the readings, data was transferred to the resistivimeter, which then takes
3 to 6 readings to obtain the smallest error average between readings. The inversion of
the electrical resistivity data obtained in the 6 lines was performed based on standards de-
fined in software RES2DINV3.59.106, namely, the application of the smoothness constraint
method in the resistivity values of the final model, calculation of the Jacobian matrix in
Water 2022, 14, 1746 5 of 18

each iteration, standard Gauss–Newton optimization method [46]. Profiles interpretation


was based on the direction of each profile over the entire length (400 m).

Figure 3. Electrical resistivity profiles and northern (WN) and southern (WS) wells studied, in the

surroundings of the Hulene-B waste dump.

2.3. Modified DRASTIC Groundwater Vulnerability


The DRASTIC model has been commonly used in areas where geographical, and
hydrogeological information is available and has been successfully applied in different
regions [15,47,48]. The word DRASTIC is an abbreviation of the initial letter of different
parameters such as ‘D’ to depth to water; ‘R’ to net recharge, ‘A’ to aquifer media, ‘S’
to soil media, ‘T’ to topography, ‘I’ the impact of the vadose zone media, and ‘C’ to the

hydraulic conductivity of the aquifer intrinsic vulnerability of groundwater is evaluated by

the DRASTIC index formula which is given below:

DRASTIC Index = Dr Dw + Rr Rw + Ar Aw + Sr Sw + Tr Tw +Ir Iw + Cr Cw (3)

where “r” is the rating value, and “w” is the weight assigned to each parameter. Each
factor is assigned a relative weight ranging from 1 to 5 (Table 1). Each DRASTIC factor
is divided into ranges that affect the contaminant potential. The range for each factor
lies from 1 to 10. The DRASTIC model depends on seven boundaries or layers, which
are used as input boundaries for modeling. Thus, the interpretation of the index follows
three categories: (i) indices <135 denote low vulnerability; (ii) indices between 135 and
Water 2022, 14, 1746 6 of 18

150 represent medium vulnerability; and (iii) indices >150 suggest a high vulnerability to
groundwater-related environmental impacts [6,12].

Table 1. DRASTIC parameters.

Factor Interval/Characteristics Value (r) Weight (w)


0–1.5 10
1.5–4.6 9
4.6–9.3 7
(D)
Groundwater depth (m) 9.3–15 5 5
15–23 3
23–30 2
>30 1
0–50 1
50–100 3
(R)
Net recharge rate (mm/year) 100–175 6 4
175–250 8
>250 9
(A)
Sand 7 3
Aquifer media
0–1.5 10
1.5–4.6 9
(S) 4.6–9.3 7
Distance between the anomalous 2
9.3–15 5
surface layer and groundwater
15–23 3
23–30 2
>30 1
0–2 10
2–6 9
(T)
Terrain slope (%) 6–12 5 1
12–18 3
>18 1
Sandstones 4–8
Limestones,
4–8
sandstones, and shales
(I)
Sands and gravels 5
Vadose Zone
with significant silt 4–8
and clay content
Sands 8
1–4.1 1
4.1–12.2 2
(C) 12.2–28.5 4
3
Hydraulic conductivity (m/day)
28.5–40.7 6
40.7–81.5 8
>81.5 10
Water 2022, 14, 1746 7 of 18


In this study we applied the modified DRASTIC model combined with electrical resis-‐
tivity data. This allows a specific assessment of the vulnerability of the dump surrounding
area, based on the distance between anomalous surface layer (low resistivity; leachate
influenced) and groundwater. The soil variable was replaced since soils around the landfill
were classified as sandy, which is characteristic of the whole surroundings [38]. Thus, ‘S’
corresponds to the distance between the superficial anomalous layer and groundwater.
Values and weights of the variables were kept the same as for soil, given processes similarity,
controlled by these factors (migration and attenuation of leachate). Layers spatial distri-
bution was not made, given detailed description for each factor in depth and superficial
slight change, as well as study area size [49,50].
For resistivity models and groundwater
− vulnerability validation, groundwater depth,
pH and sulphates (PO4 3− ) were measured in two wells in the surroundings of the dump
(Figures 3 and S1). Chemical analysis of total phosphate was performed with an HI96713,
with a resolution level of 0.01 mg/L (Figure S1).

3. Results
3.1. Resistivity Models and Potential Contamination Risk
For the analysis of the profiles, the resistivity values of the profiles were adjusted
to the same scale so that each color of the contour in the resistivity model implies the
same resistivity value (Figure 4). In this study, the electrical resistivity models were
analyzed to identify the possible influence of leachate on groundwater contamination.
Thus, anomalous zones that may reflect the leachate migration and contamination process
were identified: (i) leachate generation and migration areas (7.99–16.8ΩΩ.m); (ii) saturated
zones contaminated by leachate (4.96–7.99 Ω.m);Ω (iii) groundwater and surface water
Ω
contaminated by leachate (1.535–7.99 Ω.m); Ω
(iv) waste and lithologies local (>16.8 Ω.m).

Figure 4. Calibrated color scale showing the resistivity range of materials and their characteristics.

Profile 1, from south to north of the dump (Figure 5a), is superficially characterized
by high resistivity associated with rubble, house debris and compacted waste in non-wet ‐
environments from 0 to 35 m. From 240 to 280 m there were zones with a rather hetero-‐
geneous anomalous resistivity, which were considered leachate generation and migration‐
zones, (7.99–16.8 Ω.m)Ωgiven the strong accumulation of surface leachates resulting from

the E-W movement (profile 2) that are diluted with the surface waters, causing possible
contamination (280 to 400 m) (1.535–7.99 Ω.m). AtΩ medium depths 30–56.5 we note the
predominance of a vast layer throughout the length of the profile that we interpret as TPv‐
Ω possible influence of horizontal migra-‐
lithologies, less resistive (12.9–16.8 Ω.m) given the
tion at depth described in profile 2. At deeper levels between 47 and 56.5 m, anomalous
zones are observed, which were considered as lithologies influenced by horizontal leachate‐
migration and possible contaminated groundwater (>45 m depth) (1.535–7.99 Ω.m).
Ω along the W-E direction (Figure 5b), is quite heterogeneous, with high surface
Profile 2,
resistivities from 0 to 140 m, which represents rubble, old house debris, and waste buried in
a non-wet environment west of the dump, followed, from 140 to 160 m, by a zone of possible
migration of surface leachate into the subsurface environment, causing an extensive zone
of subsurface anomalies, which were considered as lithologies contaminated by strong
horizontal migration of leachate with E-W direction (7.99–16.8 Ω.m) and saturated zone of
contaminated groundwater (1.535–7.99 Ω.m), differentiated levels of semi-confinement of
Water 2022, 14, 1746 8 of 18

aquifers. From 245 to 400 m depths, there were local anomalies and lithologies with higher
resistivities, corresponding to highly compacted waste with diverse contents. At depths
ranging from 30 to 160 m, semi-confined anomalies (<8.35 m depth) and aquifers (>10 m
depth) separated by semi-saturated layers were noted, demonstrating the existence of a
possible continuous connection between the two. These characteristics were described as
conducive to groundwater contamination at various depths, mainly in the surroundings
of the Hulene-B waste dump, where surface leachate flows were noted with successive
enrichment of lithological layers (7.99–16.8 Ω.m).

Figure 5. Profiles of the variation of the electrical resistivity in the study area.

Profile 3 (Figure 5c) north of the dump in the S-N direction, to study possible dynamics
of groundwater contamination, ~300 m away from the dump. The profile at the surface level
exhibited generally high resistivities, alternating between rubble and highly compacted
soils. At the deeper level (>40 m) in the southern end, an anomalous zone was found and
considered as contaminated lithologies (7.99–16.8 Ω.m). This anomaly was associated with
the horizontal and vertical migration of leachate described in profile 2.
Profile 4, in the NE-SW direction (Figure 5d), from the starting point to 150 m, showed
generally high resistivities associated with compacted residues and rubble of old houses.
From 150 m depth to the end of the profile, a continuous decrease in resistivity was
Water 2022, 14, 1746 9 of 18

observed, which can be associated with saturated and wet areas with origin in a natural
receiving basin where new waste deposits were observed, a localized source of dilution,
and vertical migration of leachate and groundwater contamination (<7.99 Ω.m).
Between 300 and 310 m, relatively high resistivities were found, associated with waste
deposited in the intra dune depression with low surface moisture level. From 320 m to the
end of the profile, a shallow aquifer was evident (<7.99 Ω.m), as this area corresponds to the
end of the depression (SW) with a surface covered by humid soils, evidenced by low surface
resistivities. Between 6 and 56 m in depth, was noticed a large anomaly (16.3–7.99 Ω.m)
related to surface influence by leachates propagated to greater depths (>37.4 m) causing
possible groundwater contamination (<7.99 Ω.m).
Profile 5, with a west-east orientation, was at the southern end of the dump (Figure 5e).
The electrical resistivity results did not display significant changes at surface level and
showed higher resistivities associated with compacted soils and rubble (including road
asphalt where the profile was executed). This road, besides being in the southern boundary
of the dump, is an access route to the interior of the Hulene-B neighborhood. Between
130 and 145 m (>40 m depth) extended a resistive zone of low anomalous values, which
were interpreted as lithologies that might influence leachate migration (7.99–16.8 Ω.m Ω.m)
and possible water contamination by the vertical movement of leachates (1.535–7.99 Ω.m).
Profile 6, this profile was taken at the eastern limit of the dump, along the S-N direction,
between the dump and Julius Nyerere Avenue (Figure 5f). From the starting point to 200 m,
there were resistivities with average profile values (12.9–33.4 Ω.m). These resistivities
suggested soils with different levels of compaction and surface moisture that may be
associated with various activities south of the dump. The resistivity of 33.4–53.8 Ω.m
corresponding to a superficial but thick layer, between 160 and 200 m, indicated compacted
soils at the entrance of the dump. From 8.6 m depth, higher resistivity values (>70.3 Ω.m)
were noted which may be related to the sandstone stratum with different levels of alteration,
typical of the TPv formation [25]. Onwards, in the northern direction, resistivity starts to
decrease successively (<27.05 Ω.m) along thick layers with moisture levels that increase
until to groundwater (<10.44 Ω.m).
The saturated area occupied a large space, revealing the existence of an E-W ground-
water flow parallel, to the dune slope where the leachate is located. The groundwater
contamination process at this point may be occurring horizontally due to leachate diffusion,
causing localized anomalous resistivities (<10.445 Ω.m) close to the groundwater resistivity
(<7.99 Ω.m). From 240 m, resistivity begins to decrease, generating localized anomalous
zones in the subsurface, which are associated with vertical leachate migration, pointing to
the occurrence of two isolated “hot spots”. Between 240 and 280 m, below the first “hot
spots”, there was a tendency for a significant increase in resistivity, which may correspond
to less saturated layers up to the least conductive stratum (>33.4 Ω.m). Profile data showed
the existence of two mechanisms of possible saturated zone and groundwater contamina-
tion which were, horizontal dilution in the south and center of the profile, and vertical
migration (<16.8 Ω.m) and retention of leachate in localized “hot spots”.

3.2. Modified DRASTIC Index


3.2.1. Depth to Water Table (D)
Aller et al. [12] refer that the depth of the water table determines the depth through
which a contaminant moves before reaching the aquifer and determines the contact time
with the surrounding media. Thus, a greater possibility of contamination mitigation occurs
when the depth of the water table is greater because a deeper water table implies more
travel time and less vulnerability to contamination [13]. Thus, the deeper the phreatic
table implies more travel time and less vulnerability to contamination [13,47]. On the
eastern, southwestern and northern boundary of the dump (area covered by profiles
6, 4, 2 and 1), subsurface waters and groundwater were detected between 1.5 and 4.6 m,
and on the southern, northern and western boundary at depths >30 m (area covered by
profiles 5 and 3). Results for the eastern border southwest and northwest were similar to
Water 2022, 14, 1746 10 of 18

those published by [28], who classified the predominant aquifers in Maputo city as shallow
and phreatic aquifers, estimating their average depth between 1.5 and 9.3 m. DRASTIC
parameters of 9 and 1, respectively, were assigned, and D = 5 (Tables 1 and 2).

Table 2. Parameter values considered in the DRASTIC Index.

Characteristics of the Surroundings of the


P1 P2 P3 P4 P5 P6 Mean
Waste Dump
Depth of groundwater level 45 45 5 45 5 45 31.6
Recharge capacity 32 32 32 32 32 32 32
Sands 21 21 21 21 21 21 21
Distance of anomalous surface layer
20 20 2 20 2 20 14
and groundwater
Plan, soft dune and interdune depression 10 5 10 10 5 10 8.3
Sands 40 40 20 40 20 40 33.3
Hydraulic conductivity 6 6 6 6 6 6 6
DRASTIC index: 174 169 96 174 91 174 146.3
P—Profile.

3.2.2. Net Recharge (R)


Net recharge is the amount of surface water that infiltrated the underground and
reaches groundwater [6], indicating the amount of water from precipitation that was
available for vertical transport, dispersion, and dilution of pollutants from a given applica-
tion point [5,12]. Recharge water in the dump’s surroundings is a source of contaminant
transport within the vadose zone to the aquifer [19]. The greater the recharge, the more
vulnerable the groundwater [51]. Given the small area analyzed in this study, data used ac-
cording to Momade et al. [38] and Vicente [36] estimates the value of groundwater recharge
in Maputo city between 165 and 185 mm/year for the entire Hulene-B dump surrounding
area. A value of 8 was assigned to the area and R = 4 (Tables 1 and 2).

3.2.3. Aquifer Media (A)


Aquifer media refers to a rock in the ground that serves as water storage [52]. It
indicates material property that controls pollutant attenuation processes based on the
permeability of each layer [53]. The attenuation characteristic of the aquifer material is
reflected by the mobility of contaminants through aquifer media [47]. In the surroundings
of the Hulene-B dump, two types of semi-confined (west of profile area, 2) and shallow
(south-west of profile area 1, 4 and area covered by profile 6) aquifers were assumed to
exist, which have been described by Momade et al. [38], Vicente [36], and Cendon et al. [26],
composed of inland dune sands and semi-permeable sands, and recharge occurs mainly by
precipitation given the permeable surfaces such as dune sands. The value of 7 was assigned
to the area around the Hulene-B dump and A = 3 (Tables 1 and 2).

3.2.4. Distance of the Anomalous Surface Layer and Groundwater (S)


The distance between anomalous surface layer (low resistivities) generally represents
surfaces contaminated by leachates in areas close to landfills [42]. Anomalous surface to
groundwater band areas were characterized by intense leachate migration, which was
evidenced by transected profiles 2 (<10.445 Ω.m) and 6 (<16.8 Ω.m). The area north of
profile 1 and southwest of profile 4 show intense anomalies that we interpret as a shallow
aquifer (4) and surface soils enriched by leachates (1) that accumulate successively to the
west that can easily migrate into the confined aquifer described in (2). However, other
profiles did not show bands with the continuous connection of anomalies and groundwater.
The area covered by profiles 1, 2, 4 and 6 was assigned the value 10 and other areas value 1,
and S = 2 (Tables 1 and 2).
Water 2022, 14, 1746 11 of 18

3.2.5. Topography (T)


Topography refers to the slope of an area [47]. It controls the probability of a pollutant
being transported by runoff or remaining in the soil where it may be infiltrated [12]. The
softer the slope (slope of 0–2%), the higher the water and/or pollutant holding capacity,
while in slopes >10%, lower water and/or pollutant holding capacity occurs [12]. In the
surroundings of the dump, the relief is heterogeneous (Figure 2). The area covered by
profiles 1, 3, 4 and 6 did not present slopes. However, the area covered by profiles 2 and
5 has sloped >10%. Thus, the area covered by profiles 2 and 5 was assigned the value 5 and
the other areas were assigned the value is 10, and T = 1 (Tables 1 and 2).

3.2.6. Vadose (I)


The vadose zone is the unsaturated zone that lies below the soil horizon and above the
water table [47]. It determines the attenuation characteristics of the contaminants [12]. The
movement of contaminants into the saturated zone is controlled by this parameter. Aller
et al. [12] and Asfaw et al. [47] refer that if the flood zone consists of sand, the potential risk
of contamination of the aquifer is very high. The surrounding soils of the Hulene-B dump
are dune sands. However, the combination of electrical resistivity data showed specific
characteristics, which allowed us to classify in detail the environment of the dump. The
area covered by profiles 3 and 5 was characterized by having very resistive surface layers,
showing a low infiltration rate, so a value of 4 was assigned. Profile 1 was heterogeneous,
with half of the area covered being very resistive(south) and with very low resistivities, the
assigned value was 8. Profiles 2, 4 and 6 had a lower resistivity, showing higher infiltration,
which is typical of sandy soils, so a value of 8 was assigned, and I = 5 (Tables 1 and 2).

3.2.7. Hydraulic Conductivity (C)


Hydraulic conductivity is described as the ability of materials to transmit water to
aquifers, in turn controlling the rate of groundwater and contaminant material flow under
a given hydraulic gradient [13]. It controls contaminant migration and dispersion from the
injection point within the saturated zone [47]. In the surroundings of the Hulene-B waste
dump, the hydraulic conductivity was estimated by Momade et al. [38] as 1–5 m/d. For
the whole studied area, a value of 2 was assigned, and C = 3 (Tables 1 and 2).

3.3. Descriptive Statistics of Electrical Resistivity and DRASTIC Index


The electrical resistivity values of the areas covered by the profiles were projected with
the vulnerability index values (Table 3). In general, the areas covered by profiles 2, 4 and
6 were classified as having a high DRASTIC index. The profile areas 2 and 4, with mean
resistivity values of 20.22 and 18.1 Ω.m, respectively, suggested the predominance of lower
resistivity across the profile surface area which extends into the groundwater, suggesting
successive leachate migration.

Table 3. Mean, maximum, minimum, standard deviation of resistivity values (Ω.m) and DRASTIC index.

ID Min Max Mean SD Modified DRASTIC


P1 8.64 42.36 20.78 5.72 174
P2 5.37 119.1 20.22 9.31 169
P3 10.12 50.8 31.01 7.3 96
P4 6.98 41.19 18.1 4.64 174
P5 1.04 477.1 37.35 31.86 91
P6 3.06 207 49.2 32.78 174
SD—standard deviation.

The area covered by profile 6 had an average of 49.2 Ω.m, a minimum of 3.06 Ω.m
and a maximum of 207 Ω.m. The average resistivity value was relatively higher but
Water 2022, 14, 1746 12 of 18

showed a much higher standard deviation (32.78 Ω.m), revealing resistivity heterogeneity,
marked by the existence of surface anomalous bands that connect with the groundwater
at various points, which may be associated with a greater migration of leachates and
groundwater contamination.
Areas covered by profiles 3 and 5 revealed a low DRASTIC index and heterogeneous
electrical resistivity. Thus, the area covered by profile 1 showed heterogeneity in surface
and subsurface anomalies that may be associated with vertical and horizontal migration of
contaminants with minimum resistivity values of (1.535–7.99 Ω.m) revealing a higher risk
of contamination of the semi-confined aquifers described in profile 2. However, the greater
depth at which the groundwater was detected reveals a natural attenuation mechanism
of contamination by the underlying lithologies. The area covered by profile 3 showed an
average resistivity of 31.01 Ω.m, ranging from 10.12 to 50.8 Ω.m. These results suggest the
predominance of high resistivity values, which is translated by a reduced predominance
of resistive surfaces that were interpreted as less permeable and less contaminated sub-
strates. The minimum resistivity at great depths (7.99–10.12 Ω.m) may be associated with a
saturated area or localized influence of horizontal migration of contaminants, described
in profile 2. The area covered by profile 5, presented an average resistivity of 37.35 Ω.m,
ranging from 1.04 to 477.1 Ω.m. Groundwater was detected at depths >40 m, with less risk
of contamination by vertical migration.
The low DRASTIC values in areas covered by profiles 3, and 5 resulted from the
combination of two factors, high resistivities prevailing in the surface lithologies suggesting
low infiltration, which greatly reduces the risk of vertical migration of leachate to deep
layers, and greater depth at which, resistivities interpreted as groundwater, were found.
In general, data showed areas covered by profiles, with high average resistivity, with
lower DRASTIC index (profiles 3 and 5) (Figure 6). Exceptionally, the area covered by
profile 6 showed a high DRASTIC index with relatively high mean resistivity values due
to its higher standard deviation and predominance of resistivities <10.445 Ω.m in a large
surface strip that was associated with leachate migration to groundwater at low depth
than in all profiles.ΩHowever, in the area covered by profiles 1, 2 and 4, the presence of
resistive anomalies was considered as semi-confined and shallow ‐
aquifers and surface
anomalies that are conducive to contaminant migration were determining factors for high
DRASTIC index.

Ω
Figure 6. Electrical resistivity values (minimum, maximum, mean, standard
Ω
deviation Ω.m) and
modified DRASTIC Index. Maximum value of profile 5 (477.1 Ω.m) was set to (200) to maintain the
scale within the limits of the perceived index.

4. Data Integration ‐
The resistivity data of the areas covered by the profiles 1, 2, 4 and 6, with a predom-
inance of surface and groundwater resistivity

anomalies, was described in other studies
suggesting leachate migration [54], and groundwater contamination [7], mainly in areas
surrounding non-isolated dumps where leachate can freely circulate through adjacent
lithologies and subsequently affecting the vadose zone and groundwater [55]. Surface

Ω
Water 2022, 14, 1746 13 of 18

circulation and leachate infiltration are well evidenced in profiles 1, 2, 4 and 6, by surface
and subsurface resistive anomalies, indicative of contamination [56].
In the area covered by profiles 1 (in the north), 2 and 4, continuous anomalies from
surface to groundwater were noted, described as lithological migration bands of leachate to
groundwater [57,58]. The area transacted by profile 6, besides showing an extensive layer
with low resistivities (<10.445 Ω.m), has been considered in similar studies as a saturated
zone and groundwater under leachate influence [45,59]. In the north of the dump, two
localized points of resistive anomalies were identified as “hot spots” [60,61] resulting from
vertical migration and localized accumulation of contaminants Feng et al. [62], and may be
associated with the confined aquifer system, described in this zone as quite vulnerable to
contamination given its proximity to the surface and sandy characteristics of the vadose
zone [26].
The electrical resistivity data of the areas covered by profiles 3 and 5 were character-
ized by high resistivities, and associated with low contaminant infiltration capacity [13],
given high surface compaction [63]. The area transacted by profile 3 is located far from
the waste dump (>350 m), revealing that contaminants may not be reaching this area.
Morita et al. [64] and Touzani et al. [9] demonstrated that a resistivity increase away from
dumpsites represents a significant decrease in contamination due to the attenuating role
of soils and groundwater. Groundwater at great depths in the area covered by profiles
1, 3 and 5 has been described in similar studies as a determinant of low contamination
risk [18,65], which partly explains the low DRASTIC index in these areas. Paul et al. [53]

and Boumaiza et al. [66] mentioned that the depth of groundwater and the characteristics
of the infiltration zone are the most important factors determining the DRASTIC risk.
Gemail et al. [13], Shah et al. [15] and Nasri et al. [67] integrated electrical conduc-

tivity and DRASTIC data and concluded that areas with the highest DRASTIC index
around dumps, generally exhibit low electrical resistivity associated with the migration of
contamination through adjacent lithologies that may subsequently reach groundwater.
The spatial projection of the Vulnerability Index in the surroundings of the Hulene-

B dump was revealed to be higher to the east, southwest and northwest of the dump,
with high and transient relief, and lower to the south and west of the dump with‐low
elevation, except for the southern area which is transient (Figure 7). Tan et al. [68] and
Blarasin et al. [69] have reported that contamination of the upper levels of water tables
leads to the dispersion of contaminants to larger areas.


Figure 7. Spatial projection of the modified DRASTIC in surroundings of the Hulene-B dump: dashed
in white represents low vulnerability and in black high vulnerability.


Water 2022, 14, 1746 14 of 18

The combination of electrical resistivity data for the assessment of groundwater vulner-
ability indices is an important tool for DRASTIC factors assessment, such as: groundwater
depth, and intrinsic characteristics of the vadose zone. The modification of the “soil”
factor in the original DRASTIC by the “distance between the anomalous surface layer and
groundwater” is important in areas with potential contaminant migration to groundwater,
as well as in studies aiming to outline remediation measures in areas of groundwater
contamination flow.
Measured groundwater depth in the northern well (WN) was 6.5 m, and the southern
well (WS) was 5.1 m similar to the electrical resistivity model (Table 4). However, The WS
location was further south of the profile 4 area and results showed that groundwater levels
tended to be more at the surface when approaching intradunar depression. Groundwater
pH and PO4 3− was 6.1, 1.33 mg/L in WS and, 8.4 and 0.43 mg/L, respectively. Phosphate
results were above the natural groundwater reference value of 0.005 to 0.05 mg/L [70].
Previous studies suggested that active waste dumps have a phosphate production of
1–100 mg/L, which can be incorporated into groundwater by leaching [71,72]. The pro-
longed consumption of water contaminated with high levels of phosphates can damage
blood vessels, damage kidneys, cause osteoporosis and induce aging processes [73].

Table 4. Data validation of electrical resistivity and modified DRASTIC.

ID Depth pH RfD [70] PO4 3− RfD [70]


WS 5.1 m 6.1 1.33 mg/L
6.5–8.5 0.005–0.05 mg/L
WN 6.5 m 8.4 0.43 mg/L
RfD—Natural reference value.

5. Conclusions
In this study, the combination of electrical resistivity and modified DRASTIC models
was effective in describing the hydrogeological particularities, estimating in detail ground-
water vulnerability, and identifying areas of possible leachate migration into groundwater
in the surroundings of the Hulene-B dump. Areas covered by profiles 1 (north), 2, 4 and
6 showed strong indications of possible groundwater contamination, with a modified
DRASTIC index which was very high (169–174) due to the proximity of groundwater to the
contaminating surface (dump) and the connection of continuous anomalous layers from
the surface to the aquifers. In the area of profiles 3 and 5 the index was low (91–96) due to
the strong resistivity of the surface layers and the high depth at which groundwater was
detected (>40 m).
The overall value of the DRASTIC modified index for the surrounding area was
estimated at 146 representing a medium overall vulnerability. However, a higher vul-
nerability index in areas covered by profiles 1, 2, 4 and 6, given their relatively higher
altitude (2 and 6), suggested groundwater contamination by horizontal dilution. Studies
were underway to assess areas of suspected vertical migration of leachate by chemical
analysis as well as groundwater. Groundwater depth local validation data was similar to
the electrical resistivity model. Groundwater contamination risk identified by the modified
DRASTIC vulnerability index was confirmed by high levels of phosphates in groundwater
samples studied.
Several challenges remain for further studies, such as quantitative studies to validate
areas of leachate migration that continuously connect to groundwater in areas covered
by profiles 2, 4 and 6 and chemical analysis of surrounding wells water, and contamina-
tion studies of surrounding soils at different depths, especially in areas with suspected
contaminant migration.

Supplementary Materials: The following supporting information can be downloaded at: https://
www.mdpi.com/article/10.3390/w14111746/s1, Figure S1: Sampling and chemical analysis of well
water (a) Water collection—South well (b) Result of phosphate analysis —North well; (c) Chemical
analysis process (d) South well.
Water 2022, 14, 1746 15 of 18

Author Contributions: Methodology, B.B. and F.R.; validation, B.B., C.C. and F.R.; formal analysis,
C.C. and F.R.; writing—original draft pr appendix eparation, B.B.; writing—review and editing, B.B.,
C.C. and F.R.; supervision, F.R. and C.C. All authors have read and agreed to the published version
of the manuscript.
Funding: This work was supported by GeoBioTec (UIDB/04035/2020) Research Center, funded by
FEDER funds through the Operational Program Competitiveness Factors COMPETE and by National
funds through FCT. First author acknowledges grants from the Portuguese Institute Camões and FNI
(Investigation National Fund—Mozambique).
Data Availability Statement: Not applicable.
Conflicts of Interest: The authors declare no conflict of interest.

References
1. González-Arqueros, M.L.; Domínguez-Vázquez, G.; Alfaro-Cuevas-Villanueva, R.; Israde-Alcántara, I.; Buenrostro-Delgado, O.
Hazardous solid waste confined in closed dump of Morelia: An urgent environmental liability to attend in developing countries.
Sustainability 2021, 13, 2557. [CrossRef]
2. Rathi, B.S.; Kumar, P.S. Critical review on hazardous pollutants in water environment: Occurrence, monitoring, fate, removal
technologies and risk assessment. Sci. Total Environ. 2021, 797, 149134. [CrossRef] [PubMed]
3. Boufekane, A.; Yahiaoui, S.; Meddi, H.; Meddi, M.; Busico, G. Modified DRASTIC index model for groundwater vulnerability
mapping using geostatistic methods and GIS in the Mitidja plain area (Algeria). Environ. Forensics 2021, 1–18. [CrossRef]
4. Saranya, T.; Saravanan, S. A comparative analysis on groundwater vulnerability models—Fuzzy DRASTIC and fuzzy DRASTIC-L.
Environ. Sci. Pollut. Res. 2021, 1–15. [CrossRef] [PubMed]
5. Sresto, M.A.; Siddika, S.; Haque, M.N.; Saroar, M. Groundwater vulnerability assessment in Khulna district of Bangladesh by
integrating fuzzy algorithm and DRASTIC (DRASTIC-L) model. Model. Earth Syst. Environ. 2021, 1–15. [CrossRef]
6. Ghosh, R.; Sutradhar, S.; Mondal, P.; Das, N. Application of DRASTIC model for assessing groundwater vulnerability: A study on
Birbhum district, West Bengal, India. Model. Earth Syst. Environ. 2021, 7, 1225–1239. [CrossRef]
7. El Mouine, Y.; El Hamdi, A.; Morarech, M.; Kacimi, I.; Touzani, M.; Mohsine, I.; Tiouiouine, A.; Ouardi, J.; Zouahri, A.;
Yachou, H.; et al. Landfill pollution plume survey in the Moroccan Tadla using spontaneous potential. Water 2021, 13, 910.
[CrossRef]
8. Kayode, J.S.; Arifin, M.H.; Nawawi, M. Characterization of a proposed quarry site using multi-electrode electrical resistivity
tomography. Sains Malays. 2019, 48, 945–963. [CrossRef]
9. Touzani, M.; Mohsine, I.; Ouardi, J.; Kacimi, I.; Morarech, M.; El Bahajji, M.H.; Bouramtane, T.; Tiouiouine, A.; Yameogo, S.;
El Mahrad, B. Mapping the pollution plume using the self-potential geophysical method: Case of Oum Azza Landfill, Rabat,
Morocco. Water 2021, 13, 961. [CrossRef]
10. Zhang, J.; Zhang, J.; Xing, B.; Liu, G.D.; Liang, Y. Study on the effect of municipal solid landfills on groundwater by combining the
models of variable leakage rate, leachate concentration, and contaminant solute transport. J. Environ. Manag. 2021, 292, 112815.
[CrossRef]
11. Mepaiyeda, S.; Madi, K.; Gwavava, O.; Baiyegunhi, C. Geological and geophysical assessment of groundwater contamination at
the Roundhill landfill site, Berlin, Eastern Cape, South Africa. Heliyon 2020, 6, e04249. [CrossRef] [PubMed]
12. Aller, G.; Bennett, L.; Lehr, T.; Petty, J.H.; Hackett, R.J. DRASTIC: A Standardized Method for Evaluating Ground Water Pollution
Potential Using Hydrogeologic Settings; USEPA Report 600/2-87/035; U.S. Environmental Protection Agency: Washington, DC,
USA, 1987.
13. Gemail, K.S.; El Alfy, M.; Ghoneim, M.F.; Shishtawy, A.M.; El-Bary, M.A. Comparison of DRASTIC and DC resistivity modeling
for assessing aquifer vulnerability in the central Nile Delta, Egypt. Environ. Earth Sci. 2017, 76, 350. [CrossRef]
14. Voudouris, K.; Kazakis, N. Groundwater quality and groundwater vulnerability assessment. Environments 2021, 8, 100. [CrossRef]
15. Shah, S.H.I.A.; Yan, J.; Ullah, I.; Aslam, B.; Tariq, A.; Zhang, L.; Mumtaz, F. Classification of aquifer vulnerability by using the
DRASTIC index and geo-electrical techniques. Water 2021, 13, 2144. [CrossRef]
16. Arowoogun, K.I.; Osinowo, O.O. 3D resistivity model of 1D vertical electrical sounding (VES) data for groundwater potential and
aquifer protective capacity assessment: A case study. Model. Earth Syst. Environ. 2021, 8, 2615–2626. [CrossRef]
17. Dhakate, R.; Mogali, N.J.; Modi, D. Characterization of proposed waste disposal site of granite quarry pits near Hyderabad using
hydro-geophysical and groundwater modeling studies. Environ. Earth Sci. 2021, 80, 516. [CrossRef]
18. George, N.J. Geo-electrically and hydrogeologically derived vulnerability assessments of aquifer resources in the hinterland of
parts of Akwa Ibom State, Nigeria. Solid Earth Sci. 2021, 6, 70–79. [CrossRef]
19. Islami, N.; Irianti, M.; Fakhruddin, F.; Azhar, A.; Nor, M. Application of geoelectrical resistivity method for the assessment of
shallow aquifer quality in landfill areas. Environ. Monit. Assess. 2020, 192, 606. [CrossRef]
20. El Naqa, A. Aquifer vulnerability assessment using the DRASTIC model at Russeifa landfill, northeast Jordan. Environ. Geol.
2004, 47, 51–62. [CrossRef]
Water 2022, 14, 1746 16 of 18

21. Vosoogh, A.; Baghvand, A.; Karbassi, A.; Nasrabadi, T. Landfill site selection using pollution potential zoning of aquifers by
modified DRASTIC method: Case study in Northeast Iran. Iran. J. Sci. Technol. Trans. Civ. Eng. 2017, 41, 229–239. [CrossRef]
22. Santhosh, L.G.; Sivakumar Babu, G.L. Landfill site selection based on reliability concepts using the DRASTIC method and AHP
integrated with GIS—A case study of Bengaluru city, India. Georisk 2018, 9518, 234–252. [CrossRef]
23. Abad, P.M.S.; Pazira, E.; Abadi, M.H.M. Application AHP-PROMETHEE technic for landfill site selection on based assessment of
aquifers vulnerability to pollution. Iran. J. Sci. Technol. Trans. Civ. Eng. 2021, 45, 1011. [CrossRef]
24. Majolagbe, A.O.; Adeyi, A.A.; Osibanjo, O. Vulnerability assessment of groundwater pollution in the vicinity of an active
dumpsite (Olusosun), Lagos, Nigeria. Chem. Int. 2016, 2, 232–241.
25. Nogueira, G.; Stigter, T.Y.; Zhou, Y.; Mussa, F.; Juizo, D. Understanding groundwater salinization mechanisms to secure freshwater
resources in the water-scarce city of Maputo, Mozambique. Sci. Total Environ. 2019, 661, 723–736. [CrossRef] [PubMed]
26. Cendón, D.I.; Haldorsen, S.; Chen, J.; Hankin, S.; Nogueira, G.; Momade, F.; Achimo, M.; Muiuane, E.; Mugabe, J.; Stigter, T.Y.
Hydrogeochemical aquifer characterization and its implication for groundwater development in the Maputo district, Mozambique.
Quat. Int. 2019, 547, 113–126. [CrossRef]
27. Bernardo, B.; Candeias, C.; Rocha, F. Characterization of the dynamics of leachate contamination plumes in the surroundings of
the Hulene-B waste dump in Maputo, Mozambique. Environments 2022, 9, 19. [CrossRef]
28. Muchimbane, A.B.D. Estudo dos Indicadores da Contaminação das Aguas Subterrâneas por Sistemas de Saneamento in Situ–
Distrito Urbano 4, Cidade de Maputo, Moçambique. Master’s Thesis, University of São Paulo, Instituto de Geociências, São
Paulo, Brasil, 2010. [CrossRef]
29. Serra, C. Da Problemática Ambiental à Mudança: Rumo à Um Mundo Melhor; Editora Escolar: Maputo, Mozambique, 2012;
ISBN 9789896700300.
30. INE (Instituto Nacional de Estatistica). Boletim de Estatísticas Demográficas e Sociais, Maputo Cidade 2019. Available online:
http://www.ine.gov.mz/estatisticas/estatisticas-demograficas-e-indicadores-sociais/boletim-de-indicadores-demograficos-
22-de-julho-de-2020.pdf/at_download/file (accessed on 10 January 2022).
31. VOA. Desabamento de Lixeira Deixa 17 Mortos em Maputo. Voice of America News. Available online: https://www.
voaportugues.com/a/desabamento-lixeira-17-mortos-maputo/4260624.html (accessed on 30 September 2021).
32. Ferrão, D.A.G. Evaluation of Removal and Disposal of Solid Waste in Maputo City, Mozambique. Master’s Thesis, University of
Cape Town, Cape Town, South Africa, 2006. Available online: http://hdl.handle.net/11427/4851 (accessed on 10 August 2021).
33. Sarmento, L.; Tokai, A.; Hanashima, A. Analyzing the structure of barriers to municipal solid waste management policy planning
in Maputo city, Mozambique. Environ. Dev. 2015, 16, 76–89. [CrossRef]
34. Matsinhe, F.O.; Paulo, M. Estudo etnográfico sobre os catadores de lixo da lixeira de hulene (Maputo). Cad. Afr. Contemp. 2020, 3.
35. Palalane, J.; Segala, I.; Opressa, I. Urbanização e Desenvolvimento Municipal em Moçambique: Gestão de Resíduos Sólidos; Instituto
Brasileiro de Administração Municipal, Área de Desenvolvimento Urbano e Meio Ambiente: Rio de Janeiro, RJ, Brazil, 2008; p. 12.
Available online: https://www.scribd.com/document/419123335/Gestao-de-Residuos-Solidos-Em-Mocambique (accessed on
7 August 2021).
36. Vicente, E.M. Aspects of the Enginering Geologic of Maputo City. Ph.D. Thesis, School of Agricultural, Earth and Environmental
Sciences, Pietermaritzburg. Soth Africa, 2011; pp. 22–38. Available online: http://hdl.handle.net/10413/8078 (accessed on
10 January 2022).
37. Vicente, E.M.; Jermy, C.A.; Schreiner, H.D. Urban geology of Maputo, Mocambique. Geol. Soc. 2006, 338, 1–13. Available online:
https://citeseerx.ist.psu.edu/viewdoc/download?doi=10.1.1.606.7220&rep=rep1&type=pdf (accessed on 3 January 2022).
38. Momade, J.T.; Ferrara, F.J.; Oliveira, M. Notícia Explicativa da Carta Geológica 2532 Maputo; Escala: Maputo, Mozambique, 1996.
(In Portuguese)
39. CIAT. Climate-Smart Agriculture in Mozambique; International Center for Tropical Agriculture: Cali, Colombia, 2017; pp. 1–25.
40. Dos Muchangos, A. Paisagens e Regiões Naturais, Maputo; Editora Escolar: Lisboa, Maputo, 1999; pp. 5–163. (In Portuguese)
41. Koda, E.; Tkaczyk, A.; Lech, M.; Osiński, P. Application of electrical resistivity data sets for the evaluation of the pollution
concentration level within landfill subsoil. Appl. Sci. 2017, 7, 262. [CrossRef]
42. Lau, A.M.P.; Ferreira, F.J.F.; Stevanato, R.; da Rosa Filho, E.F. Geophysical and physicochemical investigations of an area
contaminated by tannery waste: A case study from southern Brazil. Environ. Earth Sci. 2019, 78, 517. [CrossRef]
43. Akhtar, J.; Sana, A.; Tauseef, S.M.; Chellaiah, G.; Kaliyaperumal, P.; Sarkar, H.; Ayyamperumal, R. Evaluating the groundwater
potential of Wadi Al-Jizi, Sultanate of Oman, by integrating remote sensing and GIS techniques. Environ. Sci. Pollut. Res. 2022,
1–12. [CrossRef] [PubMed]
44. Bernardo, B.; Candeias, C.; Rocha, F. Application of geophysics in geo-environmental diagnosis on the surroundings of the
hulene-b waste dump, Maputo, Mozambique. J. Afr. Earth Sci. 2022, 185, 104415. [CrossRef]
45. Adamo, N.; Al-Ansari, N.; Sissakian, V.; Laue, J.; Knutsson, S. Geophysical methods and their applications in dam safety
monitoring. J. Earth Sci. Geotech. Eng. 2020, 11, 291–345. [CrossRef]
46. Geotomo. RES2DINV ver. 3.59—Rapid 2-D Resistivity & IP Inversion using the Least-Squares Method Wenner, Dipole-
Dipole, Inline Pole-Pole, Pole-Dipole, Equatorial Dipole-Dipole, Offset Pole-Dipole, Wenner-Schlumberger, Gradient and Non-
Conventional Arrays, 2010, 1–148. Available online: http://epsc.wustl.edu/~{}epsc454/instruction-sheets/Res2dinv03.59.pdf
(accessed on 10 January 2022).
Water 2022, 14, 1746 17 of 18

47. Asfaw, D.; Mengistu, D. Modeling megech watershed aquifer vulnerability to pollution using modified DRASTIC model for
sustainable groundwater management, Northwestern Ethiopia. Groundw. Sustain. Dev. 2020, 11, 100375. [CrossRef]
48. Anshumala, K.; Shukla, J.P.; Patel, S.S.; Singh, A. Assessment of groundwater vulnerability zone in mandideep industrial area
using DRASTIC model. J. Geol. Soc. India 2021, 97, 1080–1086. [CrossRef]
49. Hosseini, M.; Saremi, A. Assessment and estimating groundwater vulnerability to pollution using a modified DRASTIC and
GODS models (Case study: Malayer plain of Iran). Civ. Eng. J. 2018, 4, 433. [CrossRef]
50. Kozłowski, M.; Sojka, M. Applying a modified DRASTIC model to assess groundwater vulnerability to pollution: A case study in
Central Poland. Pol. J. Environ. Stud. 2019, 28, 1223–1231. [CrossRef]
51. Hasan, M.A.; Ahmad, S.; Mohammed, T. Groundwater contamination by hazardous wastes. Arab. J. Sci. Eng. 2021, 46, 4191–4212.
[CrossRef]
52. Ersoy, A.; Gültekin, F. DRASTIC-based methodology for assessing groundwater vulnerability in the Gümüşhacıköy and Merzifon
basin (Amasya, Turkey). Earth Sci. Res. J. 2013, 17, 33–40. Available online: https://www.scielo.org.co/scielo.php?pid=S1794-61
902013000100006&script=sci_arttext&tlng=en (accessed on 15 January 2022).
53. Paul, S.; Surabhi, C. An investigation of groundwater vulnerability in the North 24 parganas district using DRASTIC and
hybrid-DRASTIC models: A case study. Environ. Adv. 2021, 5, 100093. [CrossRef]
54. Parvin, F.; Tareq, S.M. Impact of landfill leachate contamination on surface and groundwater of Bangladesh: A systematic review
and possible public health risks assessment. Appl. Water Sci. 2021, 11, 100. [CrossRef] [PubMed]
55. Wysocka, M.E.; Zabielska-Adamska, K. Impact of Protective Barriers on Groundwater Quality. In Proceedings of the 10th
International Conference: Environmental Engineering, Vilnius, Lithuania, 27–28 April 2017. [CrossRef]
56. Netto, L.G.; Filho, W.M.; Moreira, C.A.; di Donato, F.T.; Helene, L.P.I. Delineation of necroleachate pathways using electrical
resistivity tomography (ERT): Case study on a cemetery in Brazil. Environ. Chall. 2021, 5, 100344. [CrossRef]
57. Yap, C.K.; Chew, W.; Al-Mutairi, K.A.; Nulit, R.; Ibrahim, M.H.; Wong, K.W.; Bakhtiari, A.R.; Sharifinia, M.; Ismail, M.S.; Leong,
W.J.; et al. Assessments of the ecological and health risks of potentially toxic metals in the topsoils of different land uses: A case
study in Peninsular Malaysia. Biology 2022, 11, 2. [CrossRef]
58. Chetri, J.K.; Reddy, K.R. Advancements in municipal solid waste landfill cover system: A review. J. Indian Inst. Sci. 2021, 1,
557–588. [CrossRef]
59. Brahmi, S.; Baali, F.; Hadji, R.; Brahmi, S.; Hamad, A.; Rahal, O.; Zerrouki, H.; Saadali, B.; Hamed, Y. Assessment of groundwater
and soil pollution by leachate using electrical resistivity and induced polarization imaging survey, case of Tebessa municipal
landfill, NE Algeria. Arab. J. Geosci. 2021, 14, 249. [CrossRef]
60. Ololade, O.O.; Mavimbela, S.; Oke, S.A.; Makhadi, R. Impact of leachate from northern landfill site in bloemfontein on water and
soil quality: Implications for water and food security. Sustainability 2019, 11, 4238. [CrossRef]
61. Marques, T.; Matias, M.S.; Silva, E.F.D.; Durães, N.; Patinha, C. Temporal and spatial groundwater contamination assessment
using geophysical and hydrochemical methods: The industrial chemical complex of Estarreja (Portugal) case study. Appl. Sci.
2021, 11, 6732. [CrossRef]
62. Feng, S.J.; Wu, S.J.; Fu, W.D.; Zheng, Q.T.; Zhang, X.L. Slope stability analysis of a landfill subjected to leachate recirculation and
aeration considering bio-hydro coupled processes. Geoenvironment. Disasters 2021, 8, 29. [CrossRef]
63. Udosen, N.I. Geo-electrical modeling of leachate contamination at a major waste disposal site in south-eastern Nigeria. Model
Earth Syst. Environ. 2022, 8, 847–856. [CrossRef]
64. Morita, A.K.M.; Ibelli-Biancoa, C.; Jamil, A.A.; Jaqueline, A.; Pelinson, C.N.; Nobrega, J.; Rosalema, L.M.P.; Leitea, C.M.C.;
Niviadonski, L.M.; Manastella, C.; et al. Pollution threat to water and soil quality by dumpsites and non-sanitary landfills in
Brazil: A review. Waste Manag. 2021, 131, 163–176. [CrossRef] [PubMed]
65. Oke, S.A. Regional aquifer vulnerability and pollution sensitivity analysis of drastic application to Dahomey basin of Nigeria. Int.
J. Environ. Res. Public Health 2020, 17, 2609. [CrossRef] [PubMed]
66. Boumaiza, L.; Walter, J.; Chesnaux, R.; Brindha, K.; Elango, L. An operational methodology for determining relevant DRASTIC
factors and their relative weights in the assessment of aquifer vulnerability to contamination. Environ. Earth Sci. 2021, 80, 281.
[CrossRef]
67. Nasri, G.; Hajji, S.; Aydi, W.; Boughariou, E.; Allouche, N.; Bouri, S. Water vulnerability of coastal aquifers using AHP and
parametric models: Methodological overview and a case study assessment. Arab. J. Geosci. 2021, 14, 59. [CrossRef]
68. Tan, M.; Wang, K.; Xu, Z.; Li, H.; Qu, J. Study on heavy metal contamination in high water table coal mining subsidence ponds
that use different resource reutilization methods. Water 2020, 12, 3348. [CrossRef]
69. Blarasin, M.; Matiatos, I.; Cabrera, A.; Lutri, V.; Giacobone, D.; Quinodoz, F.B. Characterization of groundwater dynamics and
contamination in an unconfined aquifer using isotope techniques to evaluate domestic supply in an urban area. J. S. Am. Earth
Sci. 2021, 110, 103360. [CrossRef]
70. WHO. Guidelines for Drinking-Water Quality, 4th ed.; Incorporating the First Addendum; WHO: Geneva, Switzerland, 2017;
ISBN 978-92-4-154995-0.
71. Akinbile, C.O.; Yusoff, M.S. Environmental impact of leachate pollution on groundwater supplies in Akure, Nigeria. Int. J.
Environ. Sci. Dev. 2011, 2, 81–86. [CrossRef]
Water 2022, 14, 1746 18 of 18

72. Pan, S.; Dixon, K.L.; Nawaz, T.; Rahman, A.; Selvaratnam, T. Evaluation of galdieria sulphuraria for nitrogen removal and
biomass production from raw landfill leachate. Algal Res 2021, 54, 102183. [CrossRef]
73. Isiuku, B.O.; Enyoh, C.E. Pollution and health risks assessment of nitrate and phosphate concentrations in water bodies in South
Eastern, Nigeria. Environ. Adv. 2020, 2, 100018. [CrossRef]
Article

Integration of Electrical Resistivity and Modified DRASTIC


Model to Assess Groundwater Vulnerability in the
Surrounding Area of Hulene-B Waste Dump, Maputo,
Mozambique
Bernardino Bernardo 1,2, Carla Candeias 1 and Fernando Rocha 1,*

1 GeoBioTec Research Centre, Department of Geosciences, University of Aveiro, 3810-193 Aveiro, Portugal;
bernardino.bernardo@ua.pt (B.B.); candeias@ua.pt (C.C.)
2 Faculty of Earth Sciences and Environment, Pedagogic University of Maputo, Av. do Trabalho, Maputo 2482,

Mozambique
* Correspondence: tavares.rocha@ua.pt

Figure S1. Sampling and chemical analysis of well water S1. (a) Water collection - South well (b) Water collection -
North well; (c) Chemical analysis process (d) Result of phosphate analysis.
254
255

13.2.3 Appendix 4. Paper 3


256
Environmental Earth Sciences (2022) 81:542
https://doi.org/10.1007/s12665-022-10672-7

ORIGINAL ARTICLE

Soil properties and environmental risk assessment of soils


in the surrounding area of Hulene‑B waste dump, Maputo
(Mozambique)
Bernardino Bernardo1,2 · Carla Candeias1 · Fernando Rocha1

Received: 12 April 2022 / Accepted: 26 November 2022


© The Author(s), under exclusive licence to Springer-Verlag GmbH Germany, part of Springer Nature 2022

Abstract
Soils in areas surrounding landfills are constantly being enriched by heavy metals contained in the leachates, which can
subsequently migrate to groundwater. The present investigation aims to characterize soil properties of 71 soil samples col-
lected in the surroundings of Hulene-B waste dump and to determine the landfill pollution index (Ip). Soils properties studied
were texture, pH, electrical conductivity, organic matter, color, and moisture. Results revealed that soils properties in the
surroundings of Hulene-B waste dump were significantly altered when compared to local background. Ip index classified
these soils with very high pollution, indicating a possible migration of contaminants to subsoil and groundwater, suggesting
the need for intervention to mitigate the impact.

Keywords Soils · Waste dump · Soil properties · Environmental risk assessment

Introduction In urban areas of developing countries, soil and water


contamination are often associated with poor solid waste
Solid waste disposal surrounding areas are prone to envi- disposal mechanisms (Chu et al. 2020; Parvin and Tareq
ronmental contamination, especially when not planned or 2021; Helene and Moreira 2021). The rate of soil and water
monitored (Morita et al. 2020). Wastes produce leachates, an contamination around landfills is controlled by factors such
effluent from biological decomposition and dissolved miner- as landfill age, relevant for leachate production linked to
als that can contaminate the environment (soil and water) climatic conditions, lithology, and waste disposal condi-
posing a risk to humans and the environment (Kapelewska tions (Hussein et al. 2021; Wijekoon et al. 2022). Tropical
et al. 2019; Gonçalves et al. 2019). Soil is the first environ- soils, where oxides (Fe, Al, and Mn) and kaolinitic minerals
mental component affected by leachate flows around waste prevail, have a potential for metal ion uptake contained in
disposal sites (Rapti-Caputo et al. 2006). Huang et al. (2020) leachates (Saentho et al. 2022).
suggested that soil is an important receptor for pollutants Soils have specific properties, indicative of contamina-
generated from human activities, acting as a sink for metals. tion (Khomiakov 2020; Zamulina et al. 2021). These prop-
O’Riordan et al. (2021) revealed that soils act as contami- erties represent a major role in the fixation, adsorption, and
nants filter and can control the transport of chemical ele- absorption of heavy metals (Feng et al. 2021; Yap et al.
ments, and other substances, to the hydrosphere, atmosphere 2022). Factors such as pH, organic matter (OM), moisture,
and biota. color, texture, and electrical conductivity (EC), are pointed
as relevant to assess contamination processes (Fatoba et al.
2021; Hussein et al. 2021). Soil pH has a major influence
on heavy metal solubility (Sparling 2020). Slight to strong
* Carla Candeias alkaline soil pH around dumps have been reported to favor
candeias@ua.pt potentially toxic elements (PTEs; e.g., Cr, Mo) dissolution
1
GeoBioTec Research Centre, Geosciences Department, and mobility (Awa and Hadibarata 2020; Zhang and Rear-
University of Aveiro, 3810-193 Aveiro, Portugal don 2003).
2
Faculty of Earth Sciences and Environment, Pedagogic Soil texture also represents an important role in migra-
University of Maputo, Maputo, Mozambique tion and retention of heavy metals (Kosheleva et al. 2018).

13
Vol.:(0123456789)
542 Page 2 of 14 Environmental Earth Sciences (2022) 81:542

In general, sandy soils are propitious to leachates migration Tenodi et al. (2020) and Mama et al. (2021) proposed the
in depth, promoting subsoil and groundwater contamina- study of soil properties to assess the adsorption, absorption,
tion (Victor et al. 2019; Makuleke et al. 2020). Soil OM and migration processes of metals associated with leachate
influences mobility and speciation of heavy metals, where migration. Previous studies by Vicente et al. (2006) and
complexation reactions modify their accumulation potential Serra (2012) in Maputo city (Mozambique) suggested that
(Kennou et al. 2015). Soils with low OM content in areas inadequate disposal of urban solid wastes is one of the main
with contamination sources such as Pb, Cu, Hg, and Zn causes of soil contamination. Hulene-B waste dump is the
are indicative of possible leaching and migration at depth largest open-air dump in Maputo. Bernardo et al. (2022a)
(Zhang and Reardon 2003). studies showed that Hulene-B surrounding soil and subsoil
One of the parameters affecting soil color is OM con- presented electrical resistivity variations, establishing a link
tent (USDA 2001). In solid waste contaminated areas, soil to leachate plumes contamination, with origin in the waste
color is influenced by leachates, which are a potential for dump. The present study aims to characterize soils prop-
metal-organic interactions through organic ligands (Lee erties in the surrounding of Hulene-B waste dump, and to
et al. 2022). Soil color has been used to indicate the possible assess pollution risk index.
interference of leachates in soils properties (Gonçalves et al.
2019). Darker colors can be associated with soils affected by
leachates (Naveen et al. 2017) and whitish colors are often Materials and methods
indicative of leaching processes (Lee et al. 2022). Regard-
ing soil moisture, Salam et al. (2019) studied its effect on Study area
the bioavailability of Cu and Pb in contaminated soils and
observed that lowest bioavailability of Pb and Cu occurred Hulene B waste dump is located in the northern area of
with higher soil moisture content. Maputo city (Mozambique) (Fig. 1). The dump occupies an
Soil electrical conductivity (EC) is a measure of soil abil- area of ~ 17 hm2 (Serra, 2012), being surrounded by dwell-
ity to conduct an electric current, being influenced by, e.g., ings, with over 75,000 inhabitants (INE 2020). Initially, is
moisture, OM, texture, and others (Grisso et al. 2009). In was an open quarry, with an estimated variable depth rang-
general, sandy soils present EC ≤ 100 μS/cm (Lund 2008; ing 6–15 m, that was abandoned in 1973 when waste started
USDA 2014). In areas surrounding waste dumps, high EC to be deposited (Ferrão 2006; Palalane and Segala 2008).
is associated to the enrichment of soils by contaminants, The dump receives all type of wastes produced in Maputo,
mainly leachates, responsible for the addition of metal ions e.g., industrial, domestic, commercial, hospital, estimated
to soils (Gonçalves et al. 2019; Andaloussi et al. 2021). at > 1000 t/daily (Sarmento et al. 2015; Serra 2012). In the

Fig. 1 Studied area context with


location of a dune depression,
b Hulene-B waste dump, (c, d)
not-impacted areas

13
Environmental Earth Sciences (2022) 81:542 Page 3 of 14 542

last 10 years, waste has been submitted to bulldozer compac- (2002). Soils color was determined using Munsell (2009)
tion and a leachate drainage system on southern, northern, soil chart. Silt fraction granulometric distribution was deter-
and western limits of the dump was added. However, due to mined with a Micromeritics Sedigraph III Plus grain size
its relatively poor structure, leachates are dispersed to the analyzer (UAVR). This technique determines the relative
western area, particularly during rainy season. mass distribution of a sample by particle size and is based
Geologically, the area belongs to Mesocenozoic sedimen- on two physical principles: sedimentation theory (Stokes’
tary basin of southern Mozambique, in contact with two law) and the absorption of X-radiation (Beer–Lambert law).
lithologies, the Ponta Vermelha (TPv) and the Malhazine The analytical accuracy and precision of the methods were
(QMa) formations (Afonso 1978; Momade et al. 1996). determined using analyses of reference materials and dupli-
Topography is of dune type and the dump is located on a cate samples in each analytical set. Results were within 95%
slight slope with E–W orientation (Momade et al. 1996). confidence limits of the recommended values for the certi-
The soils are heterogeneous, comprising materials of the fied material. The relative standard deviation was between
upper Pleistocene QMa and upper Pleistocene to lower Pleis- 5 and 10%.
tocene TPv formations (Momade et al. 1996). QMa forma- pH classes were defined in accordance with USDA
tion consists of coarse to fine, poorly consolidated sands (1998): 6.6–7.3 neutral, 7.4–7.8 slightly alkaline, and
with whitish to reddish colors, fixed by vegetation and by 7.9–8.4 moderately alkaline. Organic matter classes were
successive consolidation processes while, TPv comprises defined in accordance with USDA (2001): > 4% high con-
ferruginous sandstone and red silty sands, which gradually tent, 2–4% medium content, and < 2% low content. Electri-
become underneath to yellow and whitish sands (Momade cal conductivity classes were defined based on internation-
et al. 1996). ally adopted reference values for sandy soils (Lund, 2008;
Predominant climate is of subtropical type, with mean USDA, 2011): ≤ 100 μS/cm natural medium conductivity,
annual precipitation of ~ 789 mm, with two climatic seasons: 101–200 μS/cm high conductivity, and > 201 μS/cm very
(a) hot (mean 25 °C) and rainy period from December to high conductivity. Soil color classes were defined taking in
March, representing > 60% of annual precipitation, with its consideration color intensity: blackish, grayish, and brown-
peak in January (~ 125 mm); and (b) dry and cold season, ish. Moisture classes were defined according to the aver-
from April to September, with lower temperatures in June age of the local background value considered standard,
and July (mean 21 °C), and scarce precipitation, with mini- 0.41–0.82% high and 0.82–1.9% very high.
mum values in August (~ 12 mm) (CIAT, 2017). The prevail-
ing winds are SE (Muchangos 1999). Statistical analysis

Sampling Descriptive statistics, principal component analysis (PCA)


and Spearman correlation were performed using SPSS® v.25
A total of 71 soil samples were collected in the surround- software (IBM, USA). PCA is a data dimensionality reduc-
ing area of Hulene-B dump, in January 2020 (Fig. 1). To tion technique that aims to explain most of data variation
determine background, 10 samples were collected in areas with a small number of independent variables called “prin-
considered not impacted by the dump. Soil samples were cipal components” (Škrbić and Đurišić-Mladenović 2010;
collected at 0–20 cm depth. Samples were georeferenced Hou et al. 2017). Promax rotation was performed which
and preserved in plastic bags until laboratorial treatment. allowed the variance of analyzed parameters to be corre-
On the laboratory, samples were oven dried ~ 40 °C (Maputo lated with each other. Spearman’s correlation was used to
Pedagogical University, Mozambique). Afterwards, samples assess the strength of pairwise correlation of the analyzed
were transported to GeoBioTec Research Center Laborato- soil properties.
ries (University of Aveiro (UAVR), Portugal), for analyses.
Risk assessment
Laboratorial analyses
The landfill pollution risk index (Ip) was proposed by Man-
Soil samples were sieved to achieve the < 2000 (sand) cini et al. (1999), is based on the determination of vulner-
and < 63 (silt) µm fractions. pH was determined in the two ability to aquifer pollution and hazard induced by the quanti-
fractions with a 1:2.5 soil/water solution using a pH meter. fied landfill as a specific parameter that allows (i) to identify
Electrical conductivity (EC) was measured under the same suitable sites to host new waste digestion sites, and (ii) to
conditions as pH in the two fractions, using a high-resolution define the priorities in the control and remediation opera-
conductivity meter. Organic matter (OM) content was deter- tions to be fulfilled in case of potential hazard landfills or
mined with method proposed by USDA (2001). Moisture sites that have already been compromised (Rapti-Caputo
was measured based on the procedures defined by Reeuwijk et al. 2006). The integration of Ip assessment has been used

13
542 Page 4 of 14 Environmental Earth Sciences (2022) 81:542

in many studies to evaluate the risk of groundwater contami- 2019), where Ip corresponds pollution risk index; Wi is
nation and understanding the properties of the soils where a weight of the risk parameter (1–5), and Ri the risk factor.
landfill is located and its surroundings has been highlighted Thus, Ip index < 3 represents low pollution potential; Ip
as relevant once soil is considered a superficial defense of 3–7 suggest risk and medium vulnerability; Ip 7–9 indicate
the hydrogeological system, where several important pro- risk and high vulnerability, and Ip > 10, high contamination
cesses take place within the soil that make up the attenuation risk and immediate intervention measures should be taken
capacity (Civita and Maio 2004). to mitigate the impact on groundwater (Rapti-Caputo et al.
The method applies eighth variables to assess risk index: 2006; Liu et al. 2019).
(a) volume of deposited waste; (b) leachate drainage; (b)
type of waste; (d) waste physical condition; (e) waste bio-
degradability; (f) monitoring system; (g) waste compac-
tion; and (h) final coating (Chaudhary et al. 2021; Liu Results and discussion
et al. 2019; Nadiri et al. 2017). Risk factor (Ri) and risk
weight parameter (Wi) for each of the eight variables are Studied soil and background samples were classified as sand
described in Table 1. Landfill risk index was determined (Fig. 2). Sand fraction ranged 90.52–99.32% in studied sam-
using Ip = ni=1 WixRi (Rapti-Caputo et al. 2006; Liu et al. ples, and 92.84–94.20% in background samples (Table 2).

Table 1 Landfill risk index Variables Weight Single risk elements Reduction
assessment (Rapti-Caputo et al. factor (R)
2006)
Volume of waste deposited 5 < 10 t/day 1
10 a 50 t/day 0.2
50 a 500 t/day 0.4
> 500 t/day 1
Leachate drainage system 5 External and internal drainage 0.1
Internal drainage 0.3
Reuse of leachate in the system 0.5
Absent drainage 1
Type of waste 3 Inert 0.1
Urban 0.5
Industrial—non-hazardous 0.8
Dangerous 1
Physical state of waste 3 Solidified with inert matrix 0.1
Solid 0.2
Mud with humidity < 70% 0.5
Stabilization typology 2 Mud with humidity > 70% 1
Non-biodegradable 0.1
Aerobic 0.3
Aerobic and anaerobic 0.5
Anaerobic 1
Monitoring system 2 Well and geomembrane monitoring 0.1
Geomembrane 0.3
Monitoring well 0.5
Absent 1
Compacting the waste 1 Compacted with pneumatic equipment 0.1
Compacted with bulldozer 0.2
Manually compacted 0.5
No compaction 1
Final coating 1 Compacted soil with clay 0.1
Compacted clay 0.2
Non-compacted soil 0.5
Absent 1

13
Environmental Earth Sciences (2022) 81:542 Page 5 of 14 542

2020). Sparling (2020) suggested that soil pH has a major


influence on heavy metals solubility. The pH of the studied
71 soil samples sand fraction ranged 6.7–8.01, with a mean
of 7.4; and in silt fraction varied between 7.0 and 8.4, with
mean 7.8 (Fig. 4). Spatially, soils pH in the two fractions,
were classified as moderately alkaline (7.9–8.4) to slightly
alkaline (7.4–7.8). Moderately alkaline class were found in
sand fraction samples towards the west of the dump while
silt fraction, in samples in the immediate edge of the dump.
Neutral pH (6.6–7.3), in silt fraction, was found in samples
across the western strip of the dump, while in sand frac-
tion in the northwest direction of the dump. Background
samples presented acid pH, ranging 4.2–6.3, and mean 5.25
in sand fraction, and in 5.7–6.3, in silt fraction. Studies sug-
gested that heavy metals mobility tends to be lower with
increasing pH, except for metal(loid)s As, Mo and Se (Zimik
et al. 2021). Alkaline pH decreases Cu, Co, Fe, Mn, and Zn
bioavailability, because of low solubility in this pH range,
Fig. 2 Soil texture classification of studied soils (blue) and back-
while Mo, Se, V, As and Cr are more available (Alexakis
ground (yellow) samples
2021). In soils with pH > 6, occurs the dissociation of H+
from OH groups in organic matter and Fe and Al oxides,
An One-way Anova analysis showed significant differences increasing metal adsorption and subsequent precipitation,
between studied soils and background samples in sand and reducing their bioavailability (Chen et al. 2022; Choppala
clay fraction (p = 0.000). The spatial distribution of sand, et al. 2018). Odom et al. (2021) found high levels of Zn in
silt and clay fractions is presented in Fig. 3, being the scale soils around Dompoase landfill in Ghana, under alkaline pH
divided in classes of 20% of each fraction range. Samples conditions (8.1) and associated it with strong soil enrichment
with higher percentage of coarser particles are distributed in by leachates. El Fadili et al. (2022) reported that soils in
the center of the studied area, corresponding to remobilized the surroundings of Benguerir landfill presented high levels
soils for construction purposes. Lower sand percentages of contamination by Zn and Cu and associated the process
were found in samples collected in the dump limits. Dakheel of adsorption of these metals with alkaline pH. In tropical
et al. (2022) suggested that areas with nearby contamina- soils, such as those in the surroundings of the Hulene-B
tion sources, sandy soils are more vulnerable to leaching dump, heavy metal retention can occur under high pH con-
and vertical migration of heavy metals to groundwater. Silt ditions (Campos 2010), due to the predominance of oxidic
fraction, ranged 0.44–7.69%, presented higher contents in (Al, Fe and Mn) and kaolinitic mineralogy in the clay frac-
samples collected near the dump (Fig. 3). Clay and silt frac- tion, which increases the metal adsorption capacity (Alleoni
tions presented similar patterns, with higher clay contents et al. 2005).
in samples on the dump proximity. This tendency can be Soil samples EC ranged 43.1–725 µS/cm in sand frac-
associated to leaching due to topographic context, with E–W tion and 37.5–217 µS/cm in silt fraction (Fig. 5). All sam-
leachates circulation (Bernardo et al. 2022a, b, c). ples showed higher EC values when compared to average
Soil pH applies to the concentration of ions (H+) pre- background samples, of 17.7 μS/cm for sand fraction, and
sent in the soil solution, being strongly attracted to negative 13.9 μS/cm for silt fraction. Sand fraction, with minimum
charges and have capability to replace other cations (Altaf EC 43.1 μS/cm was found on the northwest area and maxi-
et al. 2021; Soubra et al. 2021). This parameter control soil mum of 725 μS/cm in the immediate dump boundary to
adsorption and distribution of heavy metals (Naveed et al. the northwest of the dump. Spatially, the lowest values

Table 2 Granulometric Fraction Studied soils (n = 71) Background (n = 10)


statistical information (in %)
Min Max Mean ± SD Min Max Mean

Sand 90.52 99.32 97.03 ± 1.89 92.84 95.36 94.20 ± 0.78


Silt 0.44 7.69 2.112 ± 1.41 2.27 3.36 2.74 ± 0.34
Clay 0.22 2.60 0.85 ± 0.52 2.31 3.80 3.06 ± 0.45

SD standard deviation

13
542

13
Page 6 of 14

Fig. 3 Granulometric spatial distribution of sand, silt and clay fractions of the studied soils (in %; n = 71)
Environmental Earth Sciences
(2022) 81:542
Environmental Earth Sciences (2022) 81:542 Page 7 of 14 542

Fig. 4 Soil samples pH in sand and silt fractions

Fig. 5 Soil samples EC in sand and silt fractions

(43.1–100 μS/cm) were randomly distributed, in the center (201–725 μS/cm) were found in samples on the immediate
and north of the sampling area and well evidenced in a strip western boundary and with some dispersion to the south-
parallel to the dump, corresponding to a disturbed area dur- west, west and in the northern direction of the sampling
ing construction of the leachate drainage channel. EC rang- area. Silty fraction, revealed minimum EC 37.5 μS/cm in
ing 101–200 μS/cm, were found in samples collected near samples at the northern boundary, corresponding to a transi-
the dump, mainly at the eastern end of the dump towards tion area between dump upper and lower parts (character-
the center and north of the main sampling area. Higher EC ized by intense leaching) and can also be associated with

13
542 Page 8 of 14 Environmental Earth Sciences (2022) 81:542

the heavy rainfall of the sampling campaign, promoting the deposited changes soils color and enriches them with Hg.
dissolution of salts in the soil and consequently a decrease Color and OM showed a similar trend, being good indicators
in soil EC (USDA 2011; Chaaou et al. 2022). Maximum for locating areas with possible contamination (Dregulo and
EC 217 μS/cm was found in the northern direction of the Bobylev 2020).
sampling center. Spatially lower EC (37.5–100 μS/cm) was Studied soil samples moisture content is presented in
predominant throughout the sampling area, followed by Fig. 6. Background samples mean content was 0.4%, ranging
EC varying 101–200 μS/cm, predominant at the immediate 0.2–0.6%. Samples from dump outskirts revealed an average
western boundary and sporadically on the south and north moisture content of 0.4%, ranging 0.0–1.9%. Samples col-
of the sampling area. In both fractions, samples collected lected in the immediate edge of the dump showed very alter-
on the immediate western boundary exhibited higher EC, nating moisture contents, with a predominance of high and
suggesting an impact of waste discharges on surrounding very high values (0.4–1.9%). In the center of sampling area,
soils. The eastern boundary of Hulene-B dump was charac- occurred a prevalence of moisture contents < 0.4%, alter-
terized by high conductivity (Bernardo et al. 2022c), areas nating with higher 0.4–0.8%. One sample in the immediate
with shallow phreatic levels, which may be a factor in the north-western boundary of the dump, showed the highest
capillary rise of salts and water into the topsoil and conse- concentration of 1.9%, being the lowest concentration of
quently an EC increase (USDA 2011). Hussein et al. (2021), 0.0% in the southern boundary and central sampling area.
found that soils with high EC values were associated with In general, sandy soils have low water retention capacity
high soil contamination, by As, Cd, Pb and Cr in a study of 6 (Zeitoun et al. 2021) and in the surroundings of dumps with
dumpsites in Malaysia. Similar findings by Wu et al. (2021), surface drainage of leachate whose substrate is sandy soils,
on a study of waste dumps in China, and Fatoba et al. (2021) the process of leaching and migration of leachate in depth
on a waste dump in Nigeria, with contamination by metal- predominates (Zhang et al. 2021), being pointed as the
ion enriched leachates. cause of subsoil and groundwater contamination in many
Soil samples OM distribution is presented in Fig. 6. In studies (Stefania et al. 2019; Wu et al. 2021; Przydatek and
studied samples, OM content is scarce (mean 1.1%), with Kanownik 2021).
few samples with > 2% in soils collected closer to the dump,
in the eastern and western sections, and in western area limit
(> 2%, maximum of 4.2%). The higher OM content, suggest Principal component analysis (PCA)
an enrichment by contaminants resulting from the migration
of surface leachate from the dump and/or by the aeolian dep- Principal component analysis reduces a set of variables into
osition of ash from waste burning (Nisari et al. 2021). Lower a smaller one called principal components (PC), attempting
OM content was found in the center of the sampling area, to reveal variables correlation structure and interpret param-
corresponding to a depression, that might increase OM lea- eters that influence soil samples. In this study, PCs were
chate and surface water migration to the east–west direction, extracted from 71 samples (sand fraction) and 8 variables
especially during rainy season. Samples OM content, in the with Promax rotation. Three principal components were
western end of the sampling, may be related to vegetation. considered, accounting for 73.14% of the total variance.
Soil samples color distribution is presented in Fig. 6. First component (PC1), explained 40.86% of total variance,
Blackish samples were found predominantly in areas near defined two groups of variables: sand negatively related to
the landfill site and the northwest. Brown samples were silt and clay variables (Fig. 7), with high negative loadings
predominant in the immediate eastern, western, and south- (> 0.95). Second component (PC2) explained 21.61% of
western boundaries of the landfill. Greyish samples were total variance, with positive loading composed by OM, EC,
more prevalent throughout the western sampling area. and moisture. Third component (PC3), with 13.66% of vari-
Factors such as plants fixation, may contribute to darker ance, groups pH and color variables.
colors in some samples distant from the dump. Color has The opposition in PC1 between sand and smaller frac-
been associated to OM content, which exhibits an affinity tions, was justified by the high sand fraction content. PC2
for heavy metal uptake (Dregulo and Bobylev 2020; Spar- groups OM, EC, and moisture, influenced by incorporation
ling 2020). The background samples revealed reddish color effects of the metallic ions contained in the leachates, as
in the TPv Formation and reddish brown in the QMa For- well as deposition of incineration ashes of contaminated
mation samples. Studies reported that the sandy soils may waste from the Hulene-B dump. Samples from the immedi-
have color changes, resulting from accumulation processes ate limits of Hulene-B dump showed a spatial distribution
of OM transported by water, which subsequently settle in that expressed the correlation between high EC, OM, and
the interstitial spaces, and may fixate some contaminants moisture. PC3, suggested heterogeneous alteration con-
(Seidl et al. 2021). Hussein et al. (2021) suggested that ash ditions associated to the movement and accumulation of
resulting from burning solid waste that is transported and leachates that cause heavy metal contamination, and, the

13
Environmental Earth Sciences
(2022) 81:542

Fig. 6 Soil samples OM content, color, and moisture content


Page 9 of 14
542

13
542 Page 10 of 14 Environmental Earth Sciences (2022) 81:542

Landfill risk index assessment (Ip)

The analysis of the geo-environmental and structural context


of the dump, allowed to assess risk factor variables weights
(Table 4). For the risk factor volume of deposited waste,
a maximum score of 1 and corresponding weight is 5 was
assigned, taking in consideration a deposition > 1000 t/day
(Table 1). For the leachate drainage system, a score of 1
and corresponding weight is 5 was assigned once leachate
is dispersed on soil in the surroundings of the dump (Fig. 8).
For the type of waste, a value of 0.5 was assigned and a
corresponding weight 3. For the stabilization typology was
considered 0.3 and a weight of 3, due to dump successive
accumulation of waste and a heterogeneous process of waste
biodegradation, and continuous accumulation causing a
reduction of oxygen in the layers below. The monitoring
system was assigned a maximum of 1, with a corresponding
Fig. 7 PCA variables distributed by component value of 2 since there is no monitoring mechanism at the
dump. The compaction of the waste is done by bulldozer,
so a value of 0.2 was assigned and a corresponding weight
predominance of sandy granulometry that is conducive to of 1. For the final coating, a value of 1 and a correspond-
vertical leaching of contaminants. ing weight of 1 was given since the dump has no coating
Spearman’s correlation (Table 3), in line with PCA mechanism.
showed the same trend. The pairs silt/sand, clay/sand, The factors with the highest scores were the volume of
moisture/sand, pH/clay, color/pH (p < 0.01), color/sand, and waste deposited (> 1000 t/day) and the conditions of its storage
color/EC (p < 0.05) revealed a significant negative correla-
tion. This suggests heterogeneous processes that can be asso-
ciated with the geo-environmental conditions surrounding Table 4 Risk index (Ip) of the studied area
the dump, such as predominance of leaching in sandy soils Risk factors Situation R W R*W
which is intensified by the circulation of leachate-enriched
Landfilled waste volume > 1000 t/day 1 5 5
surface water, responsible for significant EC increase and
Drainage system Drainage absents 1 5 5
soil color alteration. Pairs clay/silt, EC/OM, pH/sand, pH/
Type of waste Urban 0.5 3 1.5
EC, moisture/silt, moisture/clay, color/clay (p < 0.01) and
State of the waste Solid 0.2 3 0.6
moisture/OM (p < 0.05), revealed a significant positive
Waste biodegradability Anaerobic 1 2 2
correlation. The impact of Hulene-B waste dump on soil
Site monitoring Non-existent 1 2 2
properties, by incorporation of metal ions from leachate and
Compacting the waste with bulldozer 0.2 1 0.2
contaminated ash, alter soil properties in a combined way at
Final coating material Absent 1 1 1
the immediate limits of the dump and in a localized manner
Ip 17.3
at various points in the surroundings of the dump.
R risk factor, W weight parameter, R*W risk factor final weight

Table 3 Spearman correlation Sand Silt Clay OM EC pH Moisture

Silt − 0.968** 1
Clay − 0.952** 0.868** 1
OM − 0.126 0.151 0.135 1
EC 0.139 − 0.038 − 0.18 0.532** 1
pH 0.303** − 0.202 − 0.341** 0.149 0.527** 1
Moisture − 0.355** 0.377** 0.349** 0.282* 0.206 − 0.11 1
Color − 0.279* 0.176 0.348** − 0.148 − 0.266* − 0.367** 0.049

**p < 0.01


*p < 0.05

13
Environmental Earth Sciences (2022) 81:542 Page 11 of 14 542

Fig. 8 a Leachate drainage channels without isolation, b dispersed leachate, c uncontrolled leachate flow

(open deposition, no isolation of the leachates from the sur- contamination and conditions for leachate migration into
rounding environment) and monitoring. These factors were the subsoil and groundwater. Electrical conductivity
described in many studies as conducive to soil contamination showed more significant alterations, with values consid-
in the surroundings of the dumpsites (Rapti-Caputo et al. 2006; ered high in all samples around the dump, what can be
Lothe and Sinha 2017; Liu et al. 2019; Rapti et al. 2021). associated with possible contamination of soils by heavy
Risk factors analysis allowed to classify Ip on the surround- metals present in leachates and ashes from waste incin-
ings of the de Hulene-B dump as very high (17.3), an indicator eration. The predominance of sandy fraction suggested
of possible groundwater contamination and need of imme- vertical migration of contaminated leachate in depth. Soils
diate intervention for impact mitigation (Rapti-Caputo et al. pH classified from neutral to slightly alkaline, and a low
2006). Structural measures were adopted by local authorities OM content suggested a higher leaching and migration
along with the present study sampling campaign, such as the capacity of contaminants in depth. The landfill pollution
construction of a surface leachate drainage system that could index (Ip) was rated very high (16.3) suggesting a possible
have reduced the risk factor to 2.5 and the overall risk to 14.8. migration of contaminants to groundwater, a similar result
Nevertheless, the constructed system does not systematically to soil properties. Results highlight the importance of the
control leachate and does not isolate the surrounding environ- dump monitoring and the systematic assessment of soil
ment (soils, groundwater, surface water and cultivated fields and groundwater contamination levels.
in the vicinity of the dump) (Fig. 8). Studies are still needed to understand the temporal
Soils properties and landfill pollution risk assessment dynamics of soils properties once this study sampling
data combination suggested complementarity to understand took place during rainy season. It is also pertinent to
soil and groundwater contamination risk by leachate. Results perform chemical and mineralogical analysis of the sur-
showed that soil properties were strongly impacted by the rounding soils as well as assess groundwater and biota
landfill site, as well as spatial areas with greatest changes. The contamination.
assessment of the landfill pollution risk was relevant, as the
factors associated to the management of the landfill site that
Author contributions Conceptualization, BB, CC, FR; methodology,
influence the change of soil properties around the Hulene-B BB, CC; validation, BB, FR, CC; formal analysis, BB, CC; investiga-
dump were identified. The combination of these results proved tion, BB, CC, FR; writing—original draft preparation, BB; writing—
to be a knowledge base to be applied in the search for measures review and editing, BB, FR, CC; supervision, FR, CC; funding acqui-
to mitigate the impacts of the Hulene-B dump on soil proper- sition, BB and FR. All authors have read and agreed to the published
version of the manuscript.
ties and the risk of groundwater contamination.
Funding This work was partially supported by GeoBioTec (UID/
GEO/ 04035/2019 + UIDB/04035/2020) Research Centre, funded by
Conclusion FEDER funds through the Operational Program Competitiveness Fac-
tors COMPETE and by National funds through FCT. The first author
acknowledges grant from the Portuguese Institute Camões and FNI
Soil properties of the samples collected in the surround- (Investigation National Fund—Mozambique).
ings of Hulene-B dump showed a strong alteration when
compared to background samples. This suggests a possible Data availability Data used is available on the manuscript.

13
542 Page 12 of 14 Environmental Earth Sciences (2022) 81:542

Declarations Chu Z, Zhou A, Ma Y, Zhuang J, Zhang L, Ma J (2020) Comparison


of municipal solid waste treatment capacity in China: a tourna-
Conflict of interest The authors declare that they have no conflict of ment graph method. J Mat Cycles Waste Manag 22(6):1913–1921.
interest. https://doi.org/10.1007/s10163-020-01077-4
Civita M, De Maio M (2004) Assessing and mapping groundwater
vulnerability to contamination: the Italian “combined” approach.
Geof Int 43(4):513–532
Dakheel AAJ, Bashir MJK, Llamas Borrajo JF (2022) Appraisal of
References groundwater contamination from surface spills of fluids asso-
ciated with hydraulic fracturing operations. Sci Total Environ
Afonso R (1978) A Geologia de Moçambique—Notícia Explicativa da 815:152949. https://doi.org/10.1016/j.scitotenv.2022.152949
Carta Geológica de Moçambique. Impr Nacional de Moçambique, Dregulo AM, Bobylev NG (2020) Heavy metals and arsenic soil
Maputo (in Portuguese) contamination resulting from wastewater sludge urban landfill
Alexakis DE (2021) Multielement contamination of land in the mar- disposal. Polish J Environ Stud 30(1):81–89. https://doi.org/10.
gin of highways. Land 10(3):1–13. https://doi.org/10.3390/land1 15244/pjoes/121989
0030230 El Fadili H, Ali MB, Touach N, El Mahi M, ElM L (2022) Ecotoxi-
Alleoni LRF, Iglesias CSM, Mello SDC, Camargo OA, Casagrande cological and pre-remedial risk assessment of heavy metals in
JC, Lavorenti NA (2005) Soil attributes related to cadmium and municipal solid wastes dumpsite impacted soil in morocco. Envi-
copper adsorption in tropical soils. Acta Sci Agron 27(4):729. ron Nanotec Monit Manag 17:100640. https://doi.org/10.1016/j.
https://doi.org/10.4025/actasciagron.v27i4.1348 enmm.2021.100640
Altaf R, Altaf S, Hussain M, Shah RU, Ullah R, Ullah MI, Datta R Fatoba JO, Eluwole AB, Sanuade OA, Hammed OS, Igboama WN,
(2021) Heavy metal accumulation by roadside vegetation and Amosun JO (2021) Geophysical and geochemical assessments of
implications for pollution control. PLoS ONE 16:1–15. https:// the environmental impact of Abule-Egba landfill, southwestern
doi.org/10.1371/journal.pone.0249147 Nigeria. Mod Earth Syst Environ 7(2):695–701. https://doi.org/
Andaloussi K, Achtak H, Nakhcha C, Haboubi K, Stitou M (2021) 10.1007/s40808-020-00991-8
Assessment of soil trace metal contamination of an uncontrolled Feng SJ, Wu SJ, Fu WD, Zheng QT, Zhang XL (2021) Slope stability
landfill and its vicinity: the case of the city of ‘Targuist’ (Northern analysis of a landfill subjected to leachate recirculation and aera-
Morocco). Moroccan J Chem 9(3):513–529. https://doi.org/10. tion considering bio-hydro coupled processes. Geoenviron Disast
48317/IMIST.PRSM/morjchem-v9i2.23680 8(1):1. https://doi.org/10.1186/s40677-021-00201-2
Awa SH, Hadibarata T (2020) Removal of heavy metals in contami- Ferrão DAG (2006) Evaluation of removal and disposal of solid waste
nated soil by phytoremediation mechanism: a review. Water Air in Maputo City, Mozambique. Master’s dissertation. University
Soil Pollut. https://doi.org/10.1007/s11270-020-4426-0 of Cape Town, Cape Town
Bernardo B, Candeias C, Rocha F (2022a) Application of geophysics in Gonçalves F, Correa CZ, Lopes DD, Rodolfo P, Vendrame S, Teixeira
geo-environmental diagnosis on the surroundings of the Hulene-B RS (2019) Monitoring of the process of waste landfill leachate dif-
waste dump, Maputo Mozambique. J Afr Earth Sci 185:104415. fusion in clay and sandy soil. Environ Monit Assess 191(577):1.
https://doi.org/10.1016/j.jafrearsci.2021.104415 https://doi.org/10.1007/s10661-019-7720-9
Bernardo B, Candeias C, Rocha F (2022b) Characterization of the Grisso R, Alley M, Holshouser D, Thomason W (2009) Precision farm-
dynamics of leachate contamination plumes in the surroundings ing tools: soil electrical conductivity. Virginia Coop Ext. https://
of the Hulene-B waste dump in Maputo, Mozambique. Environ. www.vtechworks.lib.vt.edu/bitstream/handle/10919/51377/442-
https://doi.org/10.3390/environments9020019 508.pdf?sequence=1&isAllowed=y
Bernardo B, Candeias C, Rocha F (2022c) Integration of electrical Helene LPI, Moreira CA (2021) Analysis of leachate generation
resistivity and modified DRASTIC model to assess groundwater dynamics in a closed municipal solid waste landfill by means
vulnerability in the surrounding area of Hulene-B waste dump, of geophysical data (DC resistivity and self-potential methods).
Maputo Mozambique. Water 14:1746. https://doi.org/10.3390/ Pure Appl Geoph 178(4):1355–1367. https:// doi. org/ 10. 1007/
w14111746 s00024-021-02700-7
Campos C (2010) Soil attributes and risk of leaching of heavy metals Hou D, Connor DO, Nathanail P, Tian L, Ma Y (2017) Integrated GIS
in tropical soils. Ambiência 6(3):547–565 and multivariate statistical analysis for regional scale assessment
Chaaou A, Chikhaoui M, Naimi M, Miad A, Achemrk A, Seif-Ennasr of heavy metal soil contamination : A critical review. Environ
M, Harche S (2022) Mapping soil salinity risk using the approach Pollution 231:1188–1200. https://doi.org/10.1016/j.envpol.2017.
of soil salinity index and land cover: a case study from Tadla 07.021
plain Morocco. Arab J Geosci 15:722. https://doi.org/10.1007/ Huang B, Yuan Z, Li D, Zheng M, Nie X, Liao Y (2020) Effects of soil
s12517-022-10009-5 particle size on the adsorption, distribution, and migration behav-
Chaudhary R, Nain P, Kumar A (2021) Temporal variation of lea- iors of heavy metal(loid)s in soil: a review. Environ Sci Proc Imp
chate pollution index of Indian landfill sites and associated human 22(8):1596–1615. https://doi.org/10.1039/d0em00189a
health risk. Environ Sci Pollution Res 28(22):28391–28406. Hussein M, Yoneda K, Mohd-Zaki Z, Amir A, Othman N (2021) Heavy
https://doi.org/10.1007/s11356-021-12383-1 metals in leachate, impacted soils and natural soils of different
Chen H, Wang L, Hu B, Xu J, Liu X (2022) Potential driving forces and landfills in Malaysia: an alarming threat. Chemosphere. https://
probabilistic health risks of heavy metal accumulation in the soils doi.org/10.1016/j.chemosphere.2020.128874
from an e-waste area, southeast China. Chemosphere 289:133182. INE (2020) Boletim de Estatísticas Demográficas e Sociais, Maputo
https://doi.org/10.1016/j.chemosphere.2021.133182 Cidade 2019. Instituto Nacional de Estatistica. http://www.ine.
Choppala G, Kunhikrishnan A, Seshadri B, Hee J, Bush R, Bolan N gov.mz/estatisticas/estatisticas-demograficas-e-indicadores-socia
(2018) Comparative sorption of chromium species as influenced is/boletim-de-indicadores-demograficos-22-de-julho-de-2020.pdf/
by pH, surface charge and organic matter content in contaminated at_download/file
soils. J Geochem Expl 184:255–260. https://doi.org/10.1016/j. Kapelewska J, Kotowska U, Karpińska J, Astel A, Zieliński P, Suchta
gexplo.2016.07.012 J, Algrzym K (2019) Water pollution indicators and chemomet-
ric expertise for the assessment of the impact of municipal solid

13
Environmental Earth Sciences (2022) 81:542 Page 13 of 14 542

waste landfills on groundwater located in their area. Chem Eng J of urban municipal landfill leachate. Environ Pollut 220:1–12.
359:790–800. https://doi.org/10.1016/j.cej.2018.11.137 https://doi.org/10.1016/j.envpol.2016.09.002
Kennou B, El Meray M, Romane A, Arjouni Y (2015) Assessment Nisari AR, Sujatha CH (2021) Assessment of trace metal contamina-
of heavy metal availability (Pb, Cu, Cr, Cd, Zn) and speciation tion in the Kol wetland, a Ramsar site, Southwest coast of India.
in contaminated soils and sediment of discharge by sequential Regional Stud Marine Sci 47:101953. https://doi.org/10.1016/j.
extraction. Environ Earth Sci 74(7):5849–5858. https://doi.org/ rsma.2021.101953
10.1007/s12665-015-4609-y O’Riordan R, Davies J, Stevens C, Quinton JN, Boyko C (2021)
Khomiakov DM (2020) Soil is an essential component of the biosphere The ecosystem services of urban soils: a review. Geoderma
and the global food system (critical assessment of the situation). 395:115076. https://doi.org/10.1016/j.geoderma.2021.115076
Moscow Univ Soil Sci Bull 75(4–5):147–158. https://doi.org/10. Odom F, Gikunoo E, Arthur EK, Agyemang F, Mensah-Darkwa K
3103/s0147687420040055 (2021) Stabilization of heavy metals in soil and leachate at Dom-
Kosheleva NE, Vlasov DV, Korlyakov ID, Kasimov NS (2018) Con- poase landfill site in Ghana. Environ Challenges 5:100308. https://
tamination of urban soils with heavy metals in Moscow as affected doi.org/10.1016/j.envc.2021.100308
by building development. Sci Total Environ 636:854–863. https:// Palalane J, Segala IO (2008) Urbanização e desenvolvimento municipal
doi.org/10.1016/j.scitotenv.2018.04.308 em Moçambique: gestão de resíduos sólidos. https://www.limpe
Lee H, Coulon F, Wagland ST (2022) Influence of pH, depth and humic zapublica.com.br/urbanizacao-e-desenvolvimento-municipal-em-
acid on metal and metalloids recovery from municipal solid waste mocambique-capitulo-gestao-de-residuos-solidos/
landfills. Sci Total Environ 806:150332. https://doi.org/10.1016/j. Parvin F, Tareq SM (2021) Impact of landfill leachate contamination on
scitotenv.2021.150332 surface and groundwater of Bangladesh: a systematic review and
Liu F, Yi S, Ma H, Huang J, Tang Y, Qin J, Zhou W (2019) Risk possible public health risks assessment. Appl Water Sci 11(6):1.
assessment of groundwater environmental contamination: a case https://doi.org/10.1007/s13201-021-01431-3
study of a karst site for the construction of a fossil power plant. Przydatek G, Kanownik W (2021) Physicochemical indicators of the
Environ Sci Pollut Res 26(30):30561–30574. https://doi.org/10. influence of a lined municipal landfill on groundwater quality: a
1007/s11356-017-1036-5 case study from Poland. Environ Earth Sci 80(13):1–14. https://
Lothe AG, Sinha A (2017) Development of model for prediction of doi.org/10.1007/s12665-021-09743-y
Leachate Pollution Index (LPI) in absence of leachate parameters. Rapti D, Masi S, Sdao F (2021) SIVRAD: an integrated system for
Waste Manag 63:327–336 the assessment of the environmental risk from solid waste land-
Lund ED (2008) Soil electrical conductivity. In: Logsdon S, Clay D, fills—guidelines. Acque Sotterranee Ital J Groundw 10(2):49–62.
Moore D (eds) Soil science: step-by-step field analysis. Ameri- https://doi.org/10.7343/as-2021-507
can Society of Agronomy and Soil Science Society of America, Rapti-Caputo D, Sdao F, Masi S (2006) Pollution risk assessment based
Madison on hydrogeological data and management of solid waste landfills.
Makuleke P, Ngole-jeme VM (2020) Soil heavy metal distribution Eng Geol 85:122–131. https://doi.org/10.1016/j.enggeo.2005.09.
with depth around a closed landfill and their uptake by Datura 033
stramonium. Appl Environ Soil Sci. https://doi.org/10.1155/2020/ Reeuwijk L (2002) Procedures for soil analysis, 6th edn. International
8872475 Soil Reference and Information Centre, Wageningen
Mancini IM, Sdao F, Masi S, D'Ecclesiis G (1999) Hydrogeological Saentho A, Wisawapipat W, Lawongsa P, Aramrak S, Prakongkep
pollution risk from solid waste: defining landfill siting criteria N, Klysubun W, Christl I (2022) Speciation and pH- and par-
and reclamation priority. Proceeding of 7° International landfill ticle size-dependent solubility of phosphorus in tropical sandy
symposium, 4–8 October, Cagliari, Italy, pp 579–588 soils. Geoderma 408:115590. https:// doi. org/ 10. 1016/j. geode
Mama CN, Nnaji CC, Nnam JP, Opata OC (2021) Environmental bur- rma.2021.115590
den of unprocessed solid waste handling in Enugu State, Nigeria. Salam A, Bashir S, Khan I, Hussain Q, Gao R, Hu H (2019) Biochar
Environ Sci Pollut Res 28(15):19439–19457. https://doi.org/10. induced Pb and Cu immobilization, phytoavailability attenua-
1007/s11356-020-12265-y tion in Chinese cabbage, and improved biochemical properties
Momade FJ, Ferrara M, Oliveira JT (1996) Notícia explicativa da in naturally co-contaminated soil. J Soils Sed 19:2381–2392.
carta geológica 2532 Maputo (Escala 1:50 000). Maputo (in https://doi.org/10.1007/s11368-019-02250-5 (SOILS)
Portuguese) Sarmento L, Tokai A, Hanashima A (2015) Analyzing the structure
Morita AKM, Pelinson NS, Wendland E (2020) Persistent impacts of barriers to municipal solid waste management policy plan-
of an abandoned non-sanitary landfill in its surroundings. ning in Maputo city, Mozambique. Environ Develop 16:76–89.
Environ Monit Assess 192(7):1. https:// doi. org/ 10. 1007/ https://doi.org/10.1016/j.envdev.2015.07.002
s10661-020-08451-7 Seidl M, Le Roux J, Mazerolles R, Bousserrhine N (2021) Assess-
Muchangos A (1999) Paisagens e Regiões Naturais de Moçambique, ment of leaching risk of trace metals, PAHs and PCBs from a
pp 5–163. https://docplayer.com.br/47220681-Mocambique-paisa brownfield located in a flooding zone. Environ Sci Pollut Res.
gens-e-regioes-naturais.html. Accessed 6 Mar 2022 https://doi.org/10.1007/s11356-021-15491-0
Munsell Color (2009) Munsell soil color book; color charts; munsell Serra C (2012) Da problemática Ambiental à mudança: rumo à um
colour company. Inc.: Newburgh, NY, USA, 2009 mundo melhor. Escolar Editora, Forte da Casa
Nadiri AA, Sedghi Z, Khatibi R, Gharekhani M (2017) Mapping vul- Škrbić B, Đurišić-Mladenović N (2010) Chemometric interpreta-
nerability of multiple aquifers using multiple models and fuzzy tion of heavy metal patterns in soils worldwide. Chemosphere
logic to objectively derive model structures. Sci Total Environ 80(11):1360–1369. https://doi.org/10.1016/j.chemosphere.2010.
593–594:75–90. https://doi.org/10.1016/j.scitotenv.2017.03.109 06.010
Naveed M, Bukhari SS, Mustafa A, Ditta A, Alamri S, El-Esawi MA Soubra G, Massoud MA, Alameddine I, Al Hindi M, Sukhn C (2021)
et al (2020) Mitigation of nickel toxicity and growth promotion in Assessing the environmental risk and pollution status of soil
sesame through the application of a bacterial endophyte and zeo- and water resources in the vicinity of municipal solid waste
lite in nickel contaminated soil. Int J Environ Res Public Health dumpsites. Environ Monit Assess 193(12):1. https://doi.org/10.
17(23):1–27. https://doi.org/10.3390/ijerph17238859 1007/s10661-021-09640-8
Naveen BP, Mahapatra DM, Sitharam TG, Sivapullaiah PV, Ramachan- Sparling GP (2020) Soil quality indicators. Manag Soils Terrest Syst.
dra TV (2017) Physico-chemical and biological characterization https://doi.org/10.1201/9780429346255-44

13
542 Page 14 of 14 Environmental Earth Sciences (2022) 81:542

Stefania GA, Rotiroti M, Buerge IJ, Zanotti C, Nava V, Leoni B, Yap CK, Chew W, Al-mutairi KA, Nulit R, Ibrahim MH, Wong KW
Bonomi T (2019) Identification of groundwater pollution et al (2022) Assessments of the ecological and health risks of
sources in a landfill site using artificial sweeteners, multivari- potentially toxic metals in the topsoils of different land uses: a
ate analysis and transport modeling. Waste Manag 95:116–128. case study in Peninsular Malaysia. Biology. https://doi.org/10.
https://doi.org/10.1016/j.wasman.2019.06.010 3390/biology11010002
Tenodi S, Krčmar D, Agbaba J, Zrnić K, Radenović M, Ubavin D, Zamulina IV, Gorovtsov AV, Minkina TM, Mandzhieva SS, Burachevs-
Dalmacija B (2020) Assessment of the environmental impact of kaya MV, Bauer TV (2021) Soil organic matter and biological
sanitary and unsanitary parts of a municipal solid waste land- activity under long-term contamination with copper. Environ Geo-
fill. J Environ Manag. https://doi.org/10.1016/j.jenvman.2019. chy Health. https://doi.org/10.1007/s10653-021-01044-4
110019 Zeitoun R, Vandergeest M, Vasava HB, Machado PVF, Jordan S, Par-
USDA (2001) Rangeland soil quality—organic matter. United States kin G et al (2021) In-situ estimation of soil water retention curve
Department of Agriculture. http://www.ftw.nrcs.usda.gov/glti. in silt loam and loamy sand soils at different soil depths. Sensors
Accessed 27 April 2022 21(2):1–15. https://doi.org/10.3390/s21020447
USDA (2011). Soil quality indicators. United States Department of Zhang M, Reardon EJ (2003) Removal of B, Cr, Mo, and Se from
Agriculture. https://www.nrcs.usda.gov/wps/portal/nrcs/detail/ wastewater by incorporation into hydrocalumite and ettringite.
soils/health/assessment/?cid=stelprdb1237387. Accessed 2 Mar Environ Sci Techn 37(13):2947–2952. https://doi.org/10.1021/
2022 es020969i
USDA (2014). Soil health—electrical conductivity. United States Zhang J, Zhang J, Xing B, Liu G, Liang Y (2021) Study on the effect of
Department of Agriculture. https://www.nrcs.usda.gov/Inter net/ municipal solid landfills on groundwater by combining the models
FSE_ DOCUM ENTS/ nrcs1 42p2_ 052803. pdf. Accessed 3 Feb of variable leakage rate, leachate concentration, and contaminant
2022 solute transport. J Environ Manag 292:112815. https://doi.org/10.
Vicente EM, Jermy CA, Schreiner HD (2006) Urban geology of 1016/j.jenvman.2021.112815
Maputo, Mocambique. Geol Soc London 338:1–13 Zimik HV, Farooq SH, Prusty P (2021) Source characterization of trace
Victor KK, Adjiri AO, Clement KK, Emilie KK, Honore KC (2019) elements and assessment of heavy metal contamination in the soil
Physical characterization of the superficial layers of Akouedo around Tarabalo geothermal field, Odisha, India. Arabian J Geosc
Landfill, Ivory Coast and assessment of heavy metals pollution 14(11):1. https://doi.org/10.1007/s12517-021-07366-y
risk of the underlying aquifer. J Environ Sci Pollut Res 5(3):361–
363. https://doi.org/10.30799/jespr.173.19050303 Publisher's Note Springer Nature remains neutral with regard to
Wijekoon P, Koliyabandara PA, Cooray AT, Lam SS, Athapattu BCL, jurisdictional claims in published maps and institutional affiliations.
Vithanage M (2022) Progress and prospects in mitigation of land-
fill leachate pollution: risk, pollution potential, treatment and chal- Springer Nature or its licensor (e.g. a society or other partner) holds
lenges. J Hazard Mat 421:126627. https://doi.org/10.1016/j.jhazm exclusive rights to this article under a publishing agreement with the
at.2021.126627 author(s) or other rightsholder(s); author self-archiving of the accepted
Wu L, Zhan L, Lan J, Chen Y, Zhang S, Li J, Liao G (2021) Leachate manuscript version of this article is solely governed by the terms of
migration investigation at an unlined landfill located in granite such publishing agreement and applicable law.
region using borehole groundwater TDS profiles. Engin Geol
292:106259. https://doi.org/10.1016/j.enggeo.2021.106259

13
271

13.2.4 Appendix 5. Paper 4


272
Manuscript Click here to
access/download;Manuscript;Manuscript_EnvPoll.docx
Click here to view linked References

1 Soil properties temporal variation and risk assessment in the surroundings of Hulene-B dump,
2 Maputo (Mozambique)

3 Bernardino Bernardoa,b, Carla Candeiasb Fernando Rochab

a
4 GeoBioTec Research Centre, Department of Geosciences, University of Aveiro, 3810-193, Aveiro,
5 Portugal

b
6 Faculty of Earth Sciences and Environment, Pedagogic University of Maputo, Mozambique

7 Abstract

8 The present study considers the temporal variation of soil properties in surroundings of Hulene-B waste
9 dump (Maputo, Mozambique), between a rainy season (2020) and a dry season (2021). For that purpose
10 71 superficial soil samples were collected in the two climatic seasons. Soil properties analyzed were
11 texture, pH, electrical conductivity (EC), organic matter (OM), color, and moisture. Dry season samples
12 results suggested higher tendency of potentially toxic elements accumulation in soil suface, such as high
13 EC values. The surfacewater contamination risk index (Pbci) was estimated, taking in consideration the
14 characteristics of the dump and its surroundings, revealing high values, indicative of contamination by
15 leachates. The study highlighted the need for mitigation measures for soil and surface water
16 contamination, adjusted to the geoenvironmental dynamics of Hulene-B dump surroundings, taking in
17 consideration the fragilities of the environmental components and the main vectors of contamination
18 (leachates).

19 Keywords: soil properties, waste dump, temporal variation; surface water, risk assessment

20 1. Introduction

21 The lack of planned urban solid waste disposal infrastructures, including dumping sites, is a major cause
22 of environmental pollution in many cities of underdeveloped countries (Morita et al., 2021). Audu et al.
23 (2021) and Chaudhary et al. (2021) suggested that areas surrounding waste dumps presented high levels
24 of contamination represent a risk to human health. Temporal variation and dispersion of leachtes amount
25 and composition, is of significat importance to understand the contamination processes around dumps
26 (Alghamdi et al., 2021; Zaki et al., 2022). Nilam et al. (2016) and Wijekoon et al. (2022) showed that
27 undeveloped dumps produce more leachates, especially during hot and rainy season, due to greater
28 decomposition of solid waste, leading to dispersion and contamination of soil and surface and ground
29 waters. Metal ions content in leachates, vary according to the type of waste and environmental conditions
30 (Nilam et al., 2016), being higher in scarce water conditions (Yan et al., 2021), once water is responsible
31 for dilution (Siddiqi et al., 2022).

32 The dynamic of environmental contamination around landfills is strongly influenced by soil characteristics
33 (Ihedioha et al., 2017). Soil is considered as the first environmental component to be impacted by
34 contamination caused by leachates (Xiao et al., 2021), since soils act as a filter for contaminants and
35 control chemical elements transport to the hydrosphere, atmosphere and biota (Hasan et al., 2021). The
36 hability of soils to control the circulation of potentially toxic elements (PTEs), such as metal ions, in
37 leachates depends not only on contaminants content, but also on the soil physical and chemical properties
38 that varies temporally (Campos, 2010; Chaudhary et al., 2021). Rieuwerts et al. (2006) revealed the need
39 of soil properties temporal variation information to understand attenuation, migration, adsorption and
40 absorption of PTEs by the soil matrix, clarifying the interference of attributes in the behaviour of heavy
41 metals and the reactions triggered. Relevant properties for assessing contamination processes include
42 pH, organic matter (OM), moisture, colour, soil texture and electrical conductivity (EC) (Fatoba et al.,
43 2021). A pH > 6 favour the dissociation of H+ from OH groups in organic matter and Fe and Al oxides,
44 increasing the adsorption of heavy metals and subsequent precipitation, reducing PTEs bioavailability
45 (Pagnanelli et al., 2003). Organic matter present great affinity with heavy metals, increasing its absorption
46 capacity by the soil (Ghobadi et al., 2021). Soil moisture influences metal ions dissolution contained in
47 leachates (Kumar et al., 2013), while soil texture plays a key role in the fixation and migration of heavy
48 metals by leaching processes, with sandy soils being prone to PTEs leaching (Dakheel et al., 2022).
49 Eelectrical conductivity in soils surrounding landfills has been used as an indicator of soil contamination
50 by leached metal ions (Hussein et al., 2021). Climatic conditions, with scarce or abundant precipitation,
51 are important factors for soils physical parameters, being pediods of intense precipitation in areas
52 contaminated by heavy metals, prone to leaching, thus spreading the contamination and reducing
53 concentration levels (Sun et al., 2022), while during dry season PTEs, such as Pb, Cu, Hg, tend to
54 concentrate on soil surface (Shaylinda, 2020).

55 Several methods have been developed to estimate environmental contamination risk in areas surrounding
56 waste dumps. Rapti-Caputo (2006) developed a method for pollution risk assessment based on
57 hydrogeological data and management of solid waste landfills. Calvo et al. (2005) proposed an
58 environmental diagnostic methodology for municipal waste landfills consisting in formulating a series of
59 environmental indices that provide information on the potential environmental problems of landfills and
60 the particular impact on the different environmental elements, as well as information related to location,
61 design and operation. However, it is necessary to assess in detail the soil properties to understand their
62 influence on the attenuation, migration, adsorption and absorption mechanisms of dumps contaminants.

63 In Mozambique, urban areas do not have effective solid waste disposal mechanisms, which often results
64 in waste accumulation in random areas (Serra, 2012). In Maputo, the capital city of Mozambique, waste
65 disposal occurs mostlly in Hulene-B waste dump, the largest dump of the country, receiving > 1000 tons
66 of different rejected materials daily (CMCM, 2020). Vicente et al. (2006) and Serra (2012) described the
67 dump as one of the soil and ground and surface water contamination causes. In the dump surroundings,
68 previous studies by Vicente (2006) and Bernardo et al. (2022a,b,c,d,e), revealed a strong influence of the
69 dump on environmental contamination. Bernardo et al. (2022c) compared the properties of soils collected
70 during rainy season in an area surrounding of Hulene-B landfill and areas not affected (background),
71 suggesting a significant difference associated to the dump contamination. Chemical analyses of the same
72 soils showed PTEs enrichment, such as Zn, Cu , Cr , Zr , Pb , Ni , Mn (Bernardo et al. 2022d).

73 This study aims to identify the seasonal variation of soil properties in Hulene-B dump surroundings and
74 estimate surface water contamination index. The monitoring of soil properties is pointed out as relevant
75 for understanding seasonal variation of contaminant uptake, adsorption and migration processes
76 (Chaudhary et al., 2021), allowing to identify mitigation measures to control the contamination processes
77 (Hoai et al., 2021; Ozbay et al., 2021).

78 2. Metodology

79 2.1 Study area

80 The Hulene-B waste dump, in Maputo, is one of the largest in Mozambique (Serra, 2012) (Fig. 1). It is
81 surrounded by a residential area, the Hulene-B neighborhood, with ~ 49 000 inhabitants (INE, 2020). The
82 dump is located in an abandoned quarry with no preparation for waste deposition, and with a waste
83 deposit that grew rapidly westwards (Ferrão, 2006). Wastes are also spread on heterogeneous sandy soils
84 resulting from the mixing of soils from Ponta Vermelha Formation and Malhazine (Bernardo et al., 2022a).
85 The dump receives all types of waste produced in Maputo urban area (Sarmento et al., 2015). CMCM
86 (2020) estimated that wastes deposited in Hulene-B dump contain 68 % organic matter, 13.2 % paper, 9
87 % plastics, 7.5 % glass, 2.8 % metals, and 1.5 % rubber. The height of the wastes was estimated to be ~6
88 to 15 m, occupying an area of ∼17 h. Dump surrounding soils (W border), have been used for subsistence
89 agriculture (horticulture), due to the continuous accumulation of surface water along the depression,
90 keeping soils moist.
91 Geologically, the area surrounding Hulene-B dump is inserted in the Mesocenozoic sedimentary basin of
92 southern Mozambique, in a contact zone of two lithologies, the Ponta Vermelha Formation and Malhazine
93 Formation, on a gentle dune slope with E-W orientation (Vicente, 2011; Cendón et al., 2020; Momade et
94 al., 1996). Hulene-B area hydrogeological system is part of the Tertiary - Quaternary aquifer system
95 (Cendón et al., 2020), with its substrate formed by a layer of clayey marl to grey clay (Cendón et al., 2020;
96 Muchimbane, 2010). It was identified the presence of a semi- impermeable layer (clayey sands), leading
97 to water circulation between these two sectors (Muchimbane, 2010; Bernardo et al., 2022a). There are
98 places where coarse sands lie directly on top of the clay layer, developing semi-confined conditions
99 (Nogueira et al., 2019). The water level of shallow wells varies between 1.5 and 9.3 m deep, with an
100 average of 3.8 m (Cendón et al., 2020). Bernardo et al. (2022a) suggested that groundwater in the
101 surroundings of the dump presents heterogeneous depths, with the eastern boundary varying between 3
102 and 6 m depth. The western edge of the dump, given the morphological configuration (intradune
103 depression) has permanently surface water and on the north a temporary flooding mostly during rainy
104 season (Bernardo et al. 2022b).

105 The predominant climate is of subtropical type, with mean annual precipitation of ~789 mm , with two
106 climatic seasons: (a) hot (mean 25 ºC) and rainy period from December to March with > 60 % of the annual
107 precipitation, with its peak in January (~125 mm), and (b) dry and cold season, from April to September,
108 with lower temperatures in June and July (mean 21 ºC), and scarce precipitation, whose minimum values
109 were recorded in August (~12 mm) (CIAT, 2017). The prevailing winds are SE (Muchangos, 1999).

110 2.1.1 Sampling

111 A total of 71 soil samples were collected at 0-20 cm depth in the surrounding area of Hulene-B dump, in
112 January 2020 and May 2021 (Fig. 1). The 2020 samples were collected during heavy rainfall, with a monthly
113 average of 123 mm, and 2021 samples were collected during low rainfall, with a monthly average of 25
114 mm (CIAT, 2019). For background, 10 samples were collected in areas considered not impacted by the
115 dump. Samples were georeferenced and preserved in plastic bags until laboratorial treatment. On the
116 laboratory, samples were oven dried ~40 ºC at Pedagogical University of Maputo (Mozambique).
117 Afterwards, samples were transported to the laboratory of GeoBioTec Research Center, University of
118 Aveiro (Portugal), for analyses. Soil samples were sieved to achieve the < 2000 µm (sand) and < 63 (silt)
119 µm fractions. The pH was determined in the two fractions with a 1:2.5 soil/water solution using a pH
120 meter. Electrical conductivity (EC) was measured under the same conditions in the two fractions, using an
121 high resolution conductivimeter. Organic matter (OM), and moisture contents wer determined by the
122 method described in Reeuwijk (2002). Texture was assessed using a Sedigraph III Plus grain size analyzer,
123 that determined the relative mass distribution of a sample by particle size and is based on two physical
124 principles: sedimentation theory (Stokes' law) and the absorption of X-radiation (Beer-Lambert law).

125 Organic matter and pH classes were established as defined by Sparling, (2020), and USDA (2001a)
126 respectively, with the following classes: OM > 4 % - high content, OM 2 to 4 % - medium content, and OM
127 < 2 % - low content; and pH 3.5 to 4.4 - extremely acid, pH 4.5 to 5.0 - very strongly acid, pH 5.1 to 5.5 -
128 strongly acid, pH 5.6 to 6.0 - moderately acid, pH 6.1 to 6.5 - slightly acid, pH 6.6 to 7.3 - neutral, pH 7.4
129 to 7.8 - slightly alkaline, pH 7.9 to 8.4 - moderately alkaline, and pH 8.5 to 9.0 - strongly alkaline. Electrical
130 conductivity classes were defined based on internationally adopted reference values for sandy soils (Lund,
131 2008; USDA, 2011b), being classified as EC ≤ 100 μS/cm - natural medium, 101 to 200 μS/cm - high, and
132 EC > 201 μS/cm - very high. Soil color was determined using Munsell (2009) soil chart, with classes defined
133 taking in consideration color intensity, from least to more intense: class 1 - grayish, class 2 – brownish;
134 and class 3 - blackish. Moisture classes were defined according to higher local background value (2 %), as
135 class I - < 0.41 %; class ii - 0.41 to 0.82 %; and class iii > 0.82 %.

136 2.1.2 Statistical analysis

137 Descriptive statistics, principal component analysis (PCA) and Spearman correlation were performed using
138 SPSS® v.25 software (IBM, USA). PCA is a data dimensionality reduction technique that aims to explain
139 most of data variation with a small number of independent variables. called principal components (Hou
140 et al., 2017). In this way, Promax rotation was done which allowed the variance of the analyzed
141 parameters to be correlated with each other. Spearmans correlation was used to assess the strength of
142 pairwise correlation of the analyzed soil properties. The combination PCA and Spearman's clustering was
143 relevant for analyzing the sources of the drivers of soil properties change as well as understanding of the
144 spatial variation. The data were also processed by the Geostatistic Analyst tool for ArcMap (ArcGIS). In
145 this way, spatial projection classes of the soil properties were defined, which allowed the analysis of the
146 patterns of their spatial distribution.

147 2.2 Surface water contamination risk by landfill

148 The contamination surface water risk (Pbci) by landfills are parameters used in landfill contamination
149 environmental assessment (Calvo et al., 2005). To evaluate the contamination probability, landfill
150 variables were selected taking in consideration the sensitivity to biochemical and physical processes that,
151 directly or indirectly, influence the surrounding environment (Aryampa et al., 2021). In this study, 8 factors
152 were selected with a link to surface water: (1) degree of compaction, (2) final cover, (3) leachates control,
153 (4) type of waste and % organic matter, (5) precipitation, (6) surface drainage systems, (6) waterproofing
154 discharge point, (7) distance to surface water, and (8) leachate control. These variables evaluation allow
155 to assess contamination risk in landfills through the probability of contamination for each contamination
C𝑗 ×W𝑗
156 variable (Pbcj), fiven by 𝑃𝑏𝑐𝑗 = , where Wj is the weighting of variable j, Cj the classification of
N

157 variable j, dependant on the variable state and provides information on the status of the interaction
158 between disposal processes and environmental characteristics related to the variable, and N the number
159 of variables of each parameter (Calvo et al., 2005). Weight values are assigned according to the
160 characteristics of the factor evaluated in relation to the risk of contamination (Tables S1-S8). Risk
161 classifications correspond to Pbci = 0 nonexistent; 0 < Pbci < 0.3 low; 0.3 < Pbci < 0.6 mean; 0.6 < Pbci ≤ 1
162 high (Calvo et al., 2005).

163 3. Results and Discussion

164 Textural distribution of the studied samples and background is presented in Figure 2. Background samples
165 and most of the soil samples collected in the surroundings of Hulene-B were classified as sand, with the
166 exception of samples P1II, P2II, P7II, P11II, P4II, P21II, classified as loamy sand, due to their silt content.
167 Samples with higher clay content were ranked P1II > P67II > P2II > P11II > P21II > P66II > P55II, ranging 5.4
168 to 2.6 %, samples located at the W and E boundaries of the dump. Sand content, dominant in all soil
169 samples ranged 97.9 (P25II) to 81.9 (P1II) %, being higher sand content found in samples collected on the
170 W boundary. Background soils sand content varied 95.4 to 92.1 %, silt between 3.8 to 2.3 %, and clay
171 between 3.8 and 2.3 % (Table 1). These data showed significant differences from a study conducted by
172 Bernardo et al. (2022c) at same locations during rainy season, that classified all soil samples around the
173 Hulene-B dump as sand. Silty and clayey fractions difference between dry (mean content of 4.49% and
174 1.56 %, respectively) and rainy (mean content of2.1% and 0.22%, respectively) seasons (Bernardo et al.
175 2022c), is possibly linked to precipitation absence during dry season, a relevant factor in the leaching of
176 fine particles and ash from waste incineration that accumulate on the soil surface (Dakheel et al., 2022).
177 Clay and silt fractions can act as an important factor for the adsorption of heavy metals in soils (Shaylinda,
178 2020). Thus, soils textural characteristics during dry season showed a higher metal adsorption capacity in
179 the soils surrounding the Hulene-B dump. Background samples showed no significant differences between
180 the two sampling campaigns (Bernardo et al., 2022c).

181 In general, background samples did not revealed a significant properties variation (p > 0.05) between the
182 two sampling seasons, suggesting reduced impact factors that could significantly modified the analyzed
183 parameters (Table 2). Samples collected in 2021 showed slightly higher pH, EC, and OM, than those of
184 2020. These differences can be associated to the fact that 2021 samples were collected during scarce
185 rainfall period, a relevant factor in the variation of these properties.

186 Hulene-B dump surrounding soils collected in 2020 showed, in sand fraction, a pH ranging 6.7 to 8.0, with
187 an average of 7.4, and in silt fraction ranging 7.0 to 8.4, with an average of 7.8, being predominantly
188 alkaline to slightly alkaline (Bernardo et al., 2022c). The same parameter in 2021, showed that sandy
189 fraction ranged 7.5 to 10.5, with mean 8.52, while and silt fraction varied between 7.6 and 8.9, and mean
190 7.93. Silt fraction revealed a predominantly slightly alkaline pH in most of the study area (Fig. 3). Soil pH
191 in the surroundings of the Hulene-B dump, has been attributed as favorable to migration of PTEs
192 (Rieuwerts et al., 2006). However, in tropical regions soil, heavy metal retention can occur under high pH
193 conditions (Campos, 2010), due to predominance of oxidic (Al, Fe, and Mn) and kaolinitic mineralogy in
194 the clay fraction, increasing metal adsorption capacity (Alleoni et al., 2005). Bernardo et al. (2022d)
195 suggested a predominance of Al, Fe and Mn oxides and a slight predominance of clay minerals in Hulene-
196 B dump surrounding soils, an important factor in the PTEs adsorption and absorption even under mildly
197 alkaline to moderate conditions. In a similar study by Naveen et al. (2017) near Mavallipura (India),
198 contaminated soil alkalinity was detected and linked to successive accumulation of ash from waste
199 incineration and circulation of leachates from waste biodegradation, potential factors in contaminating
200 and altering soil pH. Škrbić et al. (2004), suggested a biggest heavy metal retention in soil with high pH,
201 given a lower metals solubility.

202 Sandy soil EC of the 2020 samples, varied 43.1 to 725 µS/cm, with an average of 220.59 µS/cm, and silt
203 fraction ranged 37.5 to 217 µS/cm, and average 88.4 µS/cm (Bernardo et al., 2022c), with higher values
204 found in samples collected on the immediate limit of Hulene-B dump (Fig. 4). Samples collected in 2021
205 showed higher EC, ranging 90 to 9510 µS/cm, and average of 1150 µS/cm on sandy fraction, and ranging
206 85 to 1720 µS/cm and average 378 µS/cm on silt fraction. Spatially, in 2021, the sandy fraction had higher
207 EC values throughout the area around the dump, with higher values detected in few samples taken in the
208 northern section of the sampling area, and in one sample in the SW section of the dump. Relatively low
209 conductivities were detected in one sample located at the north. Silt fraction higher EC was detected in
210 samples located on the W of the dump while lower values in one sample on the N of the sampling area.
211 Samples collected in 2021 showed higher EC values than 2020, probably related to the different climatic
212 conditions of the sampling periods. Precipitation is a factor in leaching and migration of contaminants,
213 influencing soils EC (Li et al., 2022). Results suggested a strong soil contamination during dry period
214 (Siddiqi et al., 2022). Ashraf et al. (2022) compared the contamination levels of soils in the surrounding of
215 a landfill in Lahore, having similar conclusions. Bernardo et al. (2022a) suggested in an electrical resistivity
216 comparative study between 2020 and 2021, that anomalies during dry period were located at greater
217 depth, when compared to rainy period, and associated this phenomenon to the poor superficial leaching
218 in dry period. Other studies showed that during dry periods, there was higher accumulation of heavy
219 metals in the topsoil layer (Azeez et al., 2011; Chaudhary et al., 2021). Wu et al. (2022) revealed that areas
220 with contamination sources, e.g., waste dumps, during dry period presented higher transport and
221 deposition of contaminants to soil surface, related to higher EC in soils around dumps (Naveen et al.,
222 2017).

223 Organic matter content on 2020 samples ranged 0.0 to 4.2 %, and average 1.09% (Bernardo et al., 2022c).
224 Samples collected in 2021, ranged 1.8 to 8.6 %, and mean 1.87 %, with higher OM content than 2020
225 samples (Fig. 5). Spatially, OM content > 4 % and between 2 to 4 % were found in samples located in the
226 immediate E and W edges of the dumpsite, a similar pattern to 2020 study. Samples collected on the W
227 occupied a larger area than 2020 samples, while lower OM content was found in the center and N areas
228 of the sampling site, also with similar pattern to 2020 samples (Bernardo et al. 2022c). Temporally, 2020
229 samples showed lower OM content when compared to 2021. Studies suggested that OM in sandy soils is
230 susceptible of leaching (Siddiqi et al., 2022), explaining the relatively higher predominance of OM content
231 in 2021 (scarce rain season). However, it is thought that high OM content in topsoil is a factor that may
232 lead to the uptake of contaminants, once heavy metals bind easily to OM (USDA, 2001a) and consequently
233 decrease toxicity levels (Nisari et al., 2021).

234 Soil color, in 2020 samples, was characterized by blackish colors at the immediate limit of the dump in a
235 very narrow band, and slightly dark in the transition area between the dump limit and the W, and at the
236 center of the sampling area with lighter color (Fig. 5) (Bernardo et al. 2022c). The 2021 collected samples
237 color distribution, presented darker soil samples found at the immediate limit of the dump, occupying a
238 larger band previous study. Soil brownish color was found on a wide band near the darkened samples and
239 at NW. Greyish presented a similar pattern to the 2020 samples (Bernardo et al. 2022c). In general, areas
240 with dark and brownish soil colors, with contamination sources such as dumps, are associated with a
241 greater capacity of the soil to absorb PTES (Shaylinda et al., 2020; Nisari et al., 2021). Greyish soil color, in
242 areas with contamination sources such as leachates have been associated to leaching processes,
243 consistent with the present study, since more grayish soil samples were found in 2020 sampling, during
244 dry season. In the surrounding of dumpsites, color is also influenced by ash deposition resulting from solid
245 waste incineration, which is a vector of contamination mainly by Hg and Pb (Tao et al., 2020). Weibel et
246 al. (2017) suggested that waste burning contributed to the dispersion of Hg-contaminated ash, which was
247 subsequently deposited in the topsoil causing its enrichment and alteration of the soil color. In Hulene-B
248 dump the practice of waste burning is characteristic and may be associated with the blackish and brownish
249 colors of the predominant 2021 soils samples.

250 Samples collected in 2020 in the surrounding of the landfill moisture content ranged 0.0 to 1.9 %, and
251 average 0.4% (Fig. 5) (Bernardo et al. 2022c). Soil samples collected in 2021, showed higher moisture
252 content, ranging 0.18 to 8.7 %, and average 0.94 %. Highest moisture content was recorded at the center
253 to W of the study area, and slightly elevated content at the immediate edge of the dump. Temporally,
254 background levels were slightly different, with higher moisture content in 2021 samples. In samples
255 collected on the surroundings of Hulene-B dump, in general, was found natural (low) moisture content,
256 characteristic of sandy soils, being an indicator of strong leaching and contaminants migration, especially
257 in soils surrounding dumpsites, that may be affected by the leachates (Zeitoun et al., 2021). Möller et al.
258 (2005), and Hussein et al. (2021) suggested that low soil moisture with active sources of contamination
259 (urban areas, waste dumps) lead to increased accumulation of contaminants in topsoil. This was
260 evidenced by dry season results (Fig. 4). Samples EC of 2020 (220.59 µS/cm) and 2021 (average of 1150
261 µS/cm) suggested conditions of higher contaminants accumulation than in rainy season, while pH,
262 moisture, color, OM showed no significant changes.

263 Principal Component Analysis (PCA)

264 Principal components were extracted from 71 samples and 7 variables (sand, silt, clay, OM, EC, pH, and
265 moisture), with Promax rotation. Two principal components were considered, accounting for 73.85 % of
266 the total variance. First component (PC1), explained 56.53 % of total variance, defined two groups of
267 variables: sand negatively related to silt, clay, OM, EC, pH, and moisture (Fig. 6), with high negative loading
268 (> 0.972). OM contents were predominantly low, being more prevalent in samples close to the dumpsite,
269 associated to the incorporation of organic matter from leachates and addition of ashes from the
270 incineration of solid wastes. EC was predominantly high in all samples evidencing higher values than the
271 world average for sandy soils (< 100 µS/cm) and the pH was predominantly alkaline in all samples. Second
272 component (PC2) explained 17.32 % of total variance, with positive loading composed by pH (0.885). This
273 PC can be explained by reduced samples pH variation, predominantly alkaline. The high soil samples pH
274 can be used as an indicator of soil contamination by heavy metals (krbić et al., 2004), in particular in
275 tropical areas with a higher predominance of Al and Fe oxides (Campos, 2010)
276 Spearman correlation is presented in Table 3. The pairs silt/clay, clay/sand, OM/sand, EC/sand (p < 0.01),
277 pH/sand, pH/clay, pH/OM, pH/EC, moisture/sand and moisture/pH (p < 0.05) revealed a significant
278 negative correlation, suggesting that properties silt, clay, sand, OM, EC and pH behave oposedly. The
279 opposition suggested different soil alteration processes in the surrounding of Hulene-B dump. The
280 opposition between clay and sand is possibly linked to increased leaching and consequent PTEs migration.
281 Studies reported that pH is the most important factor in controlling mechanisms of heavy metal
282 contamination, determining their mobility in depht in the areas around dumpsites (Naveen et al. 2017).
283 Despite low silt, clay, and OM content, high EC was observed which can be an evidence of contamination
284 by metal ions contained in leachates or solid waste incineration ashes. The pairs clay/silt, OM/silt, OM/silt,
285 EC/silt, EC/clay EC/OM, moisture/silt, moisture/clay, moisture/OM (p < 0.01) and moisture/EC (p < 0. 05)
286 showed a possible correlation suggesting that the properties clay, silt, OM, moisture are properties that
287 change in the same trend, which can be evidenced by the samples that had higher silt content showed
288 higher clay and OM content thus evidencing a higher adsorption and absorption capacity of heavy metals
289 in the vicinity of Hulene-B dump, especially in samples from the immediate limits of the dump which had
290 high levels of these contents. Samples relatively distant from the dump showed the same clay, silt, OM
291 trend, that may be associated with superficial and vertical leaching processes associated with the free
292 circulation of leachates and the contaminants migration to lower soil layers. In similar studies, Zhang et
293 al. (2010) and Pikula et al. (2022) suggested that reduced clay, silt, and OM in sandy soils promoted the
294 migration of contaminants in depth. OM, EC, and moisture were properties with higher alterability in the
295 vicinity of dumps, given the circulation of leachate and ash from waste incineration that can increase these
296 properties (OM, EC, moisture), wtih moisture that can be associated with the accumulation of surface
297 water enriched by leachate to the west of the Hulene-B dump.

298 Risk of surface water contamination (Pbci)

299 The weights (W) and ranking levels (C) were assigned to the studied variables, based on the characteristics
300 of the landfill and its surroundings (Calvo et al., 2005): (1) compaction, given the precariousness of the
301 compaction processes at Hulene-B dump, characterized by being partial, it was assigned W = 2, and C = 3;
302 (2) cover material, once is an open-air landfill without any mechanism to cover the waste, it was assigned
303 W = 2, and C = 4; (3) leachate control, at the N, S and W borders of the landfill, to where leachates drains
304 and been opened and cyclically overflowing, with no underground isolation being likely that contribute to
305 migration of leachate in depth (Fig. 7), with maximum W = 2 to surface water, and W = 1 to soil, and C =
306 4; (4) age and percentage of organic matter, the landfill dates from 1973 (~50 years age), assuming little
307 leachate generation but >1000 tons of miscellaneous waste deposited daily, and an estimated 50 % of
308 organic, W = 2 and C = 4; (5) impermeability of discharge point, the landfill has neither been waterproofed
309 or its surroundings, increasing the potential risk of soil, ground and surface water contamination, W = 1
310 for soils, W = 2 for surface water, and C = 4; (6) being located in flood-prone areas, the surroundings of
311 the dump is susceptible of inundation, given the inexistence of mechanisms to adequately control surface
312 flows leading to widespread flooding by surface leachate from Hulene-B dump, W = 4 and C = 2 (Fig. 7);
313 (7) rainfall, average annual precipitation of ~789 mm, with precipitation influencing decomposition and
314 leachate production, W = 2 and C = 2 were assigned; and (8) existence and distance of surface water, W
315 of the dump surface water occupying the surface of the dune depression every year and other cyclically
316 floods at more distant areas, W = 2 and C = 4. Thus, contamination risk of surface water was 0.91 (Table
317 4), with almost all factors contributing to higher values, except for the compaction factor, given the
318 process of waste compaction by bulldozer that takes place in recent years.

319 High risk of surface water contamination, associated with Hulene-B dump characteristics and
320 contamination processes occurring by leachate free circulation at various points of the W boundary (Fig.
321 7), associated with ashes deposition of incinerated waste, are vectors of contamination and alteration of
322 soil properties. Bernardo et al. (2022b), suggested a high risk of soil and groundwater contamination
323 mainly by leachate in the surrounding area of Hulene-B waste dump. Naveen et al. (2018), showed high
324 levels of surface water and soil contamination in the surrounding a landfill in Bangalore (India), and
325 associated it to the contaminants contained in leachates. A soil contamination study conducted during
326 rainy season on the same locations (Bernardo et al., 2022f) reported an enrichment of Cr, Cu, Mn, Ni, Pb,
327 Zn, and Zr, linked to the impact of the waste dump.

328 Conclusions

329 Analysis of the temporal variation of soil properties in the surrounding area of Hulene-B dump, revealed
330 soil samples properties during rainy season (2020) characteristic of contaminant leaching, with low OM,
331 moisture, and EC content, when compared to 2021 dry season samples, with conditions for topsoil
332 contaminants accumulation. Clay, silty and organic matter content wer higher in dry period samples.
333 Blackish and brownish color occupied larger areas of dry period samples, suggesting reduced leaching
334 mechanisms. pH showed no significant differences, suggesting that was not a determining factor in
335 contaminant retention and migration. Risk of surface water contamination was high, consistent with rainy
336 season indexes, suggesting that the dump act as a vector of contamination to soils, and surface water and
337 groundwater, in the surroundings of Hulene-B dump. Future research is needed to assess soil and surface
338 water heavy metal content in both seasons, that complement physical properties data and contamination
339 risk assessment. However, these results were good indicator of the need for structural measures to
340 control the main sources of contamination, mainly surface leachates and ashes from wastes incineration.

341 Funding: This work was partially supported by GeoBioTec (UIDB/04035/2020) Research Centre, funded
342 by FEDER funds through the Operational Program Competitiveness Factors COMPETE and by National
343 funds through FCT. The first author acknowledges grant from the Portuguese Institute Camões and FNI
344 (Investigation National Fund—Mozambique).

345 Competing Interests: The authors declare no conflict of interest.

346 Author Contributions: Conceptualization, B.B., C.C., F.R.; methodology, B.B., C.C.; validation, B.B. , C.C.,
347 F.R.; formal analysis, B.B., C.C.; investigation, B.B., C.C., F.R.; writing—original draft preparation, B.B. , C.C.;
348 writing—review and editing, C.C., F.R.; supervision, C.C., F.R.; funding acquisition, B.B and F.R. All authors
349 have read and agreed to the published version of the manuscript.

350 References

351 Alghamdi AG, Aly AA, Ibrahim HM (2021) Assessing the environmental impacts of municipal solid waste
352 landfill leachate on groundwater and soil contamination in western Saudi Arabia. Arab J Geos.
353 https://doi.org/10.1007/s12517-021-06583-9

354 Alleoni L, Iglesias C, Mello S, Camargo O, Casagrande J, Lavorenti N (2005) Soil attributes related to
355 cadmium and copper adsorption in tropical soils. Acta Scientiarum. Agron.
356 https://doi.org/10.4025/actasciagron.v27i4.1348

357 Aryampa S, Maheshwari B, Sabiiti E, Olobo C, Bateganya N (2021) Journal of African Earth Sciences
358 Adaptation of EVIAVE methodology to landfill environmental impact assessment in Uganda – A case
359 study of Kiteezi landfill. J Afr Earth Sci. https://doi.org/10.1016/j.jafrearsci.2021.104310

360 Ashraf M, Zeshan M, Hafeez S, Hussain R, Qadir A, Majid M, Ahmad S (2022) Temporal variation in leachate
361 composition of a newly constructed landfill site in Lahore in context to environmental pollution and
362 risks. Environ Sci Pollut Res. https://doi.org/10.1007/s11356-022-18646-9

363 Audu P, Wuana R (2021) Evaluating the Levels and Human Health Risks of Heavy Metals in Soils around
364 Onne Landfill, Rivers State, Nigeria. Int Res J Pure Apl Chem 22(3):43–59.
365 https://doi.org/10.9734/irjpac/2021/v22i330395
366 Azeez J, Hassan O, Egunjobi P (2011) Soil contamination at dumpsites: Implication of soil heavy metals
367 distribution in municipal solid waste disposal system: A case study of Abeokuta, southwestern
368 Nigeria. Soil Sed Cont 20(4):370–386. https://doi.org/10.1080/15320383.2011.571312

369 Bernardo B, Candeias C, Rocha F (2022a) Integration of Electrical Resistivity and Modified DRASTIC Model
370 to Assess Groundwater Vulnerability in the Surrounding. Water.
371 https://doi.org/10.3390/w14111746

372 Bernardo B, Candeias C, Rocha F (2022b) Application of Geophysics in geo-environmental diagnosis on


373 the surroundings of the Hulene-B waste dump, Maputo, Mozambique. J Afr Earth Sci.
374 https://doi.org/10.1016/j.jafrearsci.2021.104415

375 Bernardo B, Candeias C, Rocha F (2022c) Soil properties and environmental risk assessment of soils in the
376 surrounding area of Hulene-B waste dump, Maputo (Mozambique). Environ Earth Sci.
377 https://doi.org/10.1007/s12665-022-10672-7

378 Bernardo B, Candeias C, Rocha F (2022d) Characterization of the Dynamics of Leachate Contamination
379 Plumes in the Surroundings of the Hulene-B Waste Dump in Maputo, Mozambique. Environments.
380 https://doi.org/10.3390/environments9020019

381 Bernardo B, Candeias C, Rocha F (2023e) The Contribution of the Hulene-B Waste Dump (Maputo,
382 Mozambique) to the Contamination of Rhizosphere Soils, Edible Plants, Stream Waters, and
383 Groundwaters. Environments. https://doi.org/10.3390/environments10030045

384 Bernardo B, Candeias C, Rocha F (2022f) Soil Risk Assessment in the Surrounding Area of Hulene-B Waste
385 Dump, Maputo (Mozambique). Geosciences. https://doi.org/10.3390/geosciences12080290

386 Calvo F, Moreno B, Zamorano M, Szanto M (2005) Environmental diagnosis methodology for municipal
387 waste landfills. Waste Manag 25:768–779. https://doi.org/10.1016/j.wasman.2005.02.019

388 Campos C (2010) Soil attributes and risk of leaching of heavy metals in tropical soils. Ambiência 6(3):547–
389 565. https://revistas.unicentro.br/index.php/ambiencia/article/view/591

390 Cendón DI, Haldorsen S, Chen J, Hankin S, Nogueira G, Momade F, Stigter T (2020) Hydrogeochemical
391 aquifer characterization and its implication for groundwater development in the Maputo district,
392 Mozambique. Quaternary Int 547:113–126. https://doi.org/10.1016/j.quaint.2019.06.024
393 Chaudhary R, Nain P, Kumar A (2021) Temporal variation of leachate pollution index of Indian landfill sites
394 and associated human health risk. Enviro Sci Poll Res 28(22):28391–28406.
395 https://doi.org/10.1007/s11356-021-12383-1

396 CIAT (2017) Climate-Smart Agriculture in Mozambique. Climate-Smart Agriculture in Mozambique. Center
397 for Tropical Agriculture. https://climateknowledgeportal.worldbank.org/sites/default/files/2019-
398 06/CSA-in-Mozambique.pdf. Acessed 8 October 2022

399 CMCM (2020) Quadro da Política de Reassentamento - QPR Novembro de 2020. Conselho Municipal da
400 Cidade de Maputo.
401 https://documents1.worldbank.org/curated/en/964181607106676324/pdf/Revised-Resettlement-
402 Framework-Maputo-Urban-Transformation-Project-P171449.pdf. Acessed 10 October 2022

403 Dakheel A, Bashir M, Borrajo J (2022) Appraisal of groundwater contamination from surface spills of fluids
404 associated with hydraulic fracturing operations. J Sci Total Env.
405 https://doi.org/10.1016/j.scitotenv.2022.152949

406 Fatoba J, Eluwole A, Sanuade O, Hammed O, Igboama W, Amosun J (2021) Geophysical and geochemical
407 assessments of the environmental impact of Abule-Egba landfill, southwestern Nigeria. Model Earth
408 Syst Env 7(2):695–701. https://doi.org/10.1007/s40808-020-00991-8

409 Ferrão D (2006) Evaluation of removal and disposal of solid waste in Maputo City, Mozambique. Master’s
410 thesis, University of Cape Town.
411 https://open.uct.ac.za/bitstream/handle/11427/4851/thesis_sci_2006_ferrao_d_a_g.pdf?sequenc
412 e=1. Acessed 10 September 2022

413 Ghobadi R, Altaee A, Zhou J, Karbassiyazdi E, Ganbat N (2021) Effective remediation of heavy metals in
414 contaminated soil by electrokinetic technology incorporating reactive filter media. J Sci Total Env.
415 https://doi.org/10.1016/j.scitotenv.2021.148668

416 Hasan M, Ahmad S, Mohammed T (2021) Groundwater Contamination by Hazardous Wastes. Arab J Sci
417 Eng 46(5):4191–4212. https://doi.org/10.1007/s13369-021-05452-7

418 Hoai S, Lan H, Viet N, Hoang G, Kawamoto K (2021) Characterizing seasonal variation in landfill leachate
419 using leachate pollution index (LPI) at nam son solid waste landfill in Hanoi, Vietnam. Environments.
420 https://doi.org/10.3390/environments8030017

421 Hussein M, Yoneda K, Mohd-Zaki Z, Amir A, Othman N (2021) Heavy metals in leachate, impacted soils
422 and natural soils of different landfills in Malaysia: An alarming threat. Chemosphere.
423 https://doi.org/10.1016/j.chemosphere.2020.128874

424 Ihedioha J, Ukoha P, Ekere N (2017) Ecological and human health risk assessment of heavy metal
425 contamination in soil of a municipal solid waste dump in Uyo, Nigeria. Enviro Geoc Health 39(3):497–
426 515. https://doi.org/10.1007/s10653-016-9830-4

427 INE (2020) Boletim de Estatísticas Demográficas e Sociais, Maputo Cidade 2019. Instituto Nacional de
428 Estatistica. http://www.ine.gov.mz/estatisticas/estatisticas-demograficas-e-indicadores-
429 sociais/boletim-de-indicadores-demograficos-22-de-julho-de-2020.pdf/at_download/file. Acessed
430 22 October 2022

431 Karimi H, Herki B, Gardi S, Galalizadeh S, Hossini H, Mirzaei K, Pirsaheb M (2022) Site selection and
432 environmental risks assessment of medical solid waste landfill for the City of Kermanshah-Iran. Int J
433 Environ Health Res 32(1):155–167. https://doi.org/10.1080/09603123.2020.1742876

434 Kumar M, Furumai H, Kurisu F, Kasuga I (2013) Potential mobility of heavy metals through coupled
435 application of sequential extraction and isotopic exchange: Comparison of leaching tests applied to
436 soil and soakaway sediment. Chemosphere 90(2):796–804.
437 https://doi.org/10.1016/j.chemosphere.2012.09.082

438 Li H, Sun J, Gui H, Xia D, Wang Y (2022) Physiochemical properties, heavy metal leaching characteristics
439 and reutilization evaluations of solid ashes from municipal solid waste incinerator plants. Waste
440 Manag 138(422):49–58. https://doi.org/10.1016/j.wasman.2021.11.035

441 Lund E (2008) Soil electrical conductivity. In: Logsdon S, Clay D, Moore D, Tsegaye T (ed) Soil Science: Step-
442 by-Step Field Analysis. https://doi.org/10.2136/2008.soilsciencestepbystep.c11

443 Mandal S, Bhattacharya S, Paul S (2022) Assessing the level of contamination of metals in surface soils at
444 thermal power area: Evidence from developing country (India). Environ Chem Ecotox 4:37–49.
445 https://doi.org/10.1016/j.enceco.2021.11.003

446 Möller A, Müller H, Abdullah A, Abdelgawad G, Utermann J (2005) Urban soil pollution in Damascus, Syria:
447 Concentrations and patterns of heavy metals in the soils of the Damascus Ghouta. Geoderma 124(1–
448 2):63–71. https://doi.org/10.1016/j.geoderma.2004.04.003

449 Momade F, Ferrara M, Oliveira J (1996) Notícia explicativa da carta geológica 2532 Maputo (Escala 1:50
450 000). Maputo. In Portuguese.

451 Morita A, Ibelli-Bianco C, Anache J, Coutinho J, Pelinson N, Nobrega J, et al. (2021). Pollution threat to
452 water and soil quality by dumpsites and non-sanitary landfills in Brazil: A review. Waste Manag
453 131:163–176. https://doi.org/10.1016/j.wasman.2021.06.004

454 Muchangos A (1999) Paisagens e Regiões Naturais de Moçambique. 5–163.


455 https://docplayer.com.br/47220681-Mocambique-paisagens-e-regioes-naturais.html. Accessed 7
456 November 2022

457 Muchimbane A (2010) Estudo dos indicadores da contaminação das aguas subterrâneas por sistemas de
458 saneamento in Situ – Distrito Urbano 4, Cidade de Maputo, Moçambique. Dissertation, São Paulo
459 University

460 Munsell (2009) Munsell Soil Color Chart. with Genuine Munsell Color Chips. Grand Rapids, MI : Munsell
461 Color

462 Nassereddine M, Rizk J, Nasserddine G (2013) Soil Resistivity Structure and Its Implication on the Pole Grid
463 Resistance for Transmission Lines. Int J Elect Comp Eng. https://doi.org/10.5281/zenodo.1063250

464 Naveen B, Mahapatra D, Sitharam T, Sivapullaiah P, Ramachandra T (2017) Physico-chemical and


465 biological characterization of urban municipal landfill leachate. Environ Poll 220:1–12.
466 https://doi.org/10.1016/j.envpol.2016.09.002

467 Naveen B, Sumalatha J, Malik R (2018) A study on contamination of ground and surface water bodies by
468 leachate leakage from a landfill in Bangalore, India. Int J Geo-Eng. https://doi.org/10.1186/s40703-
469 018-0095-x

470 Nilam T, Ibrahim T, Mahmood N, Othman F (2016) Estimation of leachaAte generation from MSW landfills
471 in estimation of leachate generation from MSW. Asian J Microbiol Biotech Environ Sci 19(1):43–48

472 Nisari A, Sujatha C (2021) Assessment of trace metal contamination in the Kol wetland , a Ramsar site ,
473 Southwest coast of India. Reg Stud Marine Sci. https://doi.org/10.1016/j.rsma.2021.101953

474 Nogueira G, Stigter T, Zhou Y, Mussa F, Juizo D (2019) Understanding groundwater salinization
475 mechanisms to secure freshwater resources in the water-scarce city of Maputo, Mozambique. Sci
476 Total Environ 661:723–736. https://doi.org/10.1016/j.scitotenv.2018.12.343

477 Ozbay G, Jones M, Gadde M, Isah S, Attarwala T (2021) Design and Operation of Effective Landfills with
478 Minimal Effects on the Environment and Human Health. J Environ Public Health.
479 https://doi.org/10.1155/2021/6921607

480 Pagnanelli F, Esposito A, Toro L (2003) Metal speciation and pH effect on Pb , Cu , Zn and Cd biosorption
481 onto Sphaerotilus natans : Langmuir-type empirical model. Water Res 37:627–633

482 Reeuwijk L (2002) Procedures for Soil Analysis. International Soil Reference and Information Centre,
483 Wageningen

484 Rieuwerts J, Ashmore M, Farago M, Thornton I (2006) The influence of soil characteristics on the
485 extractability of Cd, Pb and Zn in upland and moorland soils. Sci Total Environ 366(2–3):864–875.
486 https://doi.org/10.1016/j.scitotenv.2005.08.023

487 Sarmento L, Tokai A, Hanashima A (2015) Analyzing the structure of barriers to municipal solid waste
488 management policy planning in Maputo city , Mozambique. Environ Develop 16:76–89.
489 https://doi.org/10.1016/j.envdev.2015.07.002

490 Serra C (2012) Da problemática Ambiental à mudança: rumo à um mundo melhor. Escolar Editora, Lisboa

491 Shaylinda M (2020) Metals contamination on soil and surface water ( earth drainage ) due to leachate
492 migration from Piyungan land Metals contamination on soil and surface water ( earth drainage ) due
493 to leachate migration from Piyungan land. Mat Sci Eng. https://doi.org/10.1088/1757-
494 899X/1144/1/012063

495 Siddiqi S, Al-Mamun A, Sana A, Baawain M, Choudhury M (2022) Characterization and pollution potential
496 of leachate from urban landfills during dry and wet periods in arid regions. Water Supply 22(3):3462–
497 3483. https://doi.org/10.2166/ws.2021.392

498 Sparling G (2020) Soil Quality: Indicators. In: Fath B, Jorgensen S (ed) Soil Quality: Indicators, 2nd edn. CRC
499 Press, Boca Raton

500 Sun R, Gao Y, Yang Y (2022) Leaching of heavy metals from lead-zinc mine tailings and the subsequent
501 migration and transformation characteristics in paddy soil. Chemosphere 291:132792.
502 https://doi.org/10.1016/j.chemosphere.2021.132792

503 Tao Z, Deng H, Li M, Chai X (2020) Mercury transport and fate in municipal solid waste landfills and its
504 implications. Biogeochem 148(1):19–29. https://doi.org/10.1007/s10533-020-00642-1

505 USDA (2011a) Rangeland Soil Quality - Organic Matter. United States Department of Agriculture
506 http://www.ftw.nrcs.usda.gov/glti. Accessed 30 October 2022

507 USDA (2011b) Soil Health Quality Indicators: chemical Properties, soil electrical conductivity. United States
508 Department of Agriculture
509 https://www.nrcs.usda.gov/wps/portal/nrcs/detail/soils/health/assessment/?cid=stelprdb123738
510 7. Accessed 30 October 2022

511 Vicente E (2011) Aspects of the enginering geologic of Maputo City. Dissertation, University of KwaZulu-
512 Natal

513 Vicente E, Jermy C, Schreiner H (2006) Urban geology of Maputo, Mozambique. Geolog Soc London 338:1–
514 13.

515 Weibel G, Eggenberger U, Schlumberger S, Mäder U (2017) Chemical associations and mobilization of
516 heavy metals in fly ash from municipal solid waste incineration. Waste Manag 62:147–159.
517 https://doi.org/10.1016/j.wasman.2016.12.004

518 Wijekoon P, Koliyabandara P, Cooray A, Lam S, Athapattu B, Vithanage M (2022) Progress and prospects
519 in mitigation of landfill leachate pollution: Risk, pollution potential, treatment and challenges. J
520 Hazard Mat. https://doi.org/10.1016/j.jhazmat.2021.126627

521 Wu G, Wang L, Yang R, Hou W, Zhang S, Guo X, Zhao W (2022) Pollution characteristics and risk assessment
522 of heavy metals in the soil of a construction waste landfill site. Ecolog Inform.
523 https://doi.org/10.1016/j.ecoinf.2022.101700

524 Xiao H, Shahab A, Xi B, Chang Q, You S, Li J, Li X (2021) Heavy metal pollution, ecological risk, spatial
525 distribution, and source identification in sediments of the Lijiang River, China. Environ Pol.
526 https://doi.org/10.1016/j.envpol.2020.116189

527 Yan Z, Jiang Y, Chen X, Lu Z, Wei Z, Fan G, Qu F (2021) Evaluation of applying membrane distillation for
528 landfill leachate treatment. Desalination. https://doi.org/10.1016/j.desal.2021.115358

529 Zaki K, Karhat Y, Falaki K (2022) Temporal Monitoring and Effect of Precipitation on the Quality of Leachate
530 from the Greater Casablanca Landfill in Morocco. Pollu 2:407–433.
531 https://doi.org/10.22059/poll.2021.328358.1158

532 Zhang W, Qiu Q (2010) Analysis on contaminant migration through vertical barrier walls in a landfill in
533 China. Environ Earth Sci 61:847–852. https://doi.org/10.1007/s12665-009-0399-4

534 Zeitoun R, Vandergeest M, Vasava HB, Machado P, Jordan S, Parkin G, Biswas A (2021) In-situ estimation
535 of soil water retention curve in silt loam and loamy sand soils at different soil depths. Sensors
536 21(2):1–15. https://doi.org/10.3390/s21020447

537

538 Figures captions

539 Figure 1. Sampling location and environmental context: (a) dune depression with surface water; (b)
540 Hulene-B waste dump; and (c, d) areas considered not impacted (background).

541 Figure 2. Soil texture classification of studied soils (blue) and background (yellow) samples.

542 Figure 3. Soil samples pH variation (2020 vs. 2021) in (a) sand, and (b) silt fractions.

543 Figure 4. Soil samples EC variation (2020 vs. 2021) in (a) sand, and (b) silt fractions.

544 Figure 5. Soil samples variation (2020 vs. 2021) (a) OM content, (b) color, and (c) moisture content.

545 Figure 6. Principal component analysis variables distributed by component.

546 Figure 7. (a) Leachate at the W boundary of Hulene-B; (b) leachate at the S boundary of Hulene-B; and
547 (c) surface water.

548

549 Tables captions


550 Table 1. Granulometric statistical information (in %).

551 Table 2. Soil physical parameters of background (Bkg) 2020 and 2021 samples.

552 Table 3. Spearman correlation.

553 Table 4. Contamination risk of surface water around Hulene-B waste dump.
Figure1
Figure2
Figure3
Figure4
Figure5
Figure6
Figure7
Table1

studied soils background


fraction
min max mean SD min max mean SD
sand 81.92 97.86 93.94 3.84 92.41 95.36 93.84 0.79
silt 1.34 12.71 4.49 2.98 2.33 3.81 3.09 0.41
clay 0.59 5.37 1.56 0.93 2.32 3.78 3.07 0.39
Table2

2020 Bkg (n = 10) 2021 Bkg (n = 10)


var
min max mean±SD min max mean±SD
pH (< 2 mm) 4.2 6.3 5.3±0.88 5.2 6.2 5.6±0.38
pH (< 63 µm) 5.7 6.4 6.2±0.25 6.1 6.7 6.4±0.2
EC (< 2mm) 9.8 28.1 18.3±7.27 11.8 34.3 21.1±9.29
EC (< 63 µm) 10.8 16.5 13.9±1.91 11.0 19.0 13.5±2.33
OM 0.3 0.6 0.5±0.1 0.4 1.2 0.8±0.22
Moisture 0.2 0.6 0.4±0.1 0.2 2.0 0.7±0.47
var – variables; Bkg – background; min – minimum; max –
maximum; SD – standard deviation; Sand, Silt, and Clay in
%; EC – electrical conductivity (in μS/cm); OM – organic
matter (in %); Moisture in %.
Table3

sand silt clay OM EC pH


silt -0.995** 1
clay -0.954** 0.928** 1
OM -0.738** 0.746** 0.690** 1
EC -0.578** 0.601** 0.453** 0.637** 1
pH 0.225 -0.231 -0.201 -0.282 -0.048 1
moisture -0.650** 0.647** 0.614** 0.601** 0.315* -0.254
** p < 0.01; * p < 0.05
Table4

variables W*C
Compaction 6
Daily Cover 8
Control of leachate 8
Final coverage 8
Type of waste and % organic matter 8
Waterproofing the discharge point 8
Rainfall 4
Existence and distance of surface water 8
Contamination risk High 0.91
Supplementary material
Soil properties temporal variation and risk assessment in the surroundings of Hulene-B dump,
Maputo (Mozambique)

Bernardino Bernardoa,b, Carla Candeiasb Fernando Rochab

a
GeoBioTec Research Centre, Department of Geosciences, University of Aveiro, 3810-193, Aveiro,
Portugal

b
Faculty of Earth Sciences and Environment, Pedagogic University of Maputo, Mozambique

Table S1. Compaction Ranking - Assessment criteria and assigned scores (in green, factor selected considering the
dump characteristics).
Classification (Cj ) Condition
Very low 0 The landfill is a bale landfill with adequate exploitation. Very high compaction
Under 1 The landfill is bales with exploitation is considered to be regular. High compaction
Medium 2 The landfill is a bale dump but the operation is deficient Medium compaction
High 3 High-density landfills with poor operation. Low compaction
Very high 4 Low-density landfills with regular operation; Landfills without any compaction Very low compaction

Table S2. Daily Coverage - Assessment criteria and assigned scores (in green, factor selected considering the dump
characteristics).
Classification
Conditions
(Cj )
Very low 0 Suitable material with satisfactory installation. Very satisfactory
Adequate material with average workmanship or average material with satisfactory
Under 1 Satisfactory
workmanship.
Adequate covering material with deficient installation, average covering material
Medium 2 with satisfactory installation or no covering material at all with appropriate Regular
commissioning.
Average material with poor workmanship or unsuitable material with average
High 3 Deficient
workmanship.
Inadequate or
Very high 4 Unsuitable covering material with inadequate or no covering material at all.
non-existent
Table S3. Control Leachte - Assessment criteria and assigned scores (in green, factor selected considering the dump
characteristics).

Classification (Cj ) Conditions


There is control of the volume and composition of the leachate the drainage systems are in Very
Very low 0
good condition, there are storage tanks and there is leachate treatment. suitable
Low 1 There is control of the volume and composition of the leachate the drainage systems Adequate
There is a drainage and storage system with treatment or recirculation with design and/or
Medium 2 Medium
conservation problems.
High 3 There is no control of the volume and composition of the leachate. Under
Very high 4 There is no control, no leachate drainage, no storage and no treatment. Null

Table S4. Final Cover - Assessment criteria and assigned scores (in green, factor selected considering the dump
characteristics).
Classification (Cj ) Conditions
Very
Very low 1 Full coverage including existence of gas drainage layer. gas drainage layer.
suitable
Low 2 Properly covered and without gas drainage layers Adequate
Medium 3 Existence of weaknesses in the impermeable layer requirements Regular
The requirements for the impermeable layer and/or the requirements for the drainage
High 4 Deficient
level are not met.

Table S5. Type of waste - Assessment criteria and assigned scores (in green, factor selected considering the dump
characteristics).
Classification
Conditions
(Cj )
Very low
Very low 0 Landfill for non-hazardous waste or pre-sorting disposal.
polluting power
Landfill for non-hazardous waste. Sub-category for waste with a low degree of pre- Low polluting
Under 1
sorting and presence of organic material power
Landfill of waste with a high percentage of organic matter from waste not subject to Average
Medium 2
pre-treatment polluting power
Landfill of waste with a high percentage of organic matter from waste that has not
High polluting
High 3 been subject to pre-treatment to separate the organic fraction with the presence of
power
inert waste
Waste landfill with a high percentage of organic matter from waste without pre- Very high
Very high 4
treatment, significant presence of organic matter and hazardous waste polluting power

Table S6. Waterproofing the discharge point - Assessment criteria and assigned scores (in green, factor selected
considering the dump characteristics).
Classification
Conditions
(Cj )
Very
Very low 0 Artificial lining placed over the whole of the vessel
high
Under 1 At the base and sides of the landfill there will be a mineral layer with permeability High
The natural waterproofing in the basin and on the sides is in good condition, although the
Medium 2 Medium
artificial waterproofing is not, as it is damaged
The natural waterproofing of the basin and the sides does not meet the requirements set out
High 3 Low
in the high waterproofing point but does meet the specifications of artificial waterproofing.
None of the natural and artificial waterproofing requirements for the basin and sides of the
Very high 4 Very low
discharge point set out in point high waterproofing.
Table S7. Rainfall assessment criteria and assigned scores (in green, factor selected considering the dump
characteristics).
Classification (Cj ) Conditions
Very low 0 Less than 300 mm Very low rainfall
Under 1 300-600 mm Low rainfall
Medium 2 600-800 mm Average rainfall
High 3 800-1000 mm High rainfall
Very high 4 More than 1000 mm Very high rainfall

Table S8. Distance to surface water bodies assessment criteria and assigned scores (in green, factor selected
considering the dump characteristics).

Classification (Cj ) Conditions


Very low 0 Surface waters at distances of more than 1000m Very high
Under 1 Surface waters located at distances between 1000-700m. High
Medium 2 Surface waters located at distances between 700-300m. Medium
High 3 Surface waters between 300 - 50 m. Low
Very high 4 Surface water less than 50 m or waste in direct contact with water surface Very low
308
309

13.2.5 Appendix 6. Paper 5


310
geosciences

Article
Soil Risk Assessment in the Surrounding Area of Hulene-B
Waste Dump, Maputo (Mozambique)
Bernardino Bernardo 1,2 , Carla Candeias 1, * and Fernando Rocha 1

1 GeoBioTec Research Centre, Department of Geosciences, University of Aveiro, 3810-193 Aveiro, Portugal;
nhacundela.berna@gmail.com (B.B.); tavares.rocha@ua.pt (F.R.)
2 Faculty of Earth Sciences and Environment, Pedagogic University of Maputo, Maputo 2482, Mozambique
* Correspondence: candeias@ua.pt

Abstract: Soil contamination in areas close to unplanned dumpsites represents an increasing risk to
the ecosystems and human health. This study aimed to evaluate soil quality in the area surrounding
the Hulene-B waste dump, Maputo, Mozambique, and to estimate potential ecological and human
health risks. A total of 71 surface soil samples were collected in the surrounding area of the dump,
along with 10 samples in areas considered not impacted by the dump. Chemical and mineralogical
analyses were performed using XRF and XRD. Quartz was the most abundant mineral phase,
followed by feldspars, carbonates, clay minerals, and Fe oxides/hydroxides. Results showed a significant
contribution to ecological degradation by PTE enrichment, ranked as Zn >> Cu > Cr > Zr > Pb > Ni > Mn.
Carcinogenic risk for both children and adults was significant due to Pb soil content. Soil sample
concentrations of Cr, Cu, Mn, Ni, Pb, Zn, and Zr, posing a risk especially in children, suggested the
need for continuous monitoring, as well as the definition and implementation of mitigation measures.

Keywords: soils; waste dump; potential toxic elements; risk assessment

Citation: Bernardo, B.; Candeias, C.;


Rocha, F. Soil Risk Assessment in the
Surrounding Area of Hulene-B Waste 1. Introduction
Dump, Maputo (Mozambique).
Soil contamination is an environmental problem affecting urban areas, particularly
Geosciences 2022, 12, 290.
in developing countries [1,2]. Contamination sources have been attributed to unregulated
https://doi.org/10.3390/
overlapping of activities that influence soil quality, e.g., industrialization, urbanization,
geosciences12080290
agriculture, and solid waste [3]. Solid waste heterogeneity contributes with the enrichment
Academic Editors: Lisetskii Fedor, of potentially toxic elements (PTEs) in soil [4], being currently considered as a major threat
Javier Lillo and Jesus Martinez-Frias to soil and groundwater quality in urban areas [5]. Previous studies [6] described that the
Received: 2 June 2022
rapid growth of cities and their inhabitants has boosted global waste production, with
Accepted: 25 July 2022
an estimated global annual production of municipal solid waste (MSW), in 2025, of 2.2
Published: 28 July 2022
billion metric tons [7,8]. Heavy metals are among the most hazardous soil pollutants, being
highly reactive; when organisms are unable to eliminate them chemically, they are retained
Publisher’s Note: MDPI stays neutral
in ecosystem, representing a potential risk [9,10]. Areas surrounding waste dump sites,
with regard to jurisdictional claims in
are prone to PTE contamination [11], with several studies reporting soil contamination,
published maps and institutional affil-
e.g., ref. [12] with high soil contamination by As, Hg, and Al in soil around a dumpsite
iations.
in Italy, ref. [13] identifying Cu, Pb, Co, Mn, and Hg soil contamination in China, [14]
evidencing high levels of contamination by As, Cd, Pb, and Cr in seven dumpsites in
Malaysia, and [15] finding Cd, Zn, Cu, Pb, Ni, and Co contamination. Soil contamination
Copyright: © 2022 by the authors.
around waste dumps poses a risk to human health [16]. A study near a dumpsite in Kenya
Licensee MDPI, Basel, Switzerland. suggested that high soil PTE levels (Pb, Hg Cd, Cu, and Cr) pose a risk to children and
This article is an open access article adolescents living and studying near the dumpsite, who presented a high incidence of
distributed under the terms and respiratory diseases and blood Pb levels, along with blood cell aberrations. The authors
conditions of the Creative Commons of [17,18] systematized data from waste dumps in developing countries and concluded
Attribution (CC BY) license (https:// that the levels of contamination caused by PTEs posed potential health risks. In turn,
creativecommons.org/licenses/by/ the authors of [19] showed that the high levels of heavy-metal contamination of the soil
4.0/). environment of the Udoyo dumpsite in Nigeria constituted a risk factor for human health.

Geosciences 2022, 12, 290. https://doi.org/10.3390/geosciences12080290 https://www.mdpi.com/journal/geosciences


Geosciences 2022, 12, 290 2 of 15

The geochemical heterogeneity of soil worldwide has led to the need for a definition of
a local background [20], relevant for reducing discrepancy between world established
guidelines and local soils [21], thereby providing a reliable estimate of the influence of
contamination according to measured element concentrations [22]. In Mozambique, one
cause of soil contamination is the unregulated disposal of solid wastes, especially in the
capital, Maputo [23]. In Maputo city, MSW production is increasing, with an estimated
daily waste production of 1250 t [23,24], of which ~1000 t is deposited in the largest open-air
dump of the city, the Hulene-B dump [23]. All types of waste, e.g., domestic, construction,
medical, sanitary, and industrial, are deposited in the dump without any treatment, posing
a worrying ecological and public health risk. The authors of [23,25] linked Hulene-B waste
to the Hg, As, Pb, Cu, and Zn contamination of the surrounding soils. The authors of [24]
suggested that high electric conductivities of the soils surrounding the Hulene-B dump
may be associated with a possible enrichment of the PTEs, with origin in dump leachates.
Studies by [23,24,26] reported that the dump may be impacting the health of the population
around the Hulene-B dump. However, studies on the chemical quantification of the levels
of enrichment of the surroundings of the dump by PTEs are scarce, as are models of the
impact of ecological systems and human health. Soil quality evaluation can be assessed
by using world guidelines and local background concentrations, which can contribute to
the improvement of methodologies. The integration of ecological and human health risk
assessment models is significant to understand the impact of the Hulene-B dumpsite and
to assist in the definition of mitigation measures. The present study aimed to assess soil
contamination by Cr, Cu, Mn, Ni, Pb, Zn, and Zr in the area surrounding Hulene-B waste
dump, Maputo, and to estimate potential ecological and human health risk.

2. Materials and Methods


2.1. Study Area
The Hulene-B dump is located in the northern part of Maputo city, the capital of
Mozambique (Figure 1). The surrounding area is densely inhabited, with over 75,500 resi-
dents [27]. On the eastern edge of the dump is Julius Nyerere Avenue, a road with heavy
traffic that provides access to Maputo city center, while the largest airport in Mozambique
(Maputo International Airport) is located on the west side. The dump covers an area of
~17 hectares [23,28], occupying an abandoned (1973) open sand pit, where the population
started to deposit all type of wastes without previous preparation [29]. The height of
the waste is estimated at 6–15 m [28]. The dump is characterized by open-air deposition
and was described by [25] as the source of soil and groundwater contamination by Hg,
As, Pb, Cu, and Zn. Recently, the authors of [30] described the Hulene-B dump as one
cause of contaminated particle (re)suspension and deposition as a result of uncontrolled
waste incineration.
Geologically, the Hulene-B dump is in the Mesocenozoic sedimentary basin of southern
Mozambique [31], in contact with two lithologies, the Ponta Vermelha formation (TPv) at the
dump eastern limit and the Malhazine formation (QMa) at the dump western limit [32]. The
Malhazine formation consists of coarse to fine, poorly consolidated sands with whitish to
reddish colors, fixed by vegetation, as a result of successive consolidation processes [32,33].
The Ponta Vermelha formation is composed by ferruginous sandstone and red sedimentary
sands that gradually passe to yellow and whitish sands [32]. Arenosols are dominant in the
area, with sandy texture [34]. The predominant climate is of subtropical type, with mean
annual precipitation of ~789 mm, with two climatic seasons: (a) a hot (mean 25 ◦ C) and
rainy period from December to March with >60% of the annual precipitation, peaking in
January (~125 mm), and (b) a dry and cold season, from April to September, with lower
temperatures in June and July (mean 21 ◦ C), and scarce precipitation, whose minimum
values are recorded in August (~12 mm) [35]. The prevailing winds are SE [36].
Geosciences 2022, 12, 290 3 of 15

Figure 1. Study area with (a) dune depression, (b) Hulene-B waste dump, and (c,d) areas considered
not impacted.

2.2. Sampling and Analysis


A total of 71 soil samples were collected in the surrounding area of the Hulene-B dump,
Maputo (Mozambique), in January 2020 (Figure 1). To determine the local background,
10 samples were collected in areas considered not impacted by the dump. Local background
sampling sites were selected taking into consideration not only the similar lithologies of
the study area, but also the distance, wind direction, and topography: (a) eastern sampling
site, located ~1 km distance from Hulene-B dump on a high relief with opposite prevailing
wind direction (SE); (b) northern site, ~2 km distance from Hulene-B, with opposite wind
direction, representing an area not impacted by the dump.
Samples were collected superficially (0–20 cm), georeferenced, and preserved in plastic
bags until laboratorial analysis. In the laboratory, samples were oven-dried at ≤40 ◦ C
(Pedagogical University of Maputo, Mozambique). Afterward, samples were transported
to the laboratories of the GeoBioTec Research Center, University of Aveiro (UAVR), Portugal,
for analyses. Soil samples were sieved at <2000 and <63 µm.
The chemical composition of samples was accessed by X-ray fluorescence (XRF) spec-
trometry, using a PANalytical PW 4400/40 45 Axios with Cr Kα radiation. Mineralogical
phases were determined by powder X-ray diffraction (XRD) using a Phillips/Panalytical
powder diffractometer, model X’Pert-Pro MPD, equipped with an automatic slit. A Cu
X-ray tube was operated at 50 kV and 30 mA. Data were collected from 2◦ to 70◦ 2θ with a
step size of 1◦ and a counting interval of 0.02 s. The precision and the accuracy of analyses
and procedures were monitored using UAVR internal standards, certified reference material
(BGS119), and quality control blanks. Results were within the 95% confidence limits. The
relative standard deviation was between 5% and 10%.
Geosciences 2022, 12, 290 4 of 15

2.3. Data Analysis


Descriptive statistics, principal component analysis (PCA), Spearman correlation, one-
way ANOVA and cluster analysis were performed using SPSS v.25® software (IBM, New
York, NY, USA). Cluster analysis was applied using Ward’s method and Euclidean distance
as a measure of similarity.
Potential ecological risk (PERI) has been widely used to assess the degree of po-
tential ecological risk of heavy metals in soils, which was originally used in sediment
samples [37]. The potential ecological risk factor is defined as the sum of the monomial
potential ecological risk factors (EF), which quantitatively defines the potential ecological
risk of a contaminant in a sample [38], calculated as PERI = ∑7i =1 EFi = ∑7i =1 CFi × TF,
where PERIi is the potential ecological risk index, EFi the monomial potential ecological
risk factor of each variable, CFi is the single contamination factor, and TF us the heavy
metal toxic-response factor for the seven selected elements, computed as described by [37]
taking into consideration biochemical composition proposed by [39] that have been used
widely (e.g., [40,41]). TF values obtained were Cr = 25, Cu and Zr = 10, Ni and Zn = 5, and
Mn and Pb = 2. The five EF classes and four PERI degrees are presented in Table 1.

Table 1. Monomial potential ecological risk factor (EF) and potential ecological risk index (PERI)
classification levels [37].

EF Classification PERI Classification


0 ≤ EF < 40 Low PERI < 150 Low
40 ≤ EF < 80 Moderate 150 ≤ PERI < 300 Moderate
80 ≤ EF < 160 Considerable 300 ≤ PERI < 60 Considerable
160 ≤ EF < 32 Very high 600 ≤ PERI Very high
320 ≤ EF Very high

The soil Nemerow index (PN) is widely used to assess the contribution of pollutants
to soil contamination [42]. The Nemerow pollution index synthesizes the single pollution
index values of all elements, and it can comprehensively reflect the degree of soil pollu-
tion and qhighlight the impact of element concentration [43]. The index is calculated as
Pn = [(1/n ∑ni=1 PI)2 + (PImax )2 ]/2, where PI is the single element pollution index,
PImax is the PI maximum value in all elements, and n is the number of elements stud-
ied. Soil quality is classified as follows: nonpolluted PN ≤ 0.7, warning line of pollution
0.7 < PN ≤ 1.0, low level of pollution 1.0 < PN ≤ 2.0, moderate level of pollution 2.0 < PN ≤ 3.0,
and high level of pollution PN > 3.0. The soil Nemerow index was calculated using both
local background (PNbkg ) and world soils (PNRC ) [22].
The pollution load index (PLI) provides a simple comparative means to assess the
level of enrichment [38]. It was determined as the n-th root of the product of the nPI, where
n is the number of variables considered; PLI = (PI1 × PI2 × PI3 ×· · · × PIn )1/n . PLI > 1
implies environmental deterioration by elemental pollution.
Noncarcinogenic and carcinogenic risk assessments were calculated taking into con-
sideration that residents, both children and adults, are directly exposed to soils through
three main pathways: (a) ingestion, (b) inhalation, and (c) dermal absorption [44]. The
carcinogenic and noncarcinogenic side effects for each element were computed individ-
ually, considering reference toxicity levels for each variable, as extensively described
in, e.g., [45,46]. Equations used to estimate the chronic daily intake for each exposure
route considered, supplemented by specific quantitative information, were as follows:
CDIing = Csoil × (IngR × EF × ED)/(BW × AT) × 10−6; CDIderm = Csoil (SA × SAF × DA ×
EF × ED)/(BW × AT ) × 10−6 ; CDIinh = Csoil (InR × EF × ED)/(PEF × BW × AT), where
CDIing is the chronic daily intake dose through ingestion (mg/kg·day), Csoil (mg/kg) is the
concentration of the element in soil in this study, IngR (mg/day) is the soil ingestion rate
(100 for children and 200 for adults), InhR (m3 /day) is the inhalation rate (7.6 for children
and 20 for adults), EF (days/year) is the exposure frequency (180), ED is the exposure
Geosciences 2022, 12, 290 5 of 15

period (6 and 24 years for noncarcinogenic effects in children and adults, respectively,
with 70 years as a lifetime for of carcinogenic effects), BW is the average body weight
(15 kg for children and 70 kg for adults), AT (days) is the average time for noncarcinogenic
effects (=ED × 365 days), SA is the exposed skin area (2800 and 5700 cm2 for children and
adults, respectively), SAF is the skin adherence factor (0.2 and 0.07 mg/cm2 for children and
adults, respectively), DA is the dermal absorption factor (0.03 for As and 0.001 for all of the
selected elements), and PEF is the particulate emission factor (1.36 × 109 m3 /kg) [47–50]. For
each selected potentially toxic element and pathway, the noncarcinogenic toxic hazard
was estimated by computing the hazard quotient (HQ) for systemic toxicity (i.e., noncar-
cinogenic risk), where HQ = CDIpathway /RfD. If HQ > 1, noncarcinogenic effects might
occur as the exposure concentration exceeds the reference dose (RfD) [50]. The cumulative
non-carcinogenic hazard index (HI) corresponds to the sum of HQ for each pathway and/or
variable. Similarly, HI > 1 indicates that noncarcinogenic effects might occur, calculated
with HI = ∑HQ = HQing + HQderm + HQinh .
Carcinogenic risk, or the probability of an individual developing any type of cancer
over a lifetime due to exposure to a potential carcinogen (Risk), was estimated by the sum
of total cancer risk for the three exposure routes, Risk = Risking + Riskderm + Riskinh , as
described by [50]. A risk > 1.00 × 10−6 is classified as the carcinogenic target risk, while
values > 1.00 × 10−4 are considered unacceptable [50].

3. Results and Discussion


3.1. Mineralogical and Chemical Characterization
Mineralogical analysis showed that quartz (SiO2 ) was the dominant mineral, rang-
ing 93.1% to 99.6%, except in samples P40 (87.0%) and P60 (68.4%). Other mineral
phases present were ranked as follows: feldspars (K feldspar (KAlSi3 O8 ) and plagio-
clase ((Na,Ca)[(Si,Al)AlSi2 ]O8 ); 0.2% to 6.2%) > carbonates (calcite (CaCO3 ), dolomite
(CaMg(CO3 )2 ), and siderite (FeCO3 ); 0.2% to 5.7%) > kaolinite (Al2 (Si2 O5 )(OH)4 ;
0.1% to 1.4%) > alunite (KAl3 (SO4 )2 (OH)6 ; 0.1% to 1.4%) > anhydrite (CaSO4 ; 0.1% to
1.1%) (Figure 2). Other minerals present in a few samples were magnetite–maghemite
(Fe3 O4 -γ-Fe2 O3 ), zeolites, opal C/CT (SiO2 ·nH2 O), vaterite (CaCO3 ), and melanterite
(Fe2+ (H2 O)6 SO4 ·H2 O). Local background samples revealed a similar trend but showing
some ilmenite (Fe2+ TiO3 ; 0.2% to 0.3%). These results are similar to those described by [32],
who considered the soils surrounding Hulene-B as heterogeneous as a result of remobi-
lization and mixing of the Ponta Vermelha (90% to 95% quartz) and Malhazine formations
(~98% quartz).
The chemical composition of the areas surrounding the waste dump and local back-
ground samples is presented in Table 2. Cd was below the limit of detection (LOD 3.88
mg/kg) in 93% of the samples. A one-way ANOVA revealed significant differences be-
tween soil and local background sample concentrations of Al, Fe, K, Mn, Ni, Rb, Si, V, and
Zr (p < 0.05). Higher concentrations were found in local background samples, except for
Si and Zr, with lower content. The high concentrations of Al, Fe, K, Mn, Ni, Rb, and V in
the local background samples could be associated with the mineralogical and geochemical
composition of Ponta Vermelha formation, characterized by high levels of Al and Fe oxides,
continually remobilized to the Malhazine formation [32]. Additionally, the Malhazine
formation in local background areas does not exhibit high erosional disturbance or surface
runoff, as it is relatively flat. Soils in the surrounding of dump are distributed between a
dune slope and a depression (heterogeneous area of sediments from the Malhazine and
Ponta Vermelha formations), characterized by strong sediment mobilization and cyclic
retention of surface water from the dump and the unchanneled flow of Julius Nyerere
Avenue in the east (Figure 1). Studies by [51,52] suggested that Al, Fe, K, Mn, Ni, Rb, and V
concentrations in soils with strong surface water circulation and high porosity are prone to
strong depletion due to high oxide mobility. The relatively high concentrations of Si in soil
samples could be associated with the soil mineralogical composition, with predominance
of silicate minerals. The feldspar presence was similar to the results in [32]. Studies by [53]
Geosciences 2022, 12, 290 6 of 15

suggested that Fe and Al oxides have the ability to adsorb and retain PTEs. Te authors
of [54] suggested that SiO2 mineral also has the aptitude to adsorb PTEs, such as Pb and
Cu, while clay minerals, such as kaolinite, can adsorb and retain PTEs in soils [55].

Figure 2. Mineral phases identified on the studied soils and local background samples. Quartz was
not considered.

Table 2. Descriptive statistics of studied soil and local background samples (in mg/kg).

Soils (n = 71) Local Background (n = 10)


Var
Min Max Mean SD Min Max Mean SD
Al 6399 24,510 11,453 3974 19,101 32,200 22,406 4061
Ba 64 287 117 41 72 134 106 23
Br 0.5 12.1 3.7 2.4 1.0 3.5 2.3 0.8
Ca 486 31,240 9167 8630 243 843 471 217
Cr 10.3 238.0 41.3 41.1 20.0 59.4 33.5 11.5
Cu 3.3 1470 59.7 189 6.8 14.0 9.2 2.1
Fe 1364 27,839 4655 4204 5316 9204 6801 1255
K 3512 10,601 6278 1719 6409 14,129 9001 2263
Mg 10 3 028 979 530 609 965 675 104
Mn 23 310 83 57 95 194 140 34
Na 386 3769 1091 780 519 1803 974 343
Ni 0.5 16.1 3.4 3.3 2.8 5.2 3.9 0.7
P 79 3 029 686 677 127 284 197 53
Pb 6.4 506 30.2 59.6 6.6 9.5 7.8 0.8
Geosciences 2022, 12, 290 7 of 15

Table 2. Cont.

Soils (n = 71) Local Background (n = 10)


Var
Min Max Mean SD Min Max Mean SD
Rb 8.0 28.8 13.1 3.4 13.9 23.8 18.0 3.8
Si 363,290 451,075 430,505 18,882 398,497 433,284 422,573 13,835
Sn 0.5 110 7.0 12.9 0.5 4.9 3.4 1.6
S 50 3228 544 578 55 175 103 31
Sr 9 89 29.5 21.8 10.0 18.0 13.6 2.7
Ti 779 2812 1624 514 1121 3333 1754 696
V 3 18.4 7.3 3.3 9.6 19.9 13.6 3.3
Zn 2 1077 92 162 3 4 3 1
Zr 65 341 161 65 81 200 123 40
Var—variable; Bkg—local background; min—minimum; max—maximum; SD—standard deviation.

Principal component analysis (PCA) performed on the studied soil samples identified
four main components: (a) Group 1, with 61.2% of the total variance explained, grouping
Br, Ca, Mg, Mn, Ni, P, S, and Sr with positive loading and Si with negative loading;
(b) Group 2, explaining 12.4% of the total variance, with variables Al, Ba, Cr, K, Na, Rb,
and V; Group 3, explaining 8.0% of the total variance, with elements Cu, Fe, Pb, Sn, and Zn;
(d) Group 4, with 4.4% of the total variance grouping variables Ti and Zr.
Group 1 suggested the existence of two enrichment sources, the anthropogenic Hulene-
B dump with elements Br, Ca, Mg, Mn, Ni, P, S, and Sr and a natural source of Si, with
negative loading. The waste dump elemental contribution can be evidenced by higher mean
concentrations in studied samples when compared to local background mean contents
(Table 2). The high Si concentration may be associated with mineralogical phases identified,
with >98% quartz (SiO2 ) [32]. Group 2 variables suggested heterogeneous sources. High
Cr concentrations might be related to anthropogenic enrichment (e.g., dump and Julius
Nyerere Avenue degraded asphalt). Group 3 variables presented high concentrations of Cu,
Pb, Sn, and Zn when compared to local background samples, suggesting anthropogenic
sources (e.g., dump, Julius Nyerere Avenue degraded asphalt, and airport traffic flow). The
low Fe content can be linked to leaching of Fe oxides given the topographic nature of the
sampling area [24], promoting water circulation and infiltration. The low Ti concentration
(Group 4) could be associated with geogenic context and the Ti oxide leaching processes.
The higher Zr concentration (Group 4), when compared to local background samples,
suggested an anthropogenic source (dump).
To access elemental sources, a Spearman correlation analysis (Table 3) was applied
to elements Cr, Cu, Mn, Ni, Pb, and Zn, considered by several studies as PTEs [56]. The
analysis revealed statistically significant correlations between pairs Cr/Cu, Mn/Cr, Mn/Cu,
Ni/Cr, Ni/Cu, Ni/Mn, Pb/Cr, Pb/Cu, Pb/Mn, Pb/Ni, Zn/Cr, Zn/Cu, Zn/Mn, Zn/Ni,
Zn/Pb, and Zr/Cr (p < 0.01), suggesting a common source.

Table 3. Spearman correlation between the selected PTEs of the soil samples.

Cr Cu Mn Ni Pb Zn
Cu 0.640 ** 1
Mn 0.747 ** 0.709 ** 1
Ni 0.770 ** 0.699 ** 0.792 ** 1
Pb 0.626 ** 0.759 ** 0.669 ** 0.711 ** 1
Zn 0.695 ** 0.813 ** 0.791 ** 0.799 ** 0.734 ** 1
Zr 0.377 ** 0.14 0.218 0.152 0.175 0.155
** p < 0.01.

Cluster analysis was conducted on selected PTEs in order to assess Cr, Ni, Pb, Mn, Zn,
Cu, and Zr common sources. Two main variable groups were formed (Figure 3). Cluster 1
grouped variables Cr, Ni, Pb, Mn, Zr, and Zn, suggesting a common anthropogenic source
(Hulene-B waste dump), with elemental concentrations higher than local background
Geosciences 2022, 12, 290 8 of 15

samples (p < 0.05). Contaminants may be associated with the dispersion of leachates from
the dump, ash from burnt waste, degraded pavements and traffic-related materials from
Julius Nyerere Avenue, and the international airport (west of the dump). Cr pollution has
been associated with paints, varnishes, organic solvents, oils, and vehicle brakes [57,58],
whereas Ni pollution has been associated electronic wastes deposited on the dump [59].
Zr and Mn contamination could be associated with paints, cosmetics, pharmaceuticals,
textiles, electrical waste, traffic associated emissions, and electronic equipment [60], while
Pb could be associated with electronic wastes such as rechargeable batteries, paint cans,
varnishes, organic solvents, glass waste, fuel combustion, and air transport [61]. Cluster 2
grouped elements Cu and Zn, characterized by higher concentrations than local background
samples, suggesting a common origin (the dump site). These PTEs were linked to electronic
waste, paints, bottle caps, tire wear, and asphalt degradation. Studies by [62] found high
concentrations of heavy metals in soils around landfills and considered an association with
leachates. Furthermore, the authors of [63,64] suggested that high levels of Pb, Zn, Hg, and
Cd in soils around a dumpsite can be associated with ash deposition in soils, resulting from
contaminated waste burning.

Figure 3. Hierarchical cluster analysis of the selected PTEs.

3.2. Environmental Risk Assessment


The PTEs Cr, Cu, Mn, Ni, Pb, Zn, and Zr were selected to estimate the impact on the
environment using the potential ecological risk index (PERI), pollution load index (PLI)
and soil Nemerow index (PN) (Figure 4). PLI results showed that 15.5% and 2.8% of all
samples were classified with high and very high contamination, respectively. A greater rep-
resentative number of samples (56.3%) were classified with moderate contamination, while
25.4% were classified with low contamination. The PLI individual element contribution was
ranked as Zn >> Cu > Pb > Cr = Zr > Ni > Mn. Spatially, the environmental contamination
levels varied significantly. PLI results < 1 were found in samples located at the central,
western, and northern boundaries of the dump. In turn, samples suggesting high levels
of environmental deterioration (PLI > 1), due to Zn >> Cu > Pb > Cr concentrations, were
located at the eastern boundary and at the immediate western boundary of the dump.
Two samples with extremely high levels of deterioration (PLI > 6) were located at the
southwestern (densely inhabited) boundary and at the immediate northwestern boundary
of the dump.
Geosciences 2022, 12, 290 9 of 15

Figure 4. Potential ecological risk index (PERI), pollution load index (PLI), and soil Nemerow index
calculated with local background soils and with world soils proposed by Reimann and Caritat (1998).

PERI results revealed that 59.2% of all samples presented a low ecological risk, while
22.5%, 11.3%, and 7.0% of the samples were classified with moderate, serious, and severe
potential ecological risk, respectively. The monomial potential ecological risk factor was
ranked as Zn >> Cu > Cr > Zr > Pb > Ni > Mn. The samples with PERI index < 150 were
concentrated in the center and north of the sampling area. Medium and high risk was
found to the immediate east of the dumpsite and to the east of the sampling area. Extremely
high risk (PERI > 600) was predominant in samples to the east, west, and southwest of the
dump (densely inhabited).
The soil Nemerow index was calculated using both local background (PNbkg ) and
world soils (PNRC ) proposed by [22]. The spatial distribution of (PNRC ) showed that only
four samples, on the eastern boundary of sampling area, were classified as nonpolluted.
The warning level of pollution was found in random samples in the center, east, west, and
north of study area. Samples with low to moderate levels of pollution were predominant
at the eastern and northern boundaries, while a high level of pollution was identified
in samples at the immediate east and west boundaries of the dump and in the extreme
southwest (densely inhabited). The PNbkg suggested that only three samples were classified
Geosciences 2022, 12, 290 10 of 15

with a low pollution level, whereas four samples were classified with a moderate pollution
level, located at the eastern boundary and in the northern direction of the dump.
The eastern and western boundaries of the dump were characterized by a high pollu-
tion level, suggesting a strong contribution of the dump to soil contamination. Pollutants
were ranked as Zn >> Cu > Pb > Cr = Zr > Ni > Mn for PNbkg and Cu > Pb > Zn > Cr > Zr
> Mn = Ni for PNRC . About 90.1% of studied samples were classified as seriously polluted,
with no samples in the safety or precaution domains. PNRC revealed 46.5% of the samples
to be slightly polluted and 23.9% to be seriously polluted. The variation between local
background and world soil reference indices evidenced a bias resulting from the differing
biogeochemical characteristics of the soils [22].
Dumps have been reported to be a source of enrichment of pollutants in surrounding
soils [20]. Variables Cr, Cu, Mn, Ni, Pb, Zn, and Zr found in the studied samples could be
associated with different anthropogenic sources, such as the dump wastes [65], surrounding
traffic and pavement degradation [42], resuspension [66], and airport proximity [67]. Alu-
minum has been associated with untreated disposal of cans, household utensils, cosmetics,
and laminated packaging [7], Cr has been associated with packaging waste from paints,
varnishes, organic solvents, and oils [56,57], Cu, Pb, Zn, and Ni have been associated with
waste electronic equipment, rechargeable, paint cans, varnishes, organic solvents, and glass
waste [58,59], and Zr has been associated with hospital wastes, explosives, and lamps [68],
all deposited in the Hulene-B dump.
The spatial distribution of indices (Figure 4) suggested very high levels in the imme-
diate eastern, western, and southwestern limits of the dump, with medium to low levels
predominant in the central and northern areas. The high contamination levels found in
the southwest area of the dump may be associated with three contamination mechanisms,
namely, incineration of wastes ash deposition, a recent sporadic waste disposal area that
contributes significantly to leachates, especially during rainy season, with mobile metals
such as Mn, Zn, Ni, and Zr [69], and the Maputo International Airport in the vicinity of
the study area (Figure 1). Contamination found on samples collected west of the study
area may be associated with leachate circulation originating from the dump, an important
source of enrichment of mainly Zn, Cu, Pb, and Cr, which are easily incorporated into the
leachate [6]. High levels of contamination on the eastern border may be associated with
leachate dispersion and intense traffic flow on Julius Nyerere Avenue (Figure 1), due to
degraded asphalt, a source of Pb, Cu, Mn, Ni, Zn, and Zr. Recently, the authors of [24]
suggested that the area around the Hulene-B dump is characterized by anomalous resistive
surfaces associated with heavy-metal contamination, e.g., Cr, Cu, Mn, Ni, Pb, Zn, and Zr,
as well as pollutants reported in leachates originating from landfills [70].

3.3. Human Health Risk Assessment


Areas with contaminated soils pose a health risk to the exposed population [6,71]. For
example, Cr chronic exposure can cause nasal irritation, bleeding, ulcers, convulsions, kid-
ney, and liver damage and even death [70]. The mean Cr concentration (41.3 mg/kg) found
in the studied soils was below maximum allowable soil (MAS) target proposed by [72]
(150 mg/kg). Nevertheless, a maximum of 238 mg/kg was found in a sample located in the
eastern end of the waste dump, near Avenida Julius Nyerere. The high concentration levels
may be associated with traffic-related waste, such as tires and brakes, as well as emissions
from Julius Nyerere Avenue traffic, including degraded road pavement. The toxic effects
of Cu chronic exposure include anemia and disorders of the central nervous system and
cardiovascular system. The average Cu concentration found in this study was 59.7 mg/kg,
below the MAS (150 mg/kg) [71], and a maximum of 1470 mg/kg was found in a sample
at the immediate northwestern boundary the dump. This area is characterized by intense
tire burning. Tires are considered a source of Cu enrichment. Zinc toxicity leads to loss of
appetite, dehydration, weakness, weight loss or gain, diarrhea, and jaundice [73]. The mean
content found in the studied samples was 92 mg/kg, below the MAS (250 mg/kg) [71];
however, a maximum of 1077 mg/kg was found in a sample at the immediate northwestern
Geosciences 2022, 12, 290 11 of 15

boundary the dump. Zinc enrichment may be associated with batteries, metal alloys, paint
cans, cosmetics, pharmaceuticals, textiles, and electrical and electronic equipment, accu-
mulated over time [59]. Nickel chronic exposure has been associated with gastrointestinal
and neurological effects including lung cancer [73]. The average concentration in studied
samples was 3.4 mg/kg, below the MAS (45 mg/kg) [71]. Ingestion or inhalation of Pb can
disturb almost all organs and the nervous system, in addition to causing kidney damage,
brain damage, miscarriage, and death [74]. The average concentration found in the studied
samples was 30.2 mg/kg, below the MAS (300 mg/kg) [72], with a maximum of 506 mg/kg
in a sample at the immediate northwestern boundary the dump. Mn is classified with low
toxicity [75]; however, chronic exposure to high levels can cause memory problems, halluci-
nations, Parkinson’s disease, pulmonary embolism and bronchitis, apathy, schizophrenia,
muscle weakness, headaches, and insomnia [76]. The average concentration found was 83
mg/kg, below the MAS (330 mg/kg) [71], and a maximum of 310 mg/kg was observed
in a sample at the immediate northwestern boundary the dump. Zr is classified with
low toxicity, but chronic exposure can cause respiratory tract irritation, dermatitis, and
pulmonary fibrosis, with a few cases reported in nonindustrial settings [77]. The limits of
its concentration in soils are still in debate [77].
Considering that the surroundings of the Hulene-B dumpsite are densely popu-
lated [27], and that the dumpsite is used daily by adult and children waste-pickers [26], the
noncarcinogenic hazard, systemic toxicity, and carcinogenic risk associated with selected
PTE concentrations were calculated. The hazard index (HI) for systemic toxicity (Figure 5),
as expected, was higher in children than adults, due to the lower body mass and higher
absorption factor in children [16,78]. Elements that most contribute to HI are Zr and Al.
A higher exposure of children to PTEs present in soils around the Hulene-B dump was
recently reported by [26], who found that children around the dump were collecting solid
wastes without using proper protection, thus increasing exposure.

Figure 5. Hazard index for systemic toxicity for children and adults, as well as adjusted for both
children and adults.

Carcinogenic risk, for both children and adults, only posed a significant risk for
Pb ingestion in samples P13 (6.19 ×−610−6 ) and P33 (2.94 ×
−6
10−6 ), remaining lower than
−4 − 4
1.00 × 10 , the target level. Sample P13 was collected at the immediate northwestern
boundary of the dump, ~20 m away from dwellings, where waste pickers are common,
and sample P33 was collected in a residential area, southwest of the dump, representing a


Geosciences 2022, 12, 290 12 of 15

risk especially for children, due to hand-to-mouth habits, associated with the fact that Pb
accumulates at the soil surface in the first 2–5 cm [60]. Studies by [79,80] described that Pb
is very toxic to children and pregnant women, who are more susceptible to adverse health
effects. The population around the Hulene-B dump is young and active (14–44 years),
representing ~65% of the inhabitants [27], who spend most of their time working on solid
waste recycling inside the dump, often without adequate protection [26]. In both samples,
total risk and ingestion risk were similar, with inhalation and dermal contact considered
negligible.

4. Conclusions
Hazard indices PLI, PN, and PERI ranged from low to moderate in samples collected
in the central west area of the dumpsite, whereas samples closer to the dumpsite and at
the southwestern edge showed a predominantly high ecological hazard. The Nemerow
hazard using local background mean concentrations (PNbkg ) revealed a predominantly high
hazard (90.1%) in the studied area, suggesting a high level of environmental deterioration.
The individual elemental contributions could be ranked as Zn >> Cu > Pb > Cr = Zr > Ni > Mn
for the pollution load index, as Zn >> Cu > Cr > Cr > Zr > Pb > Ni > Mn for PERI, and
as Zn >> Cu > Pb > Cr = Zr > Ni > Mn for PNbkg . The carcinogenic risk for both children
and adults was significant only due to Pb concentrations in the immediate northwest area
of the dump (6.19 × 10−6 ) and southwest of the densely inhabited dump (2.94 × 10−6 ).
Future studies are needed to understand the bioaccessibility of metals in soils, allowing
the identification of bioaccumulation standards in indicators such as biological absorption
coefficient for each element. Given the toxicity of Cr, Cu, Mn, Ni, Pb, Zn, and Zr, especially
for children, measures to control and mitigate soil contamination in the surroundings of
the Hulene-B dump are pertinent.

Author Contributions: Conceptualization, B.B., C.C. and F.R.; methodology, B.B. and C.C.; formal
analysis, C.C. and F.R.; investigation, B.B. and C.C.; writing—original draft preparation, B.B. and
C.C.; writing—review and editing, B.B., C.C. and F.R.; supervision, F.R. and C.C. All authors have
read and agreed to the published version of the manuscript.
Funding: This work was supported by the GeoBioTec (UIDB/04035/2020) Research Center, funded
by FEDER funds through the Operational Program Competitiveness Factors COMPETE and by
National funds through FCT. The first author acknowledges grants from the Portuguese Institute
Camões and FNI (Investigation National Fund—Mozambique).
Data Availability Statement: Not applicable.
Conflicts of Interest: The authors declare no conflict of interest.

References
1. Boente, C.; Baragaño, D.; García-González, N.; Forján, R.; Colina, A.; Gallego, J. A holistic methodology to study geochemical and
geomorphological control of the distribution of potentially toxic elements in soil. Catena 2022, 208, 105730. [CrossRef]
2. Morita, A.K.M.; Pelinson, N.D.S.; Wendland, E. Persistent impacts of an abandoned non-sanitary landfill in its surroundings.
Environ. Monit. Assess. 2020, 192, 463. [CrossRef] [PubMed]
3. Khan, S.; Anjum, R.; Raza, S.T.; Bazai, N.A.; Ihtisham, M. Technologies for municipal solid waste management: Current status,
challenges, and future perspectives. Chemosphere 2022, 288, 132403. [CrossRef]
4. Lashen, Z.M.; Shams, M.S.; El-Sheshtawy, H.S.; Slaný, M.; Antoniadis, V.; Yang, X.; Sharma, G.; Rinklebe, J.; Shaheen, S.M.;
Elmahdy, S.M. Remediation of Cd and Cu contaminated water and soil using novel nanomaterials derived from sugar beet
processing- and clay brick factory-solid wastes. J. Hazard. Mater. 2022, 428, 128205. [CrossRef]
5. Hoang, S.A.; Bolan, N.; Madhubashani, A.; Vithanage, M.; Perera, V.; Wijesekara, H.; Wang, H.; Srivastava, P.; Kirkham, M.;
Mickan, B.S.; et al. Treatment processes to eliminate potential environmental hazards and restore agronomic value of sewage
sludge: A review. Environ. Pollut. 2022, 293, 118564. [CrossRef]
6. Zaynab, M.; Al-Yahyai, R.; Ameen, A.; Sharif, Y.; Ali, L.; Fatima, M.; Khan, K.A.; Li, S. Health and environmental effects of heavy
metals. J. King Saud Univ. Sci. 2022, 34, 101653. [CrossRef]
7. Morita, A.K.; Ibelli-Bianco, C.; Anache, J.A.; Coutinho, J.V.; Pelinson, N.S.; Nobrega, J.; Rosalem, L.M.; Leite, C.M.; Niviadonski,
L.M.; Manastella, C.; et al. Pollution threat to water and soil quality by dumpsites and non-sanitary landfills in Brazil: A review.
Waste Manag. 2021, 131, 163–176. [CrossRef]
Geosciences 2022, 12, 290 13 of 15

8. Cheela, V.R.S.; Goel, S.; John, M.; Dubey, B. Characterization of municipal solid waste based on seasonal variations, source and
socio-economic aspects. Waste Dispos. Sustain. Energy 2021, 3, 275–288. [CrossRef]
9. Breus, D.; Yevtushenko, O. Modeling of Trace Elements and Heavy Metals Content in the Steppe Soils of Ukraine. J. Ecol. Eng.
2022, 23, 159–165. [CrossRef]
10. Singh, A.; Pal, D.B.; Mohammad, A.; Alhazmi, A.; Haque, S.; Yoon, T.; Srivastava, N.; Gupta, V.K. Biological remediation
technologies for dyes and heavy metals in wastewater treatment: New insight. Bioresour. Technol. 2022, 343, 126154. [CrossRef]
11. Kalisz, S.; Kibort, K.; Mioduska, J.; Lieder, M.; Małachowska, A. Waste management in the mining industry of metals ores, coal,
oil and natural gas—A review. J. Environ. Manag. 2022, 304, 114239. [CrossRef]
12. Meloni, F.; Montegrossi, G.; Lazzaroni, M.; Rappuoli, D.; Nisi, B.; Vaselli, O. Total and Leached Arsenic, Mercury and Antimony
in the Mining Waste Dumping Area of Abbadia San Salvatore (Mt. Amiata, Central Italy). Appl. Sci. 2021, 11, 7893. [CrossRef]
13. Li, H.; Sun, J.; Gui, H.; Xia, D.; Wang, Y. Physiochemical properties, heavy metal leaching characteristics and reutilization
evaluations of solid ashes from municipal solid waste incinerator plants. Waste Manag. 2021, 138, 49–58. [CrossRef]
14. Hussein, M.; Yoneda, K.; Mohd-Zaki, Z.; Amir, A.; Othman, N. Heavy metals in leachate, impacted soils and natural soils of
different landfills in Malaysia: An alarming threat. Chemosphere 2021, 267, 128874. [CrossRef]
15. Soumahoro, N.S.; Kouassi, N.L.B.; Yao, K.M.; Kwa-Koffi, E.K.; Kouassi, A.M.; Trokourey, A. Impact of municipal solid waste
dumpsites on trace metal contamination levels in the surrounding area: A case study in West Africa, Abidjan, Cote d’Ivoire.
Environ. Sci. Pollut. Res. 2021, 28, 30425–30435. [CrossRef] [PubMed]
16. Yap, C.K.; Chew, W.; Al-Mutairi, K.A.; Nulit, R.; Ibrahim, M.H.; Wong, K.W.; Bakhtiari, A.R.; Sharifinia, M.; Ismail, M.S.; Leong,
W.J.; et al. Assessments of the Ecological and Health Risks of Potentially Toxic Metals in the Topsoils of Different Land Uses: A
Case Study in Peninsular Malaysia. Biology 2022, 11, 2. [CrossRef]
17. Kimani, N. Implications of the Dandora Municipal Dumping Site in Nairobi, Kenya. Environ. Pollut. Impacts Public Health 2012, 1,
14.
18. Wang, S.; Han, Z.; Wang, J.; He, X.; Zhou, Z.; Hu, X. Environmental risk assessment and factors influencing heavy metal
concentrations in the soil of municipal solid waste landfills. Waste Manag. 2022, 139, 330–340. [CrossRef]
19. Essien, J.P.; Ikpe, D.I.; Inam, E.D.; Okon, A.O.; Ebong, G.A.; Benson, N.U. Occurrence and spatial distribution of heavy metals in
landfill leachates and impacted freshwater ecosystem: An environmental and human health threat. PLoS ONE 2022, 17, e0263279.
[CrossRef] [PubMed]
20. El Fadili, H.; Ben Ali, M.; Touach, N.; El Mahi, M.; Lotfi, E.M. Ecotoxicological and pre-remedial risk assessment of heavy metals
in municipal solid wastes dumpsite impacted soil in morocco. Environ. Nanotechnol. Monit. Manag. 2022, 17, 100640. [CrossRef]
21. Liao, J.; Cui, X.; Feng, H.; Yan, S. Environmental Background Values and Ecological Risk Assessment of Heavy Metals in
Watershed Sediments: A Comparison of Assessment Methods. Water 2022, 14, 51. [CrossRef]
22. Reimann, C.; Caritat, P. Chemical Elements in the Environment: Factsheets for the Geochemist and Environmental Scientist; Springer:
Berlin, Germany, 1998; ISBN 3540636706.
23. Serra, C. Da Problemática Ambiental à Mudança: Rumo à um Mundo Melhor; Editora Escolar: Maputo, Mozambique, 2012;
ISBN 0A9789896700300. (In Portuguese)
24. Bernardo, B.; Candeias, C.; Rocha, F. Application of Geophysics in geo-environmental diagnosis on the surroundings of the
Hulene-B waste dump, Maputo, Mozambique. J. Afr. Earth Sci. 2022, 185, 104415. [CrossRef]
25. Vicente, E.; Jermy, C.; Schreiner, H. Urban geology of Maputo, Mozambique. Geol. Soc. London 2006, 338, 1–13. Available online:
https://citeseerx.ist.psu.edu/viewdoc/download?doi=10.1.1.606.7220&rep=rep1&type=pdf (accessed on 10 February 2022).
26. Matsinhe, F.; Paulo, M. Estudo Etnográfico sobre os catadores de Lixo da Lixeira de Hulene (Maputo); Cadernos África Contemporânea
3; Universidade Eduardo Mondlane: Maputo, Moçambique, 2020; ISSN 2595-5713.
27. INE. Boletim de Estatísticas Demográficas e Sociais, Maputo Cidade 2019; Instituto Nacional de Estatistica: Maputo, Mozambique, 2020.
Available online: http://www.ine.gov.mz/estatisticas/estatisticas-demograficas-e-indicadores-sociais/boletim-de-indicadores-
demograficos-22-de-julho-de-2020.pdf/at_download/file. (accessed on 21 February 2022).
28. Palalane, J.; Segala, I. Urbanização e Desenvolvimento Municipal em Moçambique: Gestão de Resíduos Sólidos. Available on-
line: https://limpezapublica.com.br/urbanizacao-e-desenvolvimento-municipal-em-mocambique-capitulo-gestao-de-residuos-
solidos/ (accessed on 22 April 2022).
29. Ferrão, D. Evaluation of Removal and Disposal of Solid Waste in Maputo City, Mozambique. Master’s Thesis, University of Cape
Town, Cape Town, South Africa, 2006. Available online: https://open.uct.ac.za/bitstream/handle/11427/4851/thesis_sci_2006
_ferrao_d_a_g.pdf?sequence=1 (accessed on 10 February 2022).
30. WHO. Global Air Quality Guidelines Particulate matter (PM2.5 and PM10), Ozone, Nitrogen Dioxide, Sulfur Dioxide and Carbon
Monoxide. World Health Organization. 2020. Available online: https://apps.who.int/iris/bitstream/handle/10665/345334/97892
40034433-eng.pdf (accessed on 10 February 2022).
31. Afonso, R. A Geologia de Moçambique—Notícia Explicativa da Carta Geológica de Moçambique 1978, 1:2,000,000; Imprensa Nacional de
Moçambique: Maputo, Moçambique, 1978. (In Portuguese)
32. Momade, F.; Ferrara, M.; Oliveira, J. Notícia Explicativa da Carta Geológica 2532 Maputo (1:50,000); Direção Nacional de Geologia:
Maputo, Mozambique, 1996. (In Portuguese)
Geosciences 2022, 12, 290 14 of 15

33. Cendón, D.I.; Haldorsen, S.; Chen, J.; Hankin, S.; Nogueira, G.; Momade, F.; Achimo, M.; Muiuane, E.; Mugabe, J.; Stigter, T.Y.
Hydrogeochemical aquifer characterization and its implication for groundwater development in the Maputo district, Mozambique.
Quat. Int. 2020, 547, 113–126. [CrossRef]
34. Nhantumbo, A.B.; Cambule, A.H. Bulk density by Proctor test as a function of texture for agricultural soils in Maputo province of
Mozambique. Soil Tillage Res. 2006, 87, 231–239. [CrossRef]
35. CIAT. Climate-Smart Agriculture in Mozambique. Climate-Smart Agriculture in Mozambique, Center for Tropical Agriculture
2017. Available online: https://climateknowledgeportal.worldbank.org/sites/default/files/2019-06/CSA-in-Mozambique.pdf
(accessed on 13 April 2022).
36. Muchangos, A. Paisagens e Regiões Naturais de Moçambique; Editora Escolar: Maputo, Mozambique, 1999; Available online:
https://docplayer.com.br/47220681-Mocambique-paisagens-e-regioes-naturais.html (accessed on 30 April 2022).
37. Håkanson, L. An ecological risk index for aquatic pollution control. A sedimentological approach. Water Res. 1980, 14, 975–1001.
[CrossRef]
38. Candeias, C.; Da Silva, E.F.; Ávila, P.F.; Teixeira, J.P. Identifying Sources and Assessing Potential Risk of Exposure to Heavy Metals
and Hazardous Materials in Mining Areas: The Case Study of Panasqueira Mine (Central Portugal) as an Example. Geosciences
2014, 4, 240–268. [CrossRef]
39. Bowen, H.J.M. Trace Elements in Biochemistry; Academic Press: London, UK, 1966.
40. Candeias, C.; Ávila, P.F.; Alves, C.; Gama, C.; Sequeira, C.; da Silva, E.F.; Rocha, F. Dust Characterization and Its Potential Impact
during the 2014–2015 Fogo Volcano Eruption (Cape Verde). Minerals 2021, 11, 1275. [CrossRef]
41. Nazzal, Y.; Bărbulescu, A.; Howari, F.; Al-Taani, A.; Iqbal, J.; Xavier, C.; Sharma, M.; Dumitriu, C. Assessment of Metals
Concentrations in Soils of Abu Dhabi Emirate Using Pollution Indices and Multivariate Statistics. Toxics 2021, 9, 95. [CrossRef]
42. Candeias, C.; Ávila, P.F.; Sequeira, C.; Manuel, A.; Rocha, F. Potentially toxic elements dynamics in the soil rhizospheric-plant
system in the active volcano of Fogo (Cape Verde) and interactions with human health. Catena 2022, 209, 105843. [CrossRef]
43. Xiao, H.; Shahab, A.; Xi, B.; Chang, Q.; You, S.; Li, J.; Sun, X.; Huang, H.; Li, X. Heavy metal pollution, ecological risk, spatial
distribution, and source identification in sediments of the Lijiang River, China. Environ. Pollut. 2021, 269, 116189. [CrossRef]
44. Candeias, C.; Vicente, E.; Tomé, M.; Rocha, F.; Ávila, P.; Célia, A. Geochemical, Mineralogical and Morphological Characterisation
of Road Dust and Associated Health Risks. Int. J. Environ. Res. Public Health 2020, 17, 1563. [CrossRef]
45. Zhang, Y.; Wang, S.; Gao, Z.; Zhang, H.; Zhu, Z.; Jiang, B.; Liu, J.; Dong, H. Contamination characteristics, source analysis and
health risk assessment of heavy metals in the soil in Shi River Basin in China based on high density sampling. Ecotoxicol. Environ.
Saf. 2021, 227, 112926. [CrossRef] [PubMed]
46. Mallongi, A.; Astuti, R.D.P.; Amiruddin, R.; Hatta, M.; Rauf, A.U. Identification source and human health risk assessment of
potentially toxic metal in soil samples around karst watershed of Pangkajene, Indonesia. Environ. Nanotechnol. Monit. Manag.
2022, 17, 100634. [CrossRef]
47. USEPA. Risk Assessment Guidance for Superfund, Volume I: Human Health Evaluation Manual; EPA 540-1-89-002; U.S. Environmental
Protection Agency: Washington, DC, USA, 1989.
48. USEPA. Screening Levels (RSL) for Chemical Contaminants at Superfund Sites. U.S.; U.S. Environmental Protection Agency: Washing-
ton, DC, USA, 2013.
49. Berg, R. Human Exposure to Soil Contamination: A Qualitative and Quantitative Analysis towards Proposals for Human Toxicological
Intervention Values (Partly Revised Edition); Report No. 725201011; National Institute for Public Health and the Environment:
Leiderdorp, The Netherlands, 1994.
50. RAIS. The Risk Assessment Information System (RAIS); U.S. Department of Energy’s Oak Ridge Operations Office (ORO): Oak
Ridge, TN, USA, 2022. Available online: https://rais.ornl.gov/ (accessed on 10 May 2022).
51. Ghasera, K.M.; Rashid, S.A. Geochemical characteristics of two contrasting weathering profiles developed at high altitude, NE
Lesser Himalaya, India: Implications for controlling factors and mobility of elements. J. Earth Syst. Sci. 2022, 131, 5. [CrossRef]
52. Han, B.; Weatherley, A.J.; Mumford, K.; Bolan, N.; He, J.-Z.; Stevens, G.W.; Chen, D. Modification of naturally abundant resources
for remediation of potentially toxic elements: A review. J. Hazard. Mater. 2022, 421, 126755. [CrossRef]
53. Kabata-Pendias, A. Trace Elements in Soils and Plants, 3rd ed.; CRC Press: Boca Raton, FL, USA, 2000. [CrossRef]
54. Bai, B.; Nie, Q.; Zhang, Y.; Wang, X.; Hu, W. Cotransport of heavy metals and SiO2 particles at different temperatures by seepage.
J. Hydrol. 2021, 597, 125771. [CrossRef]
55. Srivastava, P.; Singh, B.; Angove, M. Competitive adsorption behavior of heavy metals on kaolinite. J. Colloid Interface Sci. 2005,
290, 28–38. [CrossRef]
56. Fan, P.; Lu, X.; Yu, B.; Fan, X.; Wang, L.; Lei, K.; Yang, Y.; Zuo, L.; Rinklebe, J. Spatial distribution, risk estimation and source
apportionment of potentially toxic metal(loid)s in resuspended megacity street dust. Environ. Int. 2022, 160, 107073. [CrossRef]
[PubMed]
57. Al-Salem, S.M. Soil quality of simulated landfill exposure to plastics in context of heavy metal analysis. Environ. Sci. Pollut. Res.
2021, 28, 36904–36910. [CrossRef]
58. Safonov, A.; Popova, N.; Andrushenko, N.; Boldyrev, K.; Yushin, N.; Zinicovscaia, I. Investigation of materials for reactive
permeable barrier in removing cadmium and chromium(VI) from aquifer near a solid domestic waste landfill. Environ. Sci. Pollut.
Res. 2021, 28, 4645–4659. [CrossRef] [PubMed]
Geosciences 2022, 12, 290 15 of 15

59. Fiala, M.; Hwang, H.-M. Influence of Highway Pavement on Metals in Road Dust: A Case Study in Houston, Texas. Water Air Soil
Pollut. 2021, 232, 185. [CrossRef]
60. Islamd, S.; Idris, A.M.; Islam, A.R.M.T.; Phoungthong, K.; Ali, M.M.; Kabir, H. Geochemical variation and contamination level of
potentially toxic elements in land-uses urban soils. Int. J. Environ. Anal. Chem. 2021, 1–18. [CrossRef]
61. Alghamdi, A.G.; Aly, A.A.; Ibrahim, H.M. Assessing the environmental impacts of municipal solid waste landfill leachate on
groundwater and soil contamination in western Saudi Arabia. Arab. J. Geosci. 2021, 14, 350. [CrossRef]
62. Gujre, N.; Mitra, S.; Soni, A.; Agnihotri, R.; Rangan, L.; Rene, E.R.; Sharma, M.P. Speciation, contamination, ecological and human
health risks assessment of heavy metals in soils dumped with municipal solid wastes. Chemosphere 2021, 262, 128013. [CrossRef]
63. Obiri-Nyarko, F.; Duah, A.A.; Karikari, A.Y.; Agyekum, W.A.; Manu, E.; Tagoe, R. Assessment of heavy metal contamination
in soils at the Kpone landfill site, Ghana: Implication for ecological and health risk assessment. Chemosphere 2021, 282, 131007.
[CrossRef] [PubMed]
64. Chen, H.; Wang, L.; Hu, B.; Xu, J.; Liu, X. Potential driving forces and probabilistic health risks of heavy metal accumulation in
the soils from an e-waste area, southeast China. Chemosphere 2022, 289, 133182. [CrossRef] [PubMed]
65. Amiri, H.; Daneshvar, E.; Azadi, S. Contamination level and risk assessment of heavy metals in the topsoil around cement factory:
A case study. Environ. Eng. Res. 2021, 27, 210313. [CrossRef]
66. Brtnický, M.; Pecina, V.; Baltazár, T.; Galiová, M.V.; Baláková, L.; B˛eś, A.; Radziemska, M. Environmental Impact Assessment of
Potentially Toxic Elements in Soils Near the Runway at the International Airport in Central Europe. Sustainability 2020, 12, 7224.
[CrossRef]
67. Perks, C.; Mudd, G.M.; Currell, M. Using corporate sustainability reporting to assess the environmental footprint of titanium and
zirconium mining. Extr. Ind. Soc. 2021, 9, 101034. [CrossRef]
68. Gautam, P.; Kumar, S. Characterisation of Hazardous Waste Landfill Leachate and its Reliance on Landfill Age and Seasonal
Variation: A Statistical Approach. J. Environ. Chem. Eng. 2021, 9, 105496. [CrossRef]
69. Aydi, A.; Mhimdi, A.; Hamdi, I.; Touaylia, S.; Sdiri, A. Application of electrical resistivity tomography and hydro-chemical
analysis for an integrated environmental assessment. Environ. Nanotechnol. Monit. Manag. 2020, 14, 100351. [CrossRef]
70. Aghili, S.; Vaezihir, A.; Hosseinzadeh, M. Distribution and modeling of heavy metal pollution in the sediment and water mediums
of Pakhir River, at the downstream of Sungun mine tailing dump, Iran. Environ. Earth Sci. 2018, 77, 128. [CrossRef]
71. Ceballos, E.; Dubny, S.; Othax, N.; Zabala, M.E.; Peluso, F. Assessment of Human Health Risk of Chromium and Nitrate Pollution
in Groundwater and Soil of the Matanza-Riachuelo River Basin, Argentina. Expo. Health 2021, 13, 323–336. [CrossRef]
72. EU. Heavy Metals and Organic Compounds from Wastes Used as Organic Fertilisers; Final Report for ENV. A. 2./ETU/2001/0024.
2004. Available online: https://ec.europa.eu/environment/pdf/waste/compost/hm_finalreport.pdf (accessed on 15 January
2022).
73. Stuckey, J.W.; Neaman, A.; Verdejo, J.; Navarro-Villarroel, C.; Peñaloza, P.; Dovletyarova, E.A. Zinc Alleviates Copper Toxicity to
Lettuce and Oat in Copper-Contaminated Soils. J. Soil Sci. Plant Nutr. 2021, 21, 1229–1235. [CrossRef]
74. Naveed, M.; Bukhari, S.; Mustafa, A.; Ditta, A.; Alamri, S.; El-Esawi, M.; Rafique, M.; Ashraf, S.; Siddiqui, M. Mitigation of Nickel
Toxicity and Growth Promotion in Sesame through the Application of a Bacterial Endophyte and Zeolite in Nickel Contaminated
Soil. Int. J. Environ. Res. Public Health 2020, 17, 8859. [CrossRef] [PubMed]
75. Sattar, S.; Jehan, S.; Siddiqui, S. Potentially toxic metals in the petroleum waste contaminated soils lead to human and ecological
risks in Potwar and Kohat Plateau, Pakistan: Application of multistatistical approaches. Environ. Technol. Innov. 2021, 22, 101395.
[CrossRef]
76. Li, J.; Smith, R.L.; Xu, S.; Yang, J.; Zhang, K.; Shen, F. Manganese oxide as an alternative to vanadium-based catalysts for effective
conversion of glucose to formic acid in water. Green Chem. 2022, 24, 315–324. [CrossRef]
77. Stewart, C.; Damby, D.E.; Horwell, C.J.; Elias, T.; Ilyinskaya, E.; Tomašek, I.; Longo, B.M.; Schmidt, A.; Carlsen, H.K.; Mason, E.;
et al. Volcanic air pollution and human health: Recent advances and future directions. Bull. Volcanol. 2022, 84, 11. [CrossRef]
78. Jones, J.V.; Piatak, N.M.; Bedinger, G.M. Zirconium and hafnium, Chap. V. In Critical Mineral Resources of the United States
Economic and Environ Mental Geology and Prospects for Future Supply; Schulz, K.J., DeYoung, J.H., Jr., Seal, R.R., Bradley, D.C.,
Eds.; U.S. Geological Survey Professional: St. Petersburg, FL, USA, 2017; p. 26. Available online: https://doi.org/https:
//doi.org/10.3133/pp1802V (accessed on 21 February 2022).
79. Lian, Z.; Zhao, X.; Gu, X.; Li, X.; Luan, M.; Yu, M. Presence, sources, and risk assessment of heavy metals in the upland soils of
northern China using Monte Carlo simulation. Ecotoxicol. Environ. Saf. 2022, 230, 113154. [CrossRef] [PubMed]
80. Wixson, B.G.; Davies, B.E. Lead in Soil; CRC Press: Boca Raton, FL, USA, 2017. [CrossRef]
326
327

13.2.6 Appendix 7. Paper 6


328
environments

Article
The Contribution of the Hulene-B Waste Dump (Maputo,
Mozambique) to the Contamination of Rhizosphere Soils,
Edible Plants, Stream Waters, and Groundwaters
Bernardino Bernardo 1,2 , Carla Candeias 1, * and Fernando Rocha 1

1 GeoBioTec Research Centre, Department of Geosciences, University of Aveiro, 3810-193 Aveiro, Portugal
2 Faculty of Earth Sciences and Environment, Pedagogical University of Maputo, Maputo 2482, Mozambique
* Correspondence: candeias@ua.pt

Abstract: The contamination of ecosystems in areas around waste dumps is a major threat to the
health of surrounding populations. The aim of this study is to understand the contribution of the
Hulene-B waste dump (Maputo, Mozambique) to the contamination of edible plants, rhizosphere
soils, stream waters, and groundwater, and to assess human health risk. Soil and plant samples were
analyzed by XRD and XRF for mineralogical and chemical composition characterization, respectively.
Mineral phases identified in rhizosphere soil samples were ranked, calcite (CaCO3 ) > quartz (SiO2 ) >
phyllosilicates (micas and kaolinite) > anhydrite (CaSO4 ) > K feldspar (KAlSi3 O8 ) > opal (SiO2 ·nH2 O)
> gypsum (CaSO4 ·2H2 O), suggesting potential toxic elements low mobility. Soil environmental
indices showed pollution by Pb > Cu > Zn > Zr. The chemical composition of edible plants revealed
contamination by Ni, Cr, Mn, Fe, Ti, and Zr. Groundwaters and stream waters showed a potential
health risk by Hg and, in one irrigation water sample, by Pb content. The health hazard index
of rhizosphere soils was higher by ingestion, with children being the ones more exposed. Results
suggested a combined health risk by exposure to edible plants, rhizosphere soils, stream waters,
and groundwaters.

Keywords: edible plants; rhizosphere soil; water; risk assessment; Maputo


Citation: Bernardo, B.; Candeias, C.;
Rocha, F. The Contribution of the
Hulene-B Waste Dump (Maputo,
Mozambique) to the Contamination
1. Introduction
of Rhizosphere Soils, Edible Plants,
Stream Waters, and Groundwaters. Unplanned city development in developing countries together with demographic
Environments 2023, 10, 45. https://doi. growth increases solid urban wastes production, often deposited in abandoned areas,
org/10.3390/environments10030045 compromising the environment and health of surrounding populations [1]. Groundwater
is an important source of potable water, representing ~30% of the world’s resource for
Academic Editor: Manuel
drinking and domestics purposes [2], and stream water is an important source for irrigation
Duarte Pinheiro
purposes [3]. Stream waters and groundwaters in areas surrounding waste dumps are vul-
Received: 19 January 2023 nerable to contamination [4]. Waste dumps are a source of leachates, as a result of deposited
Revised: 20 February 2023 solid waste decomposition, compromising surrounding waters and soils. Hussein et al. [5]
Accepted: 2 March 2023 reported Hg, Cd, Pb, Mg, As, Zn, and Cu groundwater enrichment in landfills in Malaysia,
Published: 6 March 2023 posing a risk to human health. Parvin et al. [6] suggested that Pb-contaminated water
consumption can result in a range of hematological, neurological, renal, gene mutation,
central nervous system, and reproductive impairment outcomes. Chronic exposure to
Copyright: © 2023 by the authors.
water with high Zn content can induce stomach cramps, nausea, vomiting, anemia, and
Licensee MDPI, Basel, Switzerland.
damage to the pancreas [7]. Other health hazards reported in the surroundings of waste
This article is an open access article dumps were linked to contaminated agricultural products [8].
distributed under the terms and Romero et al. [9] showed that agricultural areas surrounding waste dumps were
conditions of the Creative Commons influenced by contaminants transported by the wind [10]. High concentrations of heavy
Attribution (CC BY) license (https:// metals in rhizosphere soils were identified as a potential health risk by inhalation, ingestion,
creativecommons.org/licenses/by/ and dermal contact, with farmers being the ones more exposed [11]. Additionally, the
4.0/). contamination of rhizosphere soils, i.e., soil immediately surrounding the plant root, can be

Environments 2023, 10, 45. https://doi.org/10.3390/environments10030045 https://www.mdpi.com/journal/environments


Environments 2023, 10, 45 2 of 19

transported to edible plants, through transfer mechanisms of absorbed potentially toxic


elements (PTEs) [12]. Sharma et al. [13] suggested that food crops from contaminated sites
can represent a hazard at the trophic level in the food chain, with increased bioaccumulation
of different elements. Contaminated plants can trigger several diseases in consumers,
depending on PTEs content and their bioaccessibility [14]. Zhou et al. [15] suggested that
pregnant women and children that consumed contaminated plants (e.g., Pb, Cd, Cu, Zn,
and As) were at greater risk.
In Mozambique, among the causes of urban ecosystems contamination are waste
dumps [16], with Maputo, the capital, having the largest areas of uncontrolled solid waste
disposal [17]. The production of municipal solid waste increased rapidly, with a daily
waste deposition in the Hulene-B dump of ~1000 t [18]. All types of domestic, urban,
clinical/hospital, and industrial wastes are deposited without previous treatment or con-
trol [19]. Cendón et al. [20] and Nogueira et al. [21] reported that the Maputo aquifer system
is vulnerable to contamination given the shallow groundwater depth. Bernardo et al. [22],
reported a possible migration of leachates causing contamination of surrounding soils,
stream waters, and groundwaters. Vicente et al. [23], and Bernardo et al. [24] suggested
enrichment of soils and groundwater by Cr, Cu, Mn, Pb, Zn, and Zr in the vicinity of the
Hulene-B dump. Studies on the contamination of edible plants, cultivated around dump
studies, are scarce, although the practice of horticulture is common in local houses back-
yards. The aim of this study is to understand the contribution of the Hulene-B waste dump
to the contamination of stream waters, groundwaters, edible plants, and rhizosphere soils
and to assess risk to human health. Results can contribute to understand the contamination
mechanisms of the ecosystem around the dump, which can help local governments to take
adequate measures to mitigate the impacts of the Hulene-B dump. The present study uses
recent methods and an integrated approach to the systemic study of environmental impact
and health risk posed by waste dumps on the surrounding population.

2. Methods and Materials


2.1. Study Area
Hulene-B waste dump (Figure 1) is located in the northern area of Maputo (Mozam-
bique) and occupies an area of ~17 hectares [18]. Around the dump are located Hulene-B
and Laulane neighborhoods, with 48,717 and 27,061 inhabitants, respectively [25]. Matsinhe
et al. [26], Sarmento [27], and Serra [18] reported that the dump receives all type of wastes
produced in the city, e.g., industrial, domestic, commercial, and hospital, with an estimated
~1000 t/d being continuously compressed. The dump is located in an abandoned open-air
quarry, and the first waste deposit dated from 1973 [28], having a waste height of ~6 to
15 m [29]. The dump is positioned on a dune slope with E–W orientation, with stream
water and leachates migrating in the same direction, mostly during rainy season [17]. The
dump’s W area was improved in 2018, after a large waste mass collapse that buried several
houses and caused the death of 17 inhabitants [30]. The improvement works consisted on
the removal of all houses near the dumpsite, creating an empty area. From 2020, this area
began to be occupied by crop fields around the riparian area, which accumulates surface
water, including dispersed leachates from the dumpsite and runoff water from Julius Ny-
erere Avenue (Figure 1). This wetland surrounding Hulene-B, is used for the practice of
subsistence agriculture, predominantly cabbage, lettuce, sweet potato, and pumpkin [16].
Environments 2023, 10, 45 3 of 19

Figure 1. Sampling locations, and surrounding environment, with numerous dwellings surrounding
Hulene-B waste dump (center), airport (NW), and Julius Nyerere Avenue (E).

Hulene-B hydrogeological system is part of the Tertiary to Quaternary aquifer sys-


tem [20], with a substrate of clayey marl to grey clay layers [20,21]. In the western limit of
Hulene-B dump, there is a semi-impermeable layer of clayey sands, fine-to-coarse sand,
and sandstones, facilitating water circulation between these two sectors [21]. Well water
levels, in the surroundings of the waste dump, varies between 1.5 and 9.3 m in deep [20].
Until 1990s, population use of groundwater was common in the surrounding of the dump;
however, many wells were decommissioned in recent years due to high levels of ground-
water contamination [31]. Soils on the western border of the dump are mostly composed
of materials from upper Pleistocene Malhazine Formation (QMa) and upper Pleistocene
to lower Pleistocene Red Point Formation (TPv) [32]. The Malhazine Formation, on the
western border, consists of coarse-to-fine, poorly consolidated sands with whitish to red-
dish colors, fixed by vegetation and consolidation processes. The predominant climate is
subtropical, with two seasons: (a) hot and rainy period from December to March, with
>60% of annual precipitation, with intense precipitation in January (average 125 mm);
(b) dry and cold season from April to September, with lower temperatures in June and July,
and weak and irregular precipitation with minimum intensity in August (~12 mm). The
average annual precipitation is ~789 mm, and prevailing SE winds [33].

2.2. Rhizosphere Sampling and Analysis


Four rhizosphere composite soils samples (R1 to R4; ~1.5 kg each) were collected
on agricultural areas on the W of the dump (Figure 1). Samples were georeferenced and
preserved in clean plastic bags until laboratory treatment. In the laboratory, samples were
dried in an oven ≤40 ◦ C, homogenized, and sieved (<2 mm), at Pedagogical University of
Maputo, Mozambique.≤ Afterwards, samples were transported to GeoBioTec research center
laboratories (University of Aveiro, Portugal, UAVR), for physical, chemical, and mineralog-
ical analyses. Electrical conductivity (EC) and pH were determined in a 1:2.5 soil/water
Environments 2023, 10, 45 4 of 19

solution using a high-resolution pH meter and condutivimeter. Organic matter (OM)


content was determined by the method described by USDA [34]. Color was determined
according to [35]. Soil pH was classified according to USDA [36] as extremely acidic 3.5–4.4,
very strongly acidic 4.5–5.0, strongly acidic 5.1–5.5, moderately acidic 5.6–6.0, slightly acidic
6.1–6.5, neutral 6.6–7.3, slightly alkaline 7.4–7.8, moderately alkaline 7.9–8.4, and strongly
≤ alkaline
μ 8.5–9.0. Soil organic matter was classified as:
μ high content > 4%, medium content 2–
4%, and low content < 2% [34]. Electrical conductivity was classified taking in consideration
reference values for soil salinity for horticultural crops (≤1100 µS/cm) and for sandy soils
α
EC (<100 µS/cm) [37]. Chemical composition was accessed by X-ray Fluorescence (XRF),
using a PANalytical PW 4400/40 45 Axios (Malvern Panalytical, Almelo, The Netherlands)
with Cr Kα radiation. Mineralogy was determined by powder X-ray Diffraction (XRD),
using a Phillips/Panalytical powder diffractometer (Malvern Panalytical, Almelo, The
Netherlands), θ model X’Pert-Pro MPD, equipped with an automatic slit. A Cu-X-ray tube
was operated at 50 Kv and 30 mA. Data were collected from 2 to 70◦ 2θ with a step size of
1◦ and a counting interval of 0.02 s. Results were within the 95% confidence limit, with pre-
cision and accuracy of analyses and digestion procedures monitored by internal standards,
certified reference material, and quality control blanks. Relative Standard Deviation was
between 5% and 10%.

2.3. Edible Plants Sampling and Analysis


Four cultivated edible plants (Figure 2a–c) were collected at the same rhizosphere
soil locations (Figure 1). The species collected were sample RP1: Amaranthus leaves
(Amaranthus spinosus), at the western end of the riparian zone; sample RP2: sweet potato
leaves (Ipomoea batatas), at the eastern end of the riparian zone; sample RP3: cabbage leaves
(Brassica oleracea L), at south-western end of the dump; and sample RP4: pumpkin leaves
(Cucurbita pepo), at the south-eastern end. Plant leaves are consumed by the inhabitants,
having been planted near the riparian zone due to soil humidity and proximity of stream
water used for irrigation. After plants collection (~400 g of fresh leaves), samples were
transported to the laboratory and dried in an oven at ~40 ◦ C. Chemical analyses of the
plants were performed by XRF, at GeoBioTec (UAVR) laboratories.

Figure 2. Edible plants cultivation near Hulene-B waste dump.


Figure 2. Edible plants cultivation near Hulene-B waste dump.

2.4. Water Sampling and Analysis


Four representative water samples were collected (Figure 1): 2 stream water samples
used for irrigation, in north-western (IWN) and south (IWS) areas; and 2 groundwater
samples from wells used for human consumption, located south-west (WS) and north-west
(WN) of the dump. Sample WS was collected at 5 m depth, on an uncovered well, and
sample WN, was collected at 8 m depth on a covered well. Each sample was collected and
preserved (4 ◦ C) in clean 1 L polyethylene bottles and transported to the laboratory for
Environments 2023, 10, 45 5 of 19

analysis. Waters pH was measured in situ, being classified as acidic < 7, neutral = 7, or
basic > 7 [38]. Water samples chemical analysis was performed using a Metalyser H1000
(Trace2o, Berkshire, United Kingdom) Trace-2 to analyze Pb, Hg, Zn concentrations. This
equipment is recommended in the analysis of heavy metals given its robustness, accuracy
and a 97% correlation with Atomic Absorption Spectrometry (GF-AAS) analysis [39].

2.5. Soil Indexes and Health Risk Assessment (SHRA)


Soil pollution index (PI) was calculated for PTEs Cr, Cu, Fe, Mn, Ni, Pb, Ti, Zn, and Zr,
by dividing the individual mean concentration (Ci ) by their corresponding reference values
for rhizosphere soils (Cd) − PI = Ci /Cd . Pollution index classifies soils as slightly polluted
(0 ≤ PI < 1), moderately polluted (1 ≤ PI < 3), highly polluted (3 ≤ PI < 6), and very highly
polluted (PI ≥ 6) [40].
Pollution Load index (PLI) provides a simple comparative mean to assess soils level of
PTEs contamination, by the same PTEs, as the nth root of the product of the nPI, with n
being the number of variables (PTEs) considered: PLI = (PI1 × PI2 × PI3 ×· · · × PIn )1/n .
When PLI > 1, environmental deterioration must be considered [41].
Soil health risk assessment (SHRA) was calculated taking in consideration inhabitants’
exposure, for both children and adults, through ingestion, inhalation, and dermal con-
tact [42]. Parameters considered were: ingestion rate of 200 and 100 mg/d for children and
adults, respectively; inhalation rate of 7.6 and 20 m3 /d for children and adults, respectively;
exposure frequency of 180 d/year; exposure period (EP) of 15 years for non-carcinogenic
effects in children and adults, respectively; 60 years as a lifetime (LT) for carcinogenic
effects; mean body weight of 15 and 60 kg for children and adults, respectively; mean time
to non-carcinogenic effects = EP × 365 days; exposed skin area of 2800 and 5700 cm2 for
children and adults, respectively, and skin adhesion factor of 0.2 and 0.07 mg/cm2 for chil-
dren and adults, respectively; dermal absorption factor of 0.001 for selected PTE (Cr, Cu, Fe,
Mn, Ni, Pb, Ti, Zn, and Zr), and age-adjusted dermal soil contact factor of 362 mg·yr/kg·d;
mean time to carcinogenic effects = LT × 365 days; and exposure time of 8 h/d [43]. Car-
cinogenic and non-carcinogenic side effects for each element were computed individually,
considering reference toxicity levels for each variable, as extensively described [42,44]. For
each selected PTE and pathway, non-cancer toxic hazard was estimated by computing
the Hazard Quotient (HQ) for systemic toxicity (i.e., non-carcinogenic risk). If HQ > 1,
non-carcinogenic effects might occur as the exposure concentration exceeds the reference
dose (RfD). The cumulative non-carcinogenic hazard index (HI) corresponded to the sum
of HQ for each pathway and/or variable. Similarly, HI > 1 indicates that non-carcinogenic
effects might occur. Carcinogenic risk, or the probability of an individual to develop any
type of cancer over a lifetime because of exposure to a potential carcinogen, was estimated
by the sum of total cancer risk for the three exposure routes. A risk > 1.00 × 10−6 is the
carcinogenic target, while values > 1.00 × 10−4 are considered unacceptable [43].

2.6. Soil/Plant Transfer Factor (TF)


Transfer factor allowed us to assess transfer of elements from soils to edible plants
and was calculated as follows: TF = Cplant /Csoil , where Cplant and Csoil are the individual
element concentration (mg/kg) in edible plants and soils samples, respectively. TF > 1
indicates an accumulation of elements [44].

2.7. Plant Health Risk Assessment (PHRA)


Plant health risk assessment (PHRA) is used to evaluate human health risk associated
with the consumption of contaminated plants [45]. Was estimated through the daily
elemental intake (DIM), the hazard risk index (HRI), the targeted hazard quotient (THQ),
and the hazard index (HIplant ). Plants DIM by adults does not consider body’s metabolic
response to PTEs; however, it may indicate a possible rate of intake of a selected PTE,
with DIM = Celement × Dintake /Bweight , where DIM is the daily adult elemental ingestion
(kg/day), Celement is the plant dry weight elemental concentration (mg/kg), Dintake is the
Environments 2023, 10, 45 6 of 19

daily consumption of edible plants (Amaranthus leaves = sweet potato leaves = pumpkin
leaves = cabbage = 0.025 kg/day), and Bweight is the average body weight (60 kg/adult) [45].
HRI by ingestion of edible plants for each element was calculated by HRI = DIM/RfD,
where RfD is the reference oral dose (Cr = 0.003; Cu = 0.037; Fe = 18; Mn = 0.14; Ni = 0.2;
Pb = 0.014; Ti = 0.7; Zn = 0.3; and Zr = 4.2 kg/day) [46]. Targeted hazard quotient was
the ratio between PTEs dose and a reference dose (dimensionless) indicating the non-
carcinogenic hazard through edible plants ingestion. It was calculated by THQ = (EF
× ED × Dintake × Celement )/(RfD × Bweigh t × AT), where EF is the exposure frequency
(300 days/year), ED is the exposure duration (40 years), and AT is the average exposure
time (ED × 365 days/year). Hazard index (HIplant ) is the summation of all PTEs considered
and given by HI = ∑in=1 THQn, i = n, . . . n [47]. HRI, THQ, and HI values <1.0 indicate
no risk by plant consumption, and if ≥1.0 adverse health effects should be considered [45].

2.8. Water Pollution Index (WPI)


The water pollution index was determined based on the measured concentrations of
Hg, Pb, and Zn. Single factor pollution index was calculated as: Pi = Ci /Si [48], where Pi is
the pollution index of pollution indicator i, Ci is the concentration of pollution indicator in
water (mg/L), and Si is the permissible limit for the pollution indicator in water, with Pi
classified as: <0.4, non-pollution; 0.4–1.0, slightly polluted; 1.0–2.0, moderately polluted;
2.0–5.0, heavily polluted; and >5.0, seriously polluted [49].

2.9. Water Health Risk Assessment (WHRA)


Water health risk assessment was determined for Zn, Hg and Pb content [50,51].
Exposure was determined with reference values established by WHO [38,52], of Hg 2, Zn
3.000, and Pb 1 µg/L. Adjustable parameters used were daily water consumption of 1.5
and 2 L/day for children and adults, respectively; exposure frequency number of days
exposed in a year of 365; total years of exposure of 15 and 60 for children and adults,
respectively; average exposure time 5475 and 21,900 for children and adults, respectively;
dermal permeability coefficient of 0.001 for all PTEs; surface area exposed of 852.5 and
1610 cm2 for children and adults, respectively; skin adherence factor of 0.2 and 0.07 for
children and adults, respectively. Carcinogenic risk and non-carcinogenic hazard were
determined using similar methods used for SHRA [53].

3. Results and Discussion


3.1. Soil Analysis
Rhizosphere soils pH, electrical conductivity (EC), organic matter (OM), and color are
presented in Table 1. Soil pH is an important property which controls chemical reactions
of metals [47]. Rhizosphere soils revealed a moderately alkaline pH (8.1–8.4), confirming
the findings of Naveen et al. [54] that, in general, areas impacted by heavy metals from
dumps tend to have higher pH. Gajaje et al. [55], in their study on soils around Morupule
dumpsite (Botswana), found an average pH of 8.3 in rhizosphere soils and associated it
with heavy metal contamination (Cu, Cr, Mn, and Pb). Skrbic et al. [56], suggested that the
PTEs highest retention in rhizosphere soils occurred with high pH, due to lower metals
solubility in this pH range. Alkaline pH might affect plant growth because of the difficulty
in absorbing essential plant nutrients, as well as PTEs, e.g., Fe, Mn, Cu, and Zn [36].
Environments 2023, 10, 45 7 of 19

Table 1. Rhizosphere soils pH, electrical conductivity (EC; in µS/cm), organic matter (OM; in %),
and color.

ID pH EC OM Color
R1 8.1 510 3.6 blackish
R2 8.4 571 1.7 greyish
R3 8.2 310 0.98 greyish
R4 8.4 810 2.3 greyish

EC showed high values, when compared to sandy soils guidelines (≤100 µS/cm; Lund,
2008), being ranked R4 >> R2 > R1 > R3. Bhardwaj et al. [39] suggested that rhizosphere
soils in areas near waste dumps, usually exhibit higher EC values, being associated with
contamination by metal ion enriched leachates [5], and deposition of contaminated ashes
resulting from waste incineration [57]. Bernardo et al. [24], reported high EC values and
high Cr, Cu, Mn, Ni, Pb, Zn, and Zr concentrations in the Hulene-B dump surrounding
soils, associating it with leachates. Another factor that may be associated with high EC is
the presence of shallow water [58]. In sandy soils, the presence of water during rainfall
episodes and periodic enrichment by surface leachates promote EC [37]. Studies related
EC to crop yields and found that high EC values can reduce yields [59]. However, for
horticultural crops, EC (salinity) results were considered within the recommended limit
(<1000 µS/cm) [60].
Soil OM content was ranked R1 > R4 > R2 > R3, coherent with other studies, with
generally low OM [61]. Some studies have shown that in contaminated areas, high OM
content in rhizosphere can lead to the uptake of heavy metals, which bind easily to OM [34]
and consequently decrease their toxicity levels from the rhizosphere to crops [62]. Rhizo-
sphere soils color was classified as very dark in all samples, with sample R1 the darkest,
associated with its OM content, which can be incorporated in interstitial spaces of sandy
soils [63]. Bernardo et al. [64] linked the Hulene-B surrounding soils color variation to
leachates circulating freely on the soil surface and waste incineration ashes, which can be
associated with rhizosphere soils PTEs contamination [65], which can be transferred to
edible plants [66].
Mineral phases are identified on the studied rhizosphere soils, where quartz (SiO2 ),
calcite (CaCO3 ), phyllosilicates (micas and kaolinite), anhydrite (CaSO4 ), K feldspar
(KAlSi3 O8 ), opal (SiO2 ·nH2 O), and gypsum (CaSO4 ·2H2 O) can be found (Figure 3), thus
suggesting differentiated processes between intra-dune and other soils around the Hulene-
B dump. Quartz content (18.1 to 40.5%) was similar to intra-dune deposits [32], which also
revealed high calcite content. Samples R2 and R4 showed a higher calcite content, possibly
due to the influence of soils remobilization in depth during the construction of the leachate
transport channels.
Phyllosilicates content (3.0 to 28.8%) may be linked to the predominant aeolian and
superficial deposition processes of the intra-dune area. Previous studies suggested that
phyllosilicates content present a higher PTEs adsorption capacity, e.g., [67]. Quartz, under
conditions of exceptional morphology fracture, can adsorb PTEs, such as Pb, Cu, and
Cr [68]. Calcite has been reported to have the ability to immobilize heavy metals such as
Pb, Fe, and Cd in contaminated soils near areas such as dumpsites [69].
Environments 2023, 10, 45 8 of 19

Figure 3. Mineral phases identified on rhizosphere soils (relative quantification, in %).

Soils PTEs (Table 2) revealed significant differences (one sample t-test; p < 0.05) be-
tween samples Cr, Fe, Mn, Ti, and Zr content. The cluster analysis formed two major
groups, one with sample R1 and another with samples R2 and R4, revealing an associ-
ation to mean background values, and these three variables were grouped with sample
R3. Sample R1 presented higher PTEs concentrations than other samples. The results
suggested an external Cu, Ni, Pb, Ti, Zn, and Zr contamination contribution, especially in
sample R1. This sample was collected on northwest section of the dump, relatively distant,
suggesting contamination by transported waste incineration ash, with this location being
in the direction of displacement and deposition of windborne suspended materials.

Table 2. Rhizosphere soil samples and potentially toxic elements content of background soils (Bkg)
(in mg/kg).

Var R1 R2 R3 R4 Bkg *
Cr 18.5 13.0 6.8 8.9 33.5
Cu 184 88 24 351 9.2
Fe 15,966 5954 3195 5480 6801
Mn 153 120 93 113 140
Ni 14.6 4.2 2.6 4.9 3.9
Pb 156 39.0 9.7 156 7.8
Ti 2782 2116 1775 1703 1754
Zn 35.7 38.7 3.7 9.8 3.0
Zr 289 207 170 148 123
* Bernardo et al. [70]

The rhizosphere soil pollution index (PI) of sample R1 revealed low pollution consid-
ering Cr, moderate pollution by Zr, Fe, and Ti, and high to very high pollution by Cu, Pb,
Zn, and Ni (Table 3). Sample R2 Cu, Pb, and Zn content contributed to a very high PI; Ni,
Ti, and Zr contributed to a moderate PI; and Ti contributed to a high PI. Similar results
were found for samples R3 and R4. The higher contributions of Cu, Pb, Zn and Ni PI values
in sample R1 may be associated with location, i.e., near the waste incineration and Julius
Nyerere Avenue on the W, with sources of ash and contaminated dust. High concentrations
of Cu, Pb, Zn, and Ni were reported in road dust samples next to the Hulene-B dump,
being associated with waste incineration, intense traffic, and pavement degradation [70].
Samples R2 and R4 showed higher Cu, Pb, Zn, Ni, Ti, and Zr content, possibly linked to
Environments 2023, 10, 45 9 of 19

the superficial circulation of leachates, which occurs cyclically. Bernardo et al. [70] reported
high EC values and associated a possible contamination by leachate flows. Sample R3
presented a high PI with respect Cu, associated with surface water enriched by leachates
circulation.

Table 3. Pollution index and pollution load index of the studied rhizosphere soil samples.

PI
Var PLI
R1 R2 R3 R4
Cr 0.6 0.4 0.0 0.3 0.0
Cu 20.0 9.6 2.6 38.2 4748
Fe 2.4 0.9 0.5 0.8 0.2
Mn 1.1 0.9 0.7 0.8 0.1
Ni 3.7 1.1 0.7 1.3 0.8
Pb 20.0 5.0 1.2 20.0 618
Ti 1.6 1.2 1.0 1.0 0.5
Zn 11.9 7.2 0.4 3.3 26.4
Zr 2.4 1.7 1.4 1.2 1.6

Pollution load index (PLI) was ranked Cu >> Pb >> Zn > Zr > Ni > Ti > Fe > Mn > Cr,
suggesting environmental deterioration in the surrounding area of the dumpsite. Sample
R1 was the one that contributed most PTEs and had a high rate of contamination. Sample
R1 was collected in the NW border of the dump, strongly influenced by waste incineration,
with ashes regularly transported and deposited in this location. On the dump surroundings,
the incineration of urban wastes, such as electronics, hospital waste, tires, and lamps, has
been pointed out as a source of Cr, Cu, Mn, Ni, Pb, and Zn [57]. Sample R4 was the one that
most contributed to this index, with high Cu and Pb content, and it was collected at the SW
border of the dump, where surface dispersion of leachate predominates. Leachates have
been identified as a source of PTEs in the surroundings of dumps [71] and are associated
with higher OM content, which may influence the fixation of contaminants [72]. Sample
R2, located north of the dump and close to the dump, with its relatively low contribution,
may be associated with possible leaching of contaminants in this strip. However, the high
pollution index by Cu, Pb, and Zn may be associated with the circulation of leachates that
cyclically affect this area. However, sample R3 showed a low contribution, which may be
related to its relatively distant location from the Hulene-B dump. High contributions of Cu
may be associated with new deposits in the southwest, where open-air burning and ash
deposition are recorded in the surrounding soils.
Rhizosphere soils hazard quotient (HQ) by ingestion and dermal contact for children
was considered, due to higher risk by playing on the ground and hand-to-mouth behaviors
in young age (Figure 4). All PTEs presented HQ > 1 by ingestion, except for Zr, with Mn
being the element posing a higher hazard followed by Cu, Ti, and Zn. Dermal contact
HQ was >1, except by Pb in samples R1 and R3, and Zr in all samples, with Ni and Mn
being the PTEs posing a higher hazard. Prolonged exposure to Mn can induce pneumonia,
hallucinations, pulmonary embolism, and bronchitis [71], while Ni poisoning affects lung
function, breathing difficulties, coughing, skin or eye irritation, and chronic exposure to Cu
can induce anemia and central nervous system and cardiovascular system disorders [73].
Environments 2023, 10, 45 10 of 19

Figure 4. Rhizosphere soil hazard quotient (HQ) by ingestion and dermal contact for children (relative
quantification, in %).

3.2. Water Analysis


Water pH was higher in stream waters, being higher than reference values (Table 4).
Water sample WS presented a pH slightly below recommended for consumption [38]. pH
is recognized to influence Hg, Pb, and Zn bioavailability and toxicity in drinking water [74].
Yesil et al. [75] reported higher bioavailability and toxicity of Hg and Pb in slightly acidic-to-
neutral waters, such as sample WS, suggesting a greater health risk. Irrigation water IWN
presented a higher Pb concentration than the reference value [52], which may be associated
with ash deposition, as Pb contamination is caused by burning waste such as electronic
equipment, batteries, paint cans, varnishes, and organic solvents, which is common [24].
In addition, free leachate circulation, which regularly reaches this area, is reported as a Pb
dispersion source, in environments around waste dumps [6]. IWS showed high Hg content,
which may also be associated with waste incineration [76].

Table 4. Water samples’ pH, Pb, Hg, and Zn concentrations and reference values (in µg/L).
μ
Samples Guidelines
Var
WS WN IWN IWS Drinking Water Stream Water Reference
pH 6.1 8.4 9.1 10.3 − 6.5−8.50 5.5–7.5 WHO [53]
Pb 6 nd 79 nd 10 20 WHO [53]
Hg 36 76 nd 25 6 5 WHO [53]
WHO [53];
Zn 63 13 7 9 3 ≤5.000
≤ CCME [77]
nd—not detected.

The water pollution index revealed severe Hg Pi in the two water samples collected
from wells (Table 5), while Pb Pi showed a slight pollution in sample WS and heavy Pi
in IWN. Zinc Pi was classified as medium polluted in both wells and slightly polluted in
IWS, taking in consideration Zn toxicity despite its higher concentration when compared
to Pb and Hg content. Mercury can incorporate into groundwaters and surface waters
through leachates produced at waste dumps by anaerobic decomposition of wastes or
by particulate material resulting from the burning of contaminated wastes [76]. The
practice of waste burning at the Hulene-B dump may be influencing the dispersion of ash
enriched by heavy metals that were successively deposited in the vicinity of the waste
dump and contributed to the contamination of groundwaters and stream waters. Lead
enrichment in areas surrounding landfills has been associated with the deposition and
burning of electronic equipment and rechargeable batteries wastes [78]. In the surroundings
Environments 2023, 10, 45 11 of 19

of the Hulene-B dumpsite, other potential Pb sources may be linked to, e.g., traffic-related
materials (Julius Nyerere Avenue) and an international airport (Figure 1). Zinc enrichment
has been associated with batteries, alloys, paint cans, cosmetics, pharmaceuticals, textiles,
and electrical and electronic equipment waste, which is continuously deposited and burnt
in the Hulene-B dump [79].

Table 5. Pollution index (Pi ) of groundwater and surface water.

ID Pb Hg Zn
WS 0.6 6.0 2.1
WN - 12.7 4.3
IWS 4.0 - 1.8
IWN - 5.0 0.6

3.3. Edible Plants


The potentially toxic elements concentration in edible plants (Table 6) was higher than
the daily reference dose (RfD) guidelines, suggesting a contribution from the Hulene-B
dump. Studies suggested that waste dumps can be sources of PTEs on adjacent crops [80],
with contamination also being associated with the transfer factor (TF) of contaminants from
soil to plant [81]. Another mechanism of edible plants contamination in the vicinity of
landfills is waste incineration ash deposition [82]. The consumption of these crops may
represent a potential health risk [41]. The mean PTEs concentration of the edible plant
samples was ranked RP1 > RP3 > RP2 > RP4. Sample RP1 (Amaranthus spinosus) was
collected in the NW region of the dump, an area influenced by waste incineration ash
deposition. However, Zr was not detected in edible leaves, despite its high concentrations
in rhizosphere soils, suggesting a low TF under alkaline pH conditions [75]. Sample RP3
(Brassica oleracea L), collected relatively distant from the dump, showed high PTEs levels,
despite rhizosphere soil contamination being relatively low, possibly associated with other
contamination mechanisms such as cyclic, open-air incineration of waste occurring in a
new area which may be a source of contaminated ash that accumulates on crop leaves.
Sample RP2 (Ipomoea batatas) was collected relatively close to the dump, with PTEs content
on rhizosphere soil being higher than the background mean, and plant edible leaves
revealing a PTEs concentration above reference values. Yan et al. [12] suggested that
Ipomoea batatas in contaminated soils presented higher capacity to absorb heavy metals
such as Cu, Fe, Pb, and Zn. Sample RP4 (Cucurbita pepo) was collected at the S boundary of
the landfill, and correspondent rhizosphere soil presenting higher PTEs concentration than
local background. Plant edible leaves showed a higher content than RfD.

Table 6. Edible plants PTEs (in mg/kg).

Var RP1 RP2 RP3 RP4 AL


Cr 974 852 2968 1161 2.3 a
Cu 1142 1245 778 1099 10–40 a
Fe 56,324 30,031 32,643 18,687 18 a
Mn 1030 1183 1436 1544 500 a
Ni 108 238 1128 538 10 a
Pb 394 56 190 nd 0.3 a
Ti 4990 5910 8392 3519 0.7 a
Zn 5730 2596 1511 3519 50 a
Zr nd 761 708 198 4.2 b
AL—acceptable limit; a Hurrell et al. [83]; b USEPA [84].

3.4. Data Integration


The transfer factor (TF) from soil to plants is an important factor for consumption
toxicity, with soil pH and OM being relevant in this process. Studied samples TF was
Environments 2023, 10, 45 12 of 19

>1 (Figure 5), except for Pb in R4P and Zr in R1P. High TF values suggested a higher
capacity of the plants to absorb contaminants [80,85]. Transfer factor was ranked RP3 >
RP4 > RP2 > RP1, with sample RP3 presenting higher TF for all PTEs, with Pb being the
most representative element. Lead can be absorbed from plants leaves by ash or road
dust deposition, with this plant being collected in a SE wind direction over the Hulene-B
landfill. Wang et al. [4] suggested that plant leaves grown in the surrounding of dumps are
constantly being enriched by PTEs from ashes resulting from waste incineration. Sample
RP4 Zn and Mn TF showed higher percentage rates.

Figure 5. Transfer factor (TF) from soil to plants (relative quantification, in %).

Daily elemental intake (DIM) and hazard risk index (HRI) are presented in Table 7.
Sample R1P (Amaranthus spinosus) DIM was ranked Fe >> Zn > Ti, with other PTEs being
negligible. The HRI was high for all elements, except for Ni and Zr, being ranked Cr > Cu >
Pb > Zn > Mn > Ti > Fe. Sample RP2 (Ipomoea batatas) DIM was >1 in Fe, Zr and Ti, while HRI
> 1 was ranked Cu > Cr > Mn > Pb. Sample RP3 (Cucurbita pepo) DIM > 1 was ranked Fe > Cr
> Ti, and HRI was ranked Cr > Cu > Pb > Ti > Mn > Ni > Zn. Sample RP4 (Brassica oleracea
L) DIM > 1 was ranked Fe > Zn > Ti, while HRI was ranked Cr > Cu > Pb > Zn > Mn > Ti >
Fe. The consumption of plants contaminated by PTEs can induce various diseases and even
cancer [86]. Chromium toxicity in humans has been linked to damaged blood cells, livers,
nervous systems, and kidneys [73]. Chronic Cu consumption may result in gastrointestinal
irritation, diarrhea, and liver failure [87], while Fe toxicity can cause vomiting, diarrhea,
and damage to the intestine and other organs [88], and Mn is linked to memory problems,
hallucinations, and Parkinson’s disease [89]. Nickel chronic ingestion has been associated
with gastrointestinal and neurological effects including lung cancer. Ingestion of Pb can
disturb almost all organs and the nervous system, in addition to causing kidney damage,
brain damage, miscarriage, and death [89], and Ti can cause damage to DNA, brain [90],
as well as liver and kidneys, and it can cause cancer [91], while excessive and prolonged
Zn intake can lead to anemia and affect the immune system [73,90]. Zirconium, which is
classified as having low toxicity [91], despite the presence and retention in relatively high
amounts in biological systems, has not yet been associated with any specific metabolic
function [92].
Environments 2023, 10, 45 13 of 19

Table 7. Daily elemental intake (DIM) and hazard risk index (HRI).

RP1 RP2 RP3 RP4


Var
DIM HRI DIM HRI DIM HRI DIM HRI
Cr 0.41 135.29 0.36 118.37 1.24 412.17 0.48 161.26
Cu 0.48 128.58 0.52 140.23 0.32 87.61 0.46 123.68
Fe 23.47 1.30 12.51 0.70 13.60 0.76 7.79 0.43
Mn 0.43 3.07 0.49 3.52 0.60 4.27 0.64 4.60
Ni 0.05 0.23 0.10 0.50 0.47 2.35 0.22 1.12
Pb 0.16 11.73 0.02 1.67 0.08 5.65 nd nd
Ti 2.08 2.97 2.46 3.52 3.50 5.00 1.47 2.09
Zn 2.39 7.96 1.08 3.61 0.63 2.10 1.47 4.89
Zr nd nd 0.32 0.08 0.30 0.07 0.08 0.02
nd—not detected.

Targeted hazard quotient (THQ) and hazard index (HIplant ) are presented in Figure 6,
with THQ being ranked Cr > Cu > Pb > Zn > Mn > Ni > Fe > Zr. The contribution of
sample RP1 (Amaranthus spinosus) to induce adverse health effects in children and adults
was ranked Cr > Cu > Pb> Zn > Mn > Ti > Fe, with nickel <1 and not posing potential risk
for adverse health effects. Sample RP2 (Ipomoea batatas) Fe, Ni, and Zr content do not pose
a potential health risk, with the main contributions being posed by Cu > Cr > Zn > Ti > Mn
> Pb. Sample RP3 (Brassica oleracea L) the main contributors to potential health outcomes
were ranked Cr > Cu > Pb > Ti > Mn > Ni > Zn, with Fe and Zr content representing <1.
Meanwhile, sample RP4 (Cucurbita pepo) PTEs were ranked Cr > Cu > Zn > Mn > Ti >
Ni, with Fe and Zr < 1. In all samples, the highest THQ risk index was posed to children.
Similar findings were reported as a result of their lower body mass, suggesting a higher
likelihood of developing cancer throughout life [41].

Figure 6. Targeted hazard quotient (THQ) in edible plants (relative quantification, in %).

HIplant was similar to THQ, being >1 in all samples and higher for children (Figure 7),
being ranked RP3 (6.41 × 102 to 4.27 × 102 ) > RP4 (3.68 × 102 to 2.45 × 102 ) > RP1
(3.59 × 102 to 2.39 × 102 ) > RP2 (3.36 × 102 –2.24 × 102 ). Long-term consumption of the
studied plants may induce cancer (e.g., kidney, bladder, and respiratory tract), nervous sys-
tem disorders, memory problems, hallucinations, cardiovascular diseases, and diabetes. In
Environments 2023, 10, 45 14 of 19

children, this has also been associated with cognitive development problems and aggressive
behavior [93,94].

Figure 7. Hazard index in edible plants.

Well water samples systemic toxicity (non-carcinogenic hazard) presented values


above 1.00 × 100 (Figure 8). The results are linked to Hg concentration, with Pb and Zn
being considerably negligible. Mercury is a toxic element, being ranked the third most toxic
among all PTEs [7]. It was estimated that chronic Hg consumption of about 0.00023 mg/L
per day can induce adverse renal, neurological, and respiratory effects [95,96], and it
can irreversibly damage kidneys, liver, and central nervous system [97]. Well waters do
not present a carcinogenic risk for humans by direct consumption, with Pb showing the
highest values of 1.09 × 10−−−7 and 6.55 × −−10−7 for WN and WS, respectively, both being
below the threshold of 1.00−− × 10−6 . The implication of lead on human health depends on
intensity, duration, and level of exposure, which can result in a range of toxic effects such as
hematological, neurological, psychological, renal, and genetic mutation and reproduction
are examples of the cumulative effect on the body [98]. As expected, in well waters,
children had a higher health risk than adults, due to their lower body mass [41]. Lead
toxicity in children can induce neurological complications, including difficulties in learning,
concentration, and aggressive behavior [89]. Water ingestion is considered one of the main
forms of Pb poisoning in children, accounting for about 7% [99].

Figure 8. Systemic toxicity of well water samples.


Environments 2023, 10, 45 15 of 19

4. Conclusions
Rhizosphere soils showed variable organic matter (OM) concentration, with moder-
ately alkaline pH. The identified mineral phases predominance was: calcite > quartz >
phyllosilicates (micas and kaolinite) > anhydrite > K feldspar > opal > gypsum, suggesting
a higher ability for fixing PTEs and thus ideal conditions for adsorption and absorption of
PTE in soils. The soil pollution load index (PLI) was ranked Cu >> Pb >> Zn >> Zr > Ni > Ti
> Fe > Mn > Cr. Edible plants showed PTES concentration above reference dose, suggesting
contamination by rhizosphere soil transfer factor and by contaminated ash deposition.
Hazard index was >1 for all selected elements in all plants, except for Ni (sample RP1), Ni
and Zr (sample RP2), Fe and Zr (sample R3), and Fe (sample R4), which suggests possible
adverse health events. Target hazard quotient was >1 for all elements and only Zr was less
than 1 in all plants. In all samples, the index was >1, and children were the ones with higher
risk. Groundwater and stream waters showed a high pollution index (Pi) for Hg and Zn,
with groundwater (drinking water) classified as acidic to slightly alkaline. Stream water
(irrigation) was moderately alkaline, suggesting variable toxicity given PTEs concentrations
above RfD.
The results suggested the need to control pollution mechanisms around the Hulene-B
dump, such as leachate circulation and waste incineration ash control, the main PTEs
enrichment sources of edible plants, rhizosphere soils, groundwater, and surface water.
There is a need to reduce the use of well water due to high concentrations of potentially toxic
elements. Farming should be relocated to other cropping areas, as part of the contamination
occurs through soil transfer to plants. There is a need to protect crops from ash deposition
to reduce toxicity. Future studies are needed to evaluate the levels of contamination of
the waste incineration ash, and to evaluate PTEs bioaccessibility in both soil, plants, and
waters.

Author Contributions: Conceptualization, C.C. and F.R.; formal analysis, B.B.; funding acquisition,
F.R.; investigation, B.B., C.C. and F.R.; methodology, B.B., C.C. and F.R.; resources, F.R.; supervision,
C.C. and F.R.; validation, C.C.; writing—original draft, B.B.; writing—review and editing, C.C. and
F.R. All authors have read and agreed to the published version of the manuscript.
Funding: The authors are grateful to FCT for the financial support to Research Unit GeoBioTec
(UIDB/04035/2020). The first author acknowledges grants from the Portuguese Institute Camões
and FNI (Investigation National Fund—Mozambique).
Institutional Review Board Statement: This research study does not involve animals.
Informed Consent Statement: This research study does not involve humans. This research study
does not involve individual person’s data in any form.
Data Availability Statement: Data used are available on the manuscript.
Conflicts of Interest: The authors declare no known competing financial interest or personal rela-
tionships that could have appeared to influence the work reported in this paper.

References
1. Dharwal, M.; Srivastava, A.K.; Sarin, V.; Gola, K. The state of solid waste management for sustainable development in India:
Current state and future potential. Mater. Today Proc. 2021, 60, 802–805. [CrossRef]
2. Morita, A.K.; Ibelli-Bianco, C.; Anache, J.A.; Coutinho, J.V.; Pelinson, N.S.; Nobrega, J.; Rosalem, L.M.; Leite, C.M.; Niviadonski,
L.M.; Manastella, C.; et al. Pollution threat to water and soil quality by dumpsites and non-sanitary landfills in Brazil: A review.
Waste Manag. 2021, 131, 163–176. [CrossRef] [PubMed]
3. Mitra, D.; Banerji, S. Urban hydrodynamics in the planned township of New Town, West Bengal, India. Appl. Geogr. 2020, 123,
102277. [CrossRef]
4. Wang, F.; Song, K.; He, X.; Peng, Y.; Liu, D.; Liu, J. Identification of Groundwater Pollution Characteristics and Health Risk
Assessment of a Landfill in a Low Permeability Area. Int. J. Environ. Res. Public Health 2021, 18, 7690. [CrossRef] [PubMed]
5. Hussein, M.; Yoneda, K.; Mohd-Zaki, Z.; Amir, A.; Othman, N. Heavy metals in leachate, impacted soils and natural soils of
different landfills in Malaysia: An alarming threat. Chemosphere 2020, 267, 128874. [CrossRef] [PubMed]
6. Parvin, F.; Tareq, S.M. Impact of landfill leachate contamination on surface and groundwater of Bangladesh: A systematic review
and possible public health risks assessment. Appl. Water Sci. 2021, 11, 100. [CrossRef]
Environments 2023, 10, 45 16 of 19

7. Khattak, S.A.; Rashid, A.; Tariq, M.; Ali, L.; Gao, X.; Ayub, M.; Javed, A. Potential risk and source distribution of groundwater
contamination by mercury in district Swabi, Pakistan: Application of multivariate study. Environ. Dev. Sustain. 2020, 23,
2279–2297. [CrossRef]
8. Thongyuan, S.; Khantamoon, T.; Aendo, P.; Binot, A.; Tulayakul, P. Ecological and health risk assessment, carcinogenic and
non-carcinogenic effects of heavy metals contamination in the soil from municipal solid waste landfill in Central, Thailand. Hum.
Ecol. Risk Assess. Int. J. 2020, 27, 876–897. [CrossRef]
9. Romero, A.; González, I.; Martín, J.M.; Vázquez, M.A.; Ortiz, P. Risk assessment of particle dispersion and trace element
contamination from mine-waste dumps. Environ. Geochem. Health 2014, 37, 273–286. [CrossRef]
10. Punia, A. Role of temperature, wind, and precipitation in heavy metal contamination at copper mines: A review. Environ. Sci.
Pollut. Res. 2020, 28, 4056–4072. [CrossRef]
11. Vongdala, N.; Tran, H.-D.; Xuan, T.D.; Teschke, R.; Khanh, T.D. Heavy Metal Accumulation in Water, Soil, and Plants of Municipal
Solid Waste Landfill in Vientiane, Laos. Int. J. Environ. Res. Public Health 2018, 16, 22. [CrossRef]
12. Yan, X.; An, J.; Yin, Y.; Gao, C.; Wang, B.; Wei, S. Heavy metals uptake and translocation of typical wetland plants and their
ecological effects on the coastal soil of a contaminated bay in Northeast China. Sci. Total. Environ. 2021, 803, 149871. [CrossRef]
13. Sharma, S.; Kaur, I.; Nagpal, A.K. Contamination of rice crop with potentially toxic elements and associated human health
risks—A review. Environ. Sci. Pollut. Res. 2021, 28, 12282–12299. [CrossRef]
14. Gupta, N.; Yadav, K.K.; Kumar, V.; Krishnan, S.; Kumar, S.; Nejad, Z.D.; Khan, M.M.; Alam, J. Evaluating heavy metals
contamination in soil and vegetables in the region of North India: Levels, transfer and potential human health risk analysis.
Environ. Toxicol. Pharmacol. 2020, 82, 103563. [CrossRef]
15. Zhou, H.; Yang, W.-T.; Zhou, X.; Liu, L.; Gu, J.-F.; Wang, W.-L.; Zou, J.-L.; Tian, T.; Peng, P.-Q.; Liao, B.-H. Accumulation of Heavy
Metals in Vegetable Species Planted in Contaminated Soils and the Health Risk Assessment. Int. J. Environ. Res. Public Health
2016, 13, 289. [CrossRef] [PubMed]
16. Halder, S.; Agüero, J.; Dolle, P.; Fernández, E.; Schmidt, C.; Yang, M. Perspectives of Urban Agriculture in Maputo and Cape Town.
2018. Available online: https://www.sle-berlin.de/files/sle/publikationen/S%20275-Maputo-Internet-Klein.pdf (accessed on 20
October 2021).
17. Bernardo, B.; Candeias, C.; Rocha, F. Application of Geophysics in geo-environmental diagnosis on the surroundings of the
Hulene-B waste dump, Maputo, Mozambique. J. Afr. Earth Sci. 2021, 185, 104415. [CrossRef]
18. Serra, C. Da Problemática Ambiental à Mudança: Rumo à um Mundo Melhor; Editora Escolar: Maputo, Mozambique, 2012;
ISBN 0A9789896700300. (In Portuguese)
19. Tvedten, I.; Candiracci, S. “Flooding our eyes with rubbish”: Urban waste management in Maputo, Mozambique. Environ. Urban.
2018, 30, 631–646. [CrossRef]
20. Cendón, D.I.; Haldorsen, S.; Chen, J.; Hankin, S.; Nogueira, G.; Momade, F.; Achimo, M.; Muiuane, E.; Mugabe, J.; Stigter, T.Y.
Hydrogeochemical aquifer characterization and its implication for groundwater development in the Maputo district, Mozambique.
Quat. Int. 2019, 547, 113–126. [CrossRef]
21. Nogueira, G.; Stigter, T.; Zhou, Y.; Mussa, F.; Juizo, D. Understanding groundwater salinization mechanisms to secure freshwater
resources in the water-scarce city of Maputo, Mozambique. Sci. Total. Environ. 2018, 661, 723–736. [CrossRef]
22. Bernardo, B.; Candeias, C.; Rocha, F. Characterization of the Dynamics of Leachate Contamination Plumes in the Surroundings of
the Hulene-B Waste Dump in Maputo, Mozambique. Environments 2022, 9, 19. [CrossRef]
23. Vicente, E.M.; Jermy, C.A.; Schreiner, H.D. Urban Geology of Maputo, Mozambique; The Geological Society of London: London, UK,
2006; Volume 338, pp. 1–13. Available online: https://citeseerx.ist.psu.edu/viewdoc/download?doi=10.1.1.606.7220&rep=rep1
&type=pdf (accessed on 20 October 2021).
24. Bernardo, B.; Candeias, C.; Rocha, F. Soil Risk Assessment in the Surrounding Area of Hulene-B Waste Dump, Maputo
(Mozambique). Geosciences 2022, 12, 290. [CrossRef]
25. INE. Boletim de Estatísticas Demográficas e Sociais, Maputo Cidade 2019; Instituto Nacional de Estatistica: Maputo, Mozambique, 2020.
Available online: https://www.ine.gov.mz/estatisticas/estatisticas-demograficas-e-indicadores-sociais/boletim-de-indicadores-
demograficos-22-de-julho-de-2020.pdf/at_download/fileinportuguese (accessed on 20 October 2021).
26. Matsinhe, F.O.; Paulo, M. Estudo Etnográfico sobre os catadores de Lixo da Lixeira de Hulene (Maputo). Cad. De África Contemp.
2020, 3, 11–28.
27. dos Muchangos, L.S.; Tokai, A.; Hanashima, A. Analyzing the structure of barriers to municipal solid waste management policy
planning in Maputo city, Mozambique. Environ. Dev. 2015, 16, 76–89. [CrossRef]
28. Ferrão, D.A.G. Evaluation of Removal and Disposal of Solid Waste in Maputo City, Mozambique. Master’s Thesis, University of
Cape Town, Cape Town, South Africa, 2006. Available online: https://open.uct.ac.za/bitstream/handle/11427/4851/thesis_sci_
2006_ferrao_d_a_g.pdf?sequence=1 (accessed on 20 October 2021).
29. Palalane, J.; Segala, I.O. Urbanização e Desenvolvimento Municipal em Moçambique: Gestão de Resíduos Sólidos. 2008.
Available online: https://limpezapublica.com.br/urbanizacao-e-desenvolvimento-municipal-em-mocambique-capitulo-gestao-
de-residuos-solidos/ (accessed on 20 October 2021). (In Portuguese).
30. VOA (Voice of America News). Desabamento de Lixeira Deixa 17 Mortos em Maputo. 2018. Available online: https://www.
voaportugues.com/a/desabamento-lixeira-17-mortos-maputo/4260624.html (accessed on 22 May 2022).
Environments 2023, 10, 45 17 of 19

31. Muchimbane, A.B.D.A. Estudo dos Indicadores de Contaminação das Águas Subterrâneas por Sistemas de Saneamento “In Situ”
Distrito Urbano 4, Cidade de Maputo-Moçambique. Ph.D. Thesis, Universidade de São Paulo, São Paulo, Brazil, 2010. [CrossRef]
32. Momade, F.J.; Ferrara, M.; Oliveira, J.T. Notícia Explicativa da Carta Geológica 2532 Maputo (Escala 1:50000); Instituto Nacional de
Estatistica: Maputo, Mozambique, 1996. (In Portuguese)
33. CIAT. Climate-Smart Agriculture in Mozambique. Center for Tropical Agriculture. 2017. Available online: https:
//climateknowledgeportal.worldbank.org/sites/default/files/2019-06/CSA-inMozambique.pdf (accessed on 14 May 2022).
34. USDA. Rangeland Soil Quality-Organic Matter; United States Department of Agriculture: Washington, DC, USA, 2001. Available
online: https://www.ftw.nrcs.usda.gov/glti (accessed on 12 July 2022).
35. Munsell Color. Munsell Soil Color Book; Color Charts; Munsell Color Company Inc.: Newburgh, NY, USA, 2009.
36. USDA. Soil Quality Indicators: pH. Managing Soils and Terrestrial Systems; United States Department of Agriculture: Washington,
DC, USA, 1998. Available online: http://soils.usda.gov (accessed on 24 July 2022).
37. USDA. Soil Health—Electrical Conductivity; United States Department of Agriculture: Washington, DC, USA, 2014; pp. 1–9.
Available online: https://www.nrcs.usda.gov/Internet/FSE_DOCUMENTS/nrcs142p2_052803.pdf (accessed on 22 July 2022).
38. World Health Organization. Guidelines for Drinking-Water Quality: Fourth Edition Incorporating First Addendum; World Health
Organization: Geneva, Switzerland, 2017; ISBN 978-92-4-154995-0.
39. Bhardwaj, S.; Soni, R.; Gupta, S.K.; Shukla, D.P. Mercury, arsenic, lead and cadmium in waters of the Singrauli coal mining and
power plants industrial zone, Central East India. Environ. Monit. Assess. 2020, 192, 251. [CrossRef]
40. Håkanson, L. An ecological risk index for aquatic pollution control. A sedimentological approach. Water Res. 1980, 14, 975–1001.
[CrossRef]
41. Candeias, C.; Ávila, P.F.; Sequeira, C.; Manuel, A.; Rocha, F. Potentially toxic elements dynamics in the soil rhizospheric-plant
system in the active volcano of Fogo (Cape Verde) and interactions with human health. Catena 2021, 209, 105843. [CrossRef]
42. Zhang, H.; Zhang, F.; Song, J.; Tan, M.L.; Kung, H.-T.; Johnson, V.C. Pollutant source, ecological and human health risks
assessment of heavy metals in soils from coal mining areas in Xinjiang, China. Environ. Res. 2021, 202, 111702. [CrossRef]
43. RAIS. The Risk Assessment Information System (RAIS); U.S. Department of Energy’s Oak Ridge Operations Office (ORO): Oak
Ridge, TN, USA, 2021. Available online: https://rais.ornl.gov/ (accessed on 24 July 2022).
44. Candeias, C.; Da Silva, E.F.; Ávila, P.F.; Teixeira, J.P. Identifying Sources and Assessing Potential Risk of Exposure to Heavy Metals
and Hazardous Materials in Mining Areas: The Case Study of Panasqueira Mine (Central Portugal) as an Example. Geosciences
2014, 4, 240–268. [CrossRef]
45. Pehoiu, G.; Murarescu, O.; Radulescu, C.; Dulama, I.D.; Teodorescu, S.; Stirbescu, R.M.; Bucurica, I.A.; Stanescu, S.G. Heavy
metals accumulation and translocation in native plants grown on tailing dumps and human health risk. Plant Soil 2020, 456,
405–424. [CrossRef]
46. Ogunwale, T.O.; Ogar, P.A.; Kayode, G.F.; Salami, K.D.; Oyekunle, J.A.O.; Ogunfowokan, A.O.; Akindolani, O.A. Health Risk
Assessment of Heavy Metal Toxicity Utilizing Eatable Vegetables from Poultry Farm Region of Osun State. J. Environ. Pollut.
Hum. Health 2021, 9, 6–15. [CrossRef]
47. USEPA. Risk Assessment Guidance for Superfund. In Volume I Human Health Evaluation Manual (Part A); U.S. Environmental
Protection Agency: Washington, DC, USA, 1989.
48. Lee, E.; Rout, P.R.; Bae, J. The applicability of anaerobically treated domestic wastewater as a nutrient medium in hydroponic
lettuce cultivation: Nitrogen toxicity and health risk assessment. Sci. Total. Environ. 2021, 780, 146482. [CrossRef]
49. Yan, C.-A.; Zhang, W.; Zhang, Z.; Liu, Y.; Deng, C.; Nie, N. Assessment of Water Quality and Identification of Polluted Risky
Regions Based on Field Observations & GIS in the Honghe River Watershed, China. PLoS ONE 2015, 10, e0119130. [CrossRef]
50. Ajibare, A.O.; Ogungbile, P.O.; Ayeku, P.O. Evaluation of water pollution monitoring for heavy metal contamination: A case
study of Agodi Reservoir, Oyo State, Nigeria. Environ. Monit. Assess. 2022, 194, 675. [CrossRef]
51. Adimalla, N.; Li, P.; Venkatayogi, S. Hydrogeochemical Evaluation of Groundwater Quality for Drinking and Irrigation Purposes
and Integrated Interpretation with Water Quality Index Studies. Environ. Process 2018, 5, 363–383. [CrossRef]
52. Yin, Z.; Duan, R.; Li, P.; Li, W. Water quality characteristics and health risk assessment of main water supply reservoirs in Taizhou
City, East China. Hum. Ecol. Risk Assessment Int. J. 2021, 27, 2142–2160. [CrossRef]
53. WHO. Mercury in Drinking-Water, Background Document for Development of WHO Guidelines for Drinking-Water Quality; World
Health Organization: Geneva, Switzerland, 2005; WHO/SDE/WS, WHO/SDE/WSH/05.08/10. Available online: http://www.
who.int/water_sanitation_health/dwq/chemicals/mercuryfinal.pdf (accessed on 1 July 2022).
54. Ahmad, W.; Alharthy, R.D.; Zubair, M.; Ahmed, M.; Hameed, A.; Rafique, S. Toxic and heavy metals contamination assessment in
soil and water to evaluate human health risk. Sci. Rep. 2021, 11, 17006. [CrossRef]
55. Naveen, B.; Mahapatra, D.M.; Sitharam, T.; Sivapullaiah, P.; Ramachandra, T. Physico-chemical and biological characterization of
urban municipal landfill leachate. Environ. Pollut. 2017, 220, 1–12. [CrossRef]
56. Gajaje, K.; Ultra, V.U.; David, P.W.; Rantong, G. Rhizosphere properties and heavy metal accumulation of plants growing in the
fly ash dumpsite, Morupule power plant, Botswana. Environ. Sci. Pollut. Res. 2021, 28, 20637–20649. [CrossRef]
57. Škrbić, B.; Novaković, J.; Miljević, N. Mobility of heavy metals originating from bombing of industrial sites. J. Environ. Sci. Health
Part A 2002, 37, 7–16. [CrossRef] [PubMed]
58. Li, H.; Sun, J.; Gui, H.; Xia, D.; Wang, Y. Physiochemical properties, heavy metal leaching characteristics and reutilization
evaluations of solid ashes from municipal solid waste incinerator plants. Waste Manag. 2021, 138, 49–58. [CrossRef]
Environments 2023, 10, 45 18 of 19

59. Chaaou, A.; Chikhaoui, M.; Naimi, M.; El Miad, A.K.; Achemrk, A.; Seif-Ennasr, M.; El Harche, S. Mapping soil salinity risk
using the approach of soil salinity index and land cover: A case study from Tadla plain, Morocco. Arab. J. Geosci. 2022, 15, 722.
[CrossRef]
60. Cheng, M.; Wang, H.; Fan, J.; Wang, X.; Sun, X.; Yang, L.; Zhang, S.; Xiang, Y.; Zhang, F. Crop yield and water productivity under
salty water irrigation: A global meta-analysis. Agric. Water Manag. 2021, 256, 107105. [CrossRef]
61. USDA. Soil Health Quality Indicators: Chemical Properties, Soil Electrical Conductivity, 3rd ed.; United States Department of
Agriculture: Washington, DC, USA, 2011. Available online: https://www.nrcs.usda.gov/wps/portal/nrcs/detail/soils/health/
assessment/?cid=stelprdb1237387 (accessed on 12 July 2022).
62. Huang, B.; Yuan, Z.; Li, D.; Zheng, M.; Nie, X.; Liao, Y. Effects of soil particle size on the adsorption, distribution, and migration
behaviors of heavy metal(loid)s in soil: A review. Environ. Sci. Process Impacts 2020, 22, 1596–1615. [CrossRef]
63. Nisari, A.; Sujatha, C. Assessment of trace metal contamination in the Kol wetland, a Ramsar site, Southwest coast of India. Reg.
Stud. Mar. Sci. 2021, 47, 101953. [CrossRef]
64. Seidl, M.; Le Roux, J.; Mazerolles, R.; Bousserrhine, N. Assessment of leaching risk of trace metals, PAHs and PCBs from a
brownfield located in a flooding zone. Environ. Sci. Pollut. Res. 2021, 29, 3600–3615. [CrossRef]
65. Bernardo, B.; Candeias, C.; Rocha, F. Soil properties and environmental risk assessment of soils in the surrounding area of
Hulene-B waste dump, Maputo (Mozambique). Environ. Earth Sci. 2022, 81, 542. [CrossRef]
66. Alexakis, D. Multielement Contamination of Land in the Margin of Highways. Land 2021, 10, 230. [CrossRef]
67. Alexakis, D.E. Suburban areas in flames: Dispersion of potentially toxic elements from burned vegetation and buildings.
Estimation of the associated ecological and human health risk. Environ. Res. 2020, 183, 109153. [CrossRef]
68. Asowata, I.T. Geophagic clay around Uteh-Uzalla near Benin: Mineral and trace elements compositions and possible health
implications. SN Appl. Sci. 2021, 3, 569. [CrossRef]
69. Bai, B.; Nie, Q.; Zhang, Y.; Wang, X.; Hu, W. Cotransport of heavy metals and SiO2 particles at different temperatures by seepage.
J. Hydrol. 2020, 597, 125771. [CrossRef]
70. Sharma, M.; Satyam, N.; Reddy, K.R.; Chrysochoou, M. Multiple heavy metal immobilization and strength improvement of
contaminated soil using bio-mediated calcite precipitation technique. Environ. Sci. Pollut. Res. 2022, 29, 51827–51846. [CrossRef]
71. Bernardo, B.; Candeias, C.; Rocha, F. Integration of Electrical Resistivity and Modified DRASTIC Model to Assess Groundwater
Vulnerability in the Surrounding Area of Hulene-B Waste Dump, Maputo, Mozambique. Water 2022, 14, 1746. [CrossRef]
72. Kazemi, Z.; Arani, M.H.; Panahande, M.; Kermani, M.; Kazemi, Z. Chemical quality assessment and health risk of heavy metals
in groundwater sources around Saravan landfill, the northernmost province of Iran. Int. J. Environ. Anal. Chem. 2021, 1–19.
[CrossRef]
73. Cheela, V.R.S.; Goel, S.; John, M.; Dubey, B. Characterization of municipal solid waste based on seasonal variations, source and
socio-economic aspects. Waste Dispos. Sustain. Energy 2021, 3, 275–288. [CrossRef]
74. Kennou, B.; El Meray, M.; Romane, A.; Arjouni, Y. Assessment of heavy metal availability (Pb, Cu, Cr, Cd, Zn) and speciation in
contaminated soils and sediment of discharge by sequential extraction. Environ. Earth Sci. 2015, 74, 5849–5858. [CrossRef]
75. Githaiga, K.B.; Njuguna, S.M.; Gituru, R.W.; Yan, X. Water quality assessment, multivariate analysis and human health risks of
heavy metals in eight major lakes in Kenya. J. Environ. Manag. 2021, 297, 113410. [CrossRef]
76. Yesil, H.; Molaey, R.; Calli, B.; Tugtas, A.E. Extent of bioleaching and bioavailability reduction of potentially toxic heavy metals
from sewage sludge through pH-controlled fermentation. Water Res. 2021, 201, 117303. [CrossRef] [PubMed]
77. CCME. Canadian Soil Quality Guidelines for the Protection of Environmental and Human Health-Nickel; Canadian Council of Ministers
of the Environment: Canberra, ACT, Canada, 2015. Available online: https://ccme.ca/en/res/nickel-canadian-soil-quality-
guidelines-for-the-protection-of-environmental-and-human-health-en.pdf (accessed on 10 April 2022).
78. Ruengruehan, K.; Junggoth, R.; Suttibak, S.; Sirikoon, C.; Sanphoti, N. Contamination of Cadmium, Lead, Mercury and Manganese
in Leachate from Open Dump, Controlled Dump and Sanitary Landfill Sites in Rural Thailand: A Case Study in Sakon Nakhon
Province. Nat. Environ. Pollut. Technol. 2021, 20, 1257–1261. [CrossRef]
79. Alghamdi, A.G.; Aly, A.A.; Ibrahim, H.M. Assessing the environmental impacts of municipal solid waste landfill leachate on
groundwater and soil contamination in western Saudi Arabia. Arab. J. Geosci. 2021, 14, 350. [CrossRef]
80. Islamd, S.; Idris, A.M.; Islam, A.R.M.T.; Phoungthong, K.; Ali, M.M.; Kabir, H. Geochemical variation and contamination level of
potentially toxic elements in land-uses urban soils. Int. J. Environ. Anal. Chem. 2021, 1–18. [CrossRef]
81. Alfaro, M.R.; Ugarte, O.M.; Lima, L.H.V.; Silva, J.R.; da Silva, F.B.V.; Lins, S.A.D.S.; Nascimento, C.W.A.D. Risk assessment of
heavy metals in soils and edible parts of vegetables grown on sites contaminated by an abandoned steel plant in Havana. Environ.
Geochem. Health 2021, 44, 43–56. [CrossRef]
82. Hermans, T.D.; Dougill, A.J.; Whitfield, S.; Peacock, C.L.; Eze, S.; Thierfelder, C. Combining local knowledge and soil science for
integrated soil health assessments in conservation agriculture systems. J. Environ. Manag. 2021, 286, 112192. [CrossRef]
83. Hurrell, R.; Egli, I. Iron bioavailability and dietary reference values. Am. J. Clin. Nutr. 2010, 91, S1461–S1467. [CrossRef]
84. USEPA. Silver and zirconium phosphate. In Federal Register; U.S. Environmental Protection Agency: Washington, DC, USA, 2010;
Volume 75. [CrossRef]
85. Weibel, G.; Eggenberger, U.; Schlumberger, S.; Mäder, U.K. Chemical associations and mobilization of heavy metals in fly ash
from municipal solid waste incineration. Waste Manag. 2017, 62, 147–159. [CrossRef] [PubMed]
Environments 2023, 10, 45 19 of 19

86. Balasooriya, S.; Diyabalanage, S.; Yatigammana, S.K.; Ileperuma, O.A.; Chandrajith, R. Major and trace elements in rice paddy
soils in Sri Lanka with special emphasis on regions with endemic chronic kidney disease of undetermined origin. Environ.
Geochem. Health 2021, 44, 1841–1855. [CrossRef] [PubMed]
87. Ashraf, I.; Ahmad, F.; Sharif, A.; Altaf, A.R.; Teng, H. Heavy metals assessment in water, soil, vegetables and their associated
health risks via consumption of vegetables, District Kasur, Pakistan. SN Appl. Sci. 2021, 3, 552. [CrossRef]
88. Ali, M.U.; Liu, G.; Yousaf, B.; Abbas, Q.; Ullah, H.; Munir, M.A.M.; Fu, B. Pollution characteristics and human health risks of
potentially (eco)toxic elements (PTEs) in road dust from metropolitan area of Hefei, China. Chemosphere 2017, 181, 111–121.
[CrossRef] [PubMed]
89. USEPA. Manganese Compounds Hazard Summary. In Health Effects Notebook for Hazardous Air Pollutants; U.S. Environmental
Protection Agency: Washington, DC, USA, 2016. Available online: https://www.epa.gov/sites/production/files/2016-10/
documents/manganese.pdf (accessed on 1 July 2022).
90. Wani, A.L.; Ara, A.; Usmani, J.A. Lead toxicity: A review. Interdiscip. Toxicol. 2015, 8, 55–64. [CrossRef]
91. (Ans), E.P.O.F.A.A.N.S.A.T.F.; Younes, M.; Aggett, P.; Aguilar, F.; Crebelli, R.; Dusemund, B.; Filipič, M.; Frutos, M.J.; Galtier, P.;
Gott, D.; et al. Evaluation of four new studies on the potential toxicity of titanium dioxide used as a food additive (E 171). EFSA J.
2018, 16, e05366. [CrossRef]
92. Price, G.A.V.; Stauber, J.L.; Holland, A.; Koppel, D.J.; Van Genderen, E.J.; Ryan, A.C.; Jolley, D.F. The Influence of pH on Zinc
Lability and Toxicity to a Tropical Freshwater Microalga. Environ. Toxicol. Chem. 2021, 40, 2836–2845. [CrossRef]
93. Jones, J.V.; Piatak, N.M.; Bedinger, G.M. Zirconium and Hafnium. In U.S. Geological Survey Professional Paper 1802; U.S. Geological
Survey: Reston, VA, USA, 2017; pp. V1–V26. [CrossRef]
94. WHO. Air Pollution. In Compendium of WHO and Other UN Guidance on Health and Environment Chapter 2; World Health
Organization: Geneva, Switzerland, 2021; pp. 1–25. Available online: https://cdn.who.int/media/docs/default-source/who-
compendium-on-health-and-environment/who_compendium_chapter2_01092021.pdf?sfvrsn=14f84896_5 (accessed on 19 April
2022).
95. WHO. WHO Guideline for Clinical Management of Exposure to Lead Executive Summary; World Health Organization: Geneva,
Switzerland, 2021. Available online: https://www.who.int/publications/i/item/9789240036888 (accessed on 8 July 2022).
96. Ekawanti, A.; Priyambodo, S.; Kadriyan, H.; Syamsun, A.; A Lestarini, I.; Wirasaka, G.; Ardianti, A.R. Mercury pollution in water
and its effect on renal function of school age children in gold mining area Sekotong West Lombok. IOP Conf. Series Earth Environ.
Sci. 2021, 637, 012055. [CrossRef]
97. Fonge, B.A.; Larissa, M.T.; Egbe, A.M.; Afanga, Y.A.; Fru, N.G.; Ngole-Jeme, V.M. An assessment of heavy metal exposure risk
associated with consumption of cabbage and carrot grown in a tropical Savannah region. Sustain. Environ. 2021, 7, 1909860.
[CrossRef]
98. Fiala, M.; Hwang, H.-M. Influence of Highway Pavement on Metals in Road Dust: A Case Study in Houston, Texas. Water Air Soil
Pollut. 2021, 232, 185. [CrossRef]
99. Obasi, P.N.; Akudinobi, B.B. Potential health risk and levels of heavy metals in water resources of lead–zinc mining communities
of Abakaliki, southeast Nigeria. Appl. Water Sci. 2020, 10, 184. [CrossRef]

Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual
author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to
people or property resulting from any ideas, methods, instructions or products referred to in the content.
348
349

13.2.7 Appendix 8. Paper 7


350
Manuscript File Click here to view linked References

1 Maputo urban area (Mozambique) road dust characterization and risk


2 assessment
3 Bernardino Bernardo1,2, Carla Candeias1, Fernando Rocha1

4 1
GeoBioTec Research Centre, Department of Geosciences, University of Aveiro, 3810-193 Aveiro, Portugal

5 2
Faculty of Earth Sciences and Environment, Pedagogic University of Maputo, Mozambique

6 Abstract

7 Roads are important source sources of pollutants, generating by wind and turbulence generated by cars
8 road dust particles with impact in the environment and human health. In this study, road dust samples,
9 fractions < 63 and 2000 µm, were collected in Maputo urban area (Mozambique). Dust was characterized
10 on their physical, mineralogical (XRD), and geochemical (XRF) content. Environmental and human health
11 risk was assessed. Mineralogical analysis revealed the predominance of quartz (SiO2) >> phyllosilicates
12 (mainly micas and kaolinite) > K-feldspars (KalSi3O8) > plagioclase ((Na,Ca)((Si,Al)AlSi2)O8) > carbonates
13 (calcite (CaCO3), dolomite (CaMg(CO3)2) > sulphates (alunite (KAl3(SO4)2(OH)6), anhydrite (CaSO4)).
14 Contamination index taking in consideration Potentially toxic elements (PTEs) Cr, Cu, Fe, Mn, Ni, Pb, Ti,
15 Zn, and Zr revealed concentration above guidelines. Carcinogenic risk of the <63 µm fraction was due to
16 road dust Ni and Pb concentration. Same fraction hazard quotient, or non-carcinogenic hazard, suggested
17 that ingestion was the main exposure route for children, mostly due to, Cu, Fe, Ni and Pb content.

18 Keywords: road dust, mineralogy, geochemistry, risk assessment, Maputo urban area

19 Introduction

20 Environmental pollution is a growing challenge worldwide, particularly in urban areas of developing


21 countries, where road and air traffic emissions, re-suspended dust, open-air waste deposition, and
22 degraded roads pavement, contribute significantly to air quality (Kraidy et al., 2022; Nematollahi et al.,
23 2020; Shahab et al., 2022). Road dust has been described as a complex mixture of particles containing
24 various organic and inorganic chemical components (Heo et al., 2021; Li et al., 2021). Zhang et al. (2019)
25 refer that road dust consists of soil, airborne particles, building construction materials, soot and smoke
26 from industry and vehicles related material, being one of the most important sources of environmental
27 pollutants emission (Pio et al., 2020; Zhao et al., 2021). Road dust with high concentrations of potentially
28 toxic elements (PTEs), such as As, Cd, Cu, Cr, Hg, Pb, Ni, Mn, Zn and Zr, is specially worrying for
29 environmental and human health (Alves et al., 2018; Candeias et al., 2021; Khan and Strand, 2018). In
30 addition to anthropogenic contribution, soil particles, enriched in Si, Ca, and Al, can also induce health
31 outcomes (Dehghani et al., 2017; Johansson et al., 2009).

32 Frequently reported adverse health effects related to road dust include chronic obstructive pulmonary
33 disease, asthma, allergy, carcinoma, emergency cardiovascular disease problems, increased mortality due
34 to cardiovascular disease, low birth weight and non-specific carcinoma (Candeias et al., 2020). The most
35 harmful effects were reported in dry climates regions or during dry season, where road dust is often
36 enriched by PTEs (Amato et al., 2014). Dust enriched in Ni and Pb, for example, when inhaled or ingested
37 chronically, has the capacity to cause bronchitis, abdominal pain, prolonged decrease of lung function and
38 cancer (Kazemi et al., 2021; Zhao et al., 2021). In the case of dust with high percentage of quartz,
39 prolonged inhalation may induce silicosis, pulmonary fibrosis, airway obstruction, and lung cancer (Leung
40 et al., 2012).

41 In recent years, African countries have recorded an increase in road systems in urban centers (Naidja et
42 al., 2018). However, most of the asphalted roads and cars are in a relatively degraded state (Fayiga et al.,
43 2018). Increased vehicle flow generates air pollution through road dust, linked to exhaust emissions,
44 resuspension of pavement surfaces, and brakes and tires related materials (Beji et al., 2020; Candeias et
45 al. 2020; IQAir, 2020; Kumar et al. 2020; Naidja et al., 2018; Sahu et al. 2016). Cars exhaust emissions has
46 been identified as a source of Pb in many urban areas of African cities, being related to the occurrence of
47 respiratory and cardiovascular diseases and allergies (Kalisa et al., 2019; Fayiga et al., 2018; Perera, 2018;
48 Sharma et al., 2004). Brake wear has been considered a major source of Cu in road dust, due to common
49 braking and acceleration occurrence on degraded roads, increasing emissions and resuspension (Beji et
50 al., 2020; Grigoratos et al., 2015). Han et al. (2014) suggested that tires wear and road pavement were a
51 major source of Mn, Ni, Cu, Zn, Pb concentration, on a study of Kuala Lumpur, representing a negative
52 impact on human health. In Africa, urban areas systematic studies on air contamination levels have been
53 deficient (Fayiga et al., 2018; Joubert et al., 2020), especially once population morbidity and mortality rate
54 associated with road dust has been estimated to be high (Coker et al. 2018; Joubert et al. 2020; Mboera
55 et al. 2014). Previous studies associated road dust and human health deprivation, being children and
56 elderly the risk groups, as well as people living in houses close to roads with deficient ventilation systems
57 (Alghamdi et al., 2022; Candeias et al., 2021; Fayiga et al. 2018; Karuppasamy et al., 2022).

58 In Mozambique, Maputo urban area has one of the highest levels of vehicle circulation and a dense road
59 network (Rwakarehe, 2022). However, the main roads providing access to the city and its outskirts are in
60 a relatively high stage of asphalt degradation, contributing to road dust contamination, and thus posing a
61 risk to the population living and working on roadsides and engaging in commercial activities (Bernardo et
62 al., 2022a; Gascon et al., 2016). This research aims to characterize Maputo urban area road dust and to
63 estimate the potential human health risk. Results can provide a basis for spatial perception of road dust
64 contamination and assist policy makers in defining measures to minimize human health risk.

65 2. Materials and methods

66 2.1 Study area

67 Maputo is the capital of Mozambique, with a population of about 1,127,565 inhabitants (INE, 2020) and
68 the highest growth rate in the country (WB, 2017; Serra, 2012; Cruz et al., 2005;). Most of the population
69 is young and female, often without formal employment and therefore engaged in commercial activities in
70 markets, streets, and squares (Cruz et al., 2005). These areas were characterized by intense traffic,
71 degraded pavement, uncontrolled solid waste disposal, and scarce vegetation, promoting dust dispersion
72 and exposing population to road dust by inhalation, ingestion, and dermal contact (Naidja et al., 2018;
73 Gascon et al., 2016; WHO, 2020; IQAir, 2020). The predominant climate is of subtropical type, with mean
74 annual precipitation of ~789 mm , with two climatic seasons: (a) hot (mean 25 ºC) and rainy period from
75 December to March, representing > 60 % of the annual precipitation, with its peak in January (~125 mm),
76 and (b) dry and cold season with lower temperatures in June and July (mean 21 ºC), and scarce
77 precipitation (CIAT, 2017). Prevailing winds were SE (Muchangos 1999).

78 2.2 Sampling

79 Road dust samples were collected, on the same day on a 1 m2 area, at Maputo urban area main roads,
80 characterized by heavy traffic, informal roadside commercial activities, and numerous dwellings (Fig. S1).
81 Other existing features in the surroundings were Maputo International Airport, and the largest open-air
82 dumpsite of the city (Hulene-B waste dump (Bernardo et al., 2022a) (Fig. 1).

83 Sample 1 (Fig. S2a) was collected southeast of Maputo International Airport, characterized by intense
84 traffic and street food commercial activities. Sample 2 (Fig. S2b) was collected in Beira Street, in the bus
85 stop, with diverse commercial small shops and dwellings. Sample 3 (Fig. S2c) was collected between the
86 eastern boundary of Hulene-B waste dump and Julius Nyerere Avenue, with heterogeneous features and
87 densely inhabited. Sample 4 (Fig. S2d) was collected in the southwest of Maputo airport, characterized by
88 heavy traffic and densely populated. Samples 5 and 6 (Fig. S2e,f) were collected in bus stops along the
89 National Road n.º 1, characterized by intense traffic and several commercial activities. Sample 7 (Fig. S2g)
90 was collected at a bus stop, between the National Road 1 (intense traffic) and the beginning of Lurdes
91 Mutola Avenue (quite degraded pavement, intense traffic, commercial street activities). Samples 8 and 9
92 (Figs. S2h,i) were collected in bus stops, with intense commerce on the sidewalks. All samples were
93 collected in June 2021. This period was characterized by scarce precipitation.

94 2.3 Samples preparation and physical parameters

95 Each of the 9 samples was collected with new plastic broom and shovel, and georeferenced. Samples were
96 sieved to achive fractions < 2000 and 63 µm. Particle size distribution of the road dust < 63 µm was
97 determined by the X-ray sedimentation technique with a Micromeritics Sedigraph III Plus grain size
98 analyzer. This technique determined the relative mass distribution of the samples by particle size and is
99 based on two physical principles: sedimentation theory (Stokes' law) and the absorption of X-radiation
100 (Beer-Lambert law). The pH was determined in with a 1:2.5 soil/water solution, using a pH meter, and pH
101 classes were defined according to USDA (1998): extremely acid 3.5 - 4.4, very strongly acid 4.5 - 5.0, highly
102 acid 5.1 - 5.5, moderately acid 5.6 - 6.0, slightly acid 6.1 - 6.5, neutral 6.6 - 7.3, slightly alkaline 7.4 - 7.8,
103 moderately alkaline 7.9 - 8.4, and highly alkaline 8.5 - 9.0. Electrical conductivity (EC) was measured under
104 the same conditions as pH in the two fractions, using a high-resolution conductivity meter. Soils color was
105 determined using Munsell Color soil chart (Munsell, 2009). The color was determined in the two fractions
106 analyzed (< 63 and 2000 µm). Sieving was done at the Pedagogical University of Maputo (Mozambique)
107 and remaining procedures and analysis at GeoBioTec Research Centre/Department of Geosciences,
108 University of Aveiro (Portugal).

109 2.4 Samples and data analysis

110 The chemical composition of samples was accessed by X-ray fluorescence (XRF) spectrometry, using a
111 PANalytical PW 4400/40 45 Axios with Cr Kα radiation. Mineralogical phases were determined by powder
112 X-ray diffraction (XRD) using a Phillips/Panalytical powder diffractometer, model X’Pert-Pro MPD,
113 equipped with an automatic slit. A Cu X-ray tube was operated at 50 kV and 30 mA. Data were collected
114 from 2◦ to 70◦ 2θ with a step size of 1◦ and a counting interval of 0.02 s. The precision and the accuracy
115 of analyses and procedures were monitored using UAVR internal standards, certified reference material
116 (BGS119), and quality control blanks. Results were within the 95% confidence limits. The relative standard
117 deviation was between 5% and 10%.

118 Nemerov index (PN) was calculated for selected PTEs Cr, Cu, Mn, Ni, Pb, Zn, and Zr on < 63 µm fraction
119 samples data, using background soil of Cr = 41.3, Cu= 189, Mn = 83, Ni = 3.4, Pb = 30.2 , Zn = 92 in mg/kg
120 (Bernardo et al., 202b2). The soil content was used in order to identify other sources of dust PTEs
121 enrichment. This index has been used to assess elemental contribution to dust contamination,
122 synthesizing a single pollution index value for all elements selected and being comprehensively reflected
123 on the degree of dust pollution, highlighting the impact of elements content (Xiao et al., 2021). It is
. 𝑛
124 calculated as 𝑃𝑁 = √[(1/𝑛 ∑𝑖=1 𝑃𝐼)2 + (𝑃𝐼𝑚𝑎𝑥 )2 ]/2, where PI is the single element pollution index,

125 PImax the maximum elements PI value, and n the number of elemental species selected. The index classifies
126 pollution as PN ≤ 0.7 non-polluted, 0.7 < PN ≤ 1.0 warning line of pollution, 1.0 < PN ≤ 2.0 low level of
127 pollution, 2.0 < PN ≤ 3.0 moderate level of pollution, and PN > 3.0 high level of pollution (Candeias et al.,
128 2022).

129 Non-carcinogenic and carcinogenic risk assessment was calculated taking in consideration that residents,
130 both children and adults, were directly exposed to dust through three main pathways (a) ingestion, (b)
131 inhalation and (c) dermal absorption. Road dust <63 μm fraction was analyzed because of its higher
132 influence on inhalable materials and on exposure via the gastrointestinal tract, easily resuspended by the
133 wind and can be deposited on human skin, considered to have the highest PTEs adsorption capacity when
134 compared to the <2000 μm fraction. Carcinogenic and non-carcinogenic side effects for each element
135 were computed individually, considering reference toxicity levels for each variable, as extensively
136 described by other authors (e.g., Candeias et al., 2020; RAIS 2022). For each selected PTE (Cr, Cu, Mn, Ni,
137 Pb, Zn, Zr) and pathway, the non-cancer toxic risk was estimated by computing the Hazard Quotient (HQ)
138 for systemic toxicity (i.e., non-carcinogenic risk). If HQ > 1, non-carcinogenic effects might occur once
139 exposure concentration exceeds the reference dose (RfD). The cumulative non-carcinogenic hazard index
140 (HI) corresponds to the sum of HQ for each pathway and/or variable. Similarly, HI > 1 indicates that non-
141 carcinogenic effects might occur. Carcinogenic risk, or the probability of an individual to develop any type
142 of cancer over a lifetime as a result of exposure to a potential carcinogen, was estimated by the sum of
143 total cancer risk for the three exposure routes. A risk > 1.00E−06 is classified as the carcinogenic target
144 risk, while values > 1.00E−04 are considered unacceptable.

145 Results and discussion

146 Dust samples pH, electrical conductivity (EC), and color are presented in Table 1. Samples < 63 µm fraction
147 presented a pH predominantly neutral, except for sample 9, slightly acid, and samples 6 and 8, highly
148 alkaline to slightly alkaline. Sample 9 was collected on a roads intersection of entrance and exit of the
149 urban area, with intense traffic. Samples 6 and 8 were taken on smaller roads with squares, with roads
150 significant degradation, being sample 8 from a more degraded road. Fraction < 2000 µm samples 1, 2, 5,
151 and 9 presented neutral pH, samples 3, 9, and 7 slightly acidic, and samples 6, and 8 highly alkaline. A
152 slightly acidic pH (Sample 3, 9 and 7) has been as favorable for heavy metals retention in soils and being
153 common in solid waste contaminated areas (Campos 2010). Sample 3 was collected near Hulene-B waste
154 dump, revealing and acid pH that reflected the dump influence. Sample 7 was collected in n area with
155 intense traffic and degraded pavement, similar to sample 9, collected on a road with degraded pavement
156 with intense traffic. In general, acidic, and neutral pH was identified in samples collected on degraded
157 pavement roads and close to contamination sources such as, Hulene-B dump and International Airport
158 (samples 1, 2, 3, 4, 5, 7 and 9). Alkaline pH was found in samples located on less degraded roads (6 and 8)
159 and relatively far from previous contamination sources. Previous studies in road dust suggested a
160 prevalence of acidic pH as an indicator of higher PTEs concentration (Nematollahi et al., 2020).

161 Dust samples EC presented higher values in < 2000 µm fraction. Road dust is a complex mixture that
162 includes organic matter (OM), which is pointed out as a factor for EC increase in sandy materials (Kasimov
163 et al. 2019; Pariente et al., 2019). Samples with lower EC values were the ones collected in less degraded
164 roads (samples 4, and 5) and widened roads (sample 8). Samples EC in < 63 µm fraction was ranked 9 > 3
165 > 2 > 6 > 1 > 5 > 4 > 7 > 8, and in fraction < 2000 µm 3 > 9 > 2 > 6 > 1 > 7 > 4 > 5 > 8, being the highest EC
166 found in samples collected in locations with degraded roads, intense traffic, near International Airport,
167 and Hulene-B waste dump. Especially in sample 3, EC results suggested that is an accumulation of
168 materials that rise EC resulting in higher retention of contaminants (Bernardo et al., 2022b). Dust samples
169 color varied slightly in both fractions analyzed (Table 1). Road dust color is influenced by particles origin
170 and composition (Yukioka et al. 2020). In general, darker colors are linked to the presence of organic
171 matter and clay materials, pointed as relevant for adsorption and absorption of PTEs (Aryal et al., 2015;
172 Vlasov et al., 2022). Dust samples granulometric distribution is presented in Table 2. Sand fraction ranged
173 93.2 to 98.3 %, in samples 2 and 7, respectively, with all samples classified as sand. Predominance of the
174 sandy sizes particles is associated with source materials, and surrounding soils (Momade et al., 1996).

175 Identified mineralogical phases of the <2000, and 63 µm fractions, is presented in Figure 2. In all samples,
176 quartz (SiO2) was the predominant mineral, ranked 97.6 > 91.0 > 90.6 > 89.7 > 89.5 > 87.3 > 86.4 > 77.3 >
177 73.7 % in < 2000 µm fraction samples 3, 1, 6, 9, 8, 7, 2, 4, and 5, respectively. In fraction < 63 µm, also
178 quartz was the main mineral identified, ranging 20.6 to 66.6 %. Quartz content, in all samples and
179 fractions, is linked to local soils and transported dusts mineralogical composition (Momade et al., 1996;
180 Bernardo et al., 2022b). Silt fraction revealed smaller quartz content, suggesting that its particles had
181 higher sizes than 63 µm. Quartz presents higher structural hardness than some other minerals, that helps
182 preventing physical weathering and abrasion (Candeias et al., 2022). However, Bai et al. (2021) suggested
183 that, in smaller quartz particles with structural cracks at temperatures > 20 ºC, adsorption of PTEs, such
184 as Pb, can occur. Prolonged contact with high quartz content dust can induce diseases such as silicosis and
185 pulmonary fibrose (Leung et al., Yu; Chen 2012; Koga et al., 2021). Respirable crystalline silica poses a risk
186 to human health when exposed chronically (Horwell et al., 2012). Other minerals identified were,
187 phyllosilicates (mainly micas and kaolinite) detected all samples < 2000 µm fraction, with exception of
188 sample 3, collected near Hulene-B waste dump; K-feldspars (KAlSi3O8), and plagioclase
189 ((Na,Ca)((Si,Al)AlSi2)O8) were detected in all samples < 2000 µm, with highest contents in samples 3 and
190 4. Studies suggested that micas and kaolinite present higher PTEs adsorption capacity (Murray, 2000;
191 Odukoya et al., 2023), which can reach the human body through the inhalation, dermal and ingestion
192 routes, leading to health outcomes (Alghamadi et al., 2023). In the < 63 µm fraction, K-feldspars were
193 detected in considerable amounts in all samples. Carbonates (calcite (CaCO3), dolomite (CaMg(CO3)2) were
194 detected only on the thinner fraction. Sulphates alunite (KAl3(SO4)2(OH)6) and anhydrite (CaSO4) were
195 detected in all samples, with higher alunite concentrations in samples 3, 5 and 2. Oxides magnetite-
196 maghemite (Fe2O3-γFe2O3), ilmenite (FeTiO3), and pyrite (FeS2) were detected in almost all < 63 µm
197 fraction samples, suggesting higher adsorption of Fe oxides in the thinner fractions. Nsutite
198 (Mn4+,Mn2+)(O,OH)2) was detected in sample 7 revealing a higher affinity of Mn oxides. Results showed a
199 heterogeneous mineralogical composition of the sampling areas, with remobilization and deposition of
200 materials transported from other locations (Momade et al., 1996).

201 Independent samples T-test showed significant differences (p < 0.05) between fraction < 2000 µm and <
202 63 µm, with higher concentration in the thinner fraction (Table 3). Previous studies reported that thinner
203 fractions presented higher toxicity, with higher ability to adsorb PTEs (Al-Shidi et al., 2021; Bian et al.,
204 2015; Lin and Wu, 2015). This fraction is also easily resuspended by traffic and wind, allowing
205 incorporation and dispersion in the air, and dermal adherence, by exposed population (Aguilera et al.,
206 2021; Al-Shidi et al., 2021; Candeias et al., 2021).

207 Lowest Cr concentration, in both fractions, was found in sample 9, collected on an intense traffic road,
208 while highest values were detected in samples 3, and 4, collected on degraded asphalt roads near Hulene-
209 B waste dump, and near the airport, respectively. Cr has been associated with intense traffic, asphalt
210 degradation, and wastes incineration that contain paints, varnishes, organic solvents, and oils (Chaudhary
211 et al., 2021). Copper presented similar pattern with sample 3 revealing the higher concentration also on
212 the < 63 µm fraction, above the 27.1 mg/kg guideline (Table 3), while samples 9 and 8 with lower Cu
213 content. This suggests a common enrichment source, e.g., degraded pavements, brakes on bus stops, tires
214 (ZnO and Cu/Zn layers formed during vulcanization). Particulate wear and resuspension have been
215 considered main sources of Zn, and Cu in urban areas (Valotto et al., 2015). Samples with highest Fe
216 concentrations in both fractions were collected near the waste dump (sample 3), and degraded
217 pavements and heavy traffic (samples 4 and 8), suggesting anthropogenic sources of enrichment, when
218 compared to local soils content (Bernardo et al., 2022b). Iron has been linked to metal alloys production,
219 that are commonly used in vehicles components, e.g., main component on steel and associated rust.
220 Samples with higher Mn concentration were collected on roads with less degraded pavement but intense
221 traffic, i.e., National nº 1 road, the main road for entry and exit of Maputo urban area. Mn has been
222 associated with tire and brake wear and is used to prevent corrosion and deformation of car components
223 (Candeias et al., 2020).
224 In all samples, Ni content was higher than guide values of 39.1 mg/kg (Table 3), differences between
225 analyzed fractions may be associated to a higher adsorption capacity by thinner materials (Bisht et al.,
226 2022). Ni has been used in the production of stainless steel and other Ni alloys, giving high corrosion and
227 temperature resistance in automobiles, and in waste electronic equipment, rechargeable batteries, paint
228 cans, varnishes, organic solvents, and waste glass (CCME, 2015), materials that are deposited in Hulene-B
229 waste dump (Bernardo et al., 2022a). Lead higher concentration was found in samples < 63 µm fraction
230 samples 6 and 3, located in areas with intense traffic and degraded asphalt, both near Hulene-B dump
231 and Maputo international airport, considered anthropogenic sources of Pb in urban areas (Bernardo et
232 al.,2023). However, when detected, Pb content was higher than guideline 19.6 mg/kg in samples 1, 3, 4
233 and 6 (Table 3), suggesting differentiated anthropogenic sources. PbO4 is a component used in brake
234 friction materials, and a gasoline additive used in Mozambique until 2005. In samples 2, 7, 8 and 9, Pb was
235 below detection limit, what may be associated with leaching processes and topsoil removal by wind, given
236 the square characteristics and the influence on wind continuous removal of surface dust at these points.
237 Kasimov et al. (2019) presented a similar study in Moscow suggesting that more open areas present a
238 greater capacity to remove and leach out road dust enriched with PTEs, such as Pb.
239 Zn was detected mainly in < 63 µm fraction, being samples 3 and 4 the ones with highest content, collected
240 in degraded asphalts with intense traffic, higher than the guideline (82.4 mg/kg). Zn has been used in road
241 pavement materials, and as an engine oil additive (Sternbeck et al., 2002), suggesting that fuel combustion
242 may be a significant source. Zirconium concentration in urban environments has been linked to vehicle
243 exhaust emissions, as mixed Zr oxides (CeO2/ZrO2) have become an essential component of three-way
244 catalysts (Piatak et al., 2017).
245 Nemerov Index was calculated for < 63 µm fraction, using both soils background (PNbkg), and soils
246 surrounding Hulene-B dump (PNHB) mean concentration (Bernardo et al., 2022b), to assess the
247 contribution of soils to road dust Cr, Cu, Mn, Ni, Pb, Zn, and Zr concentration (Fig. 3). Dust PN using
248 background values revealed higher contamination than when considering Hulene-B surrounding soils,
249 suggesting an anthropogenic contribution of the Hulene-B dump associated with leachates and waste
250 incineration ashes. The high PTEs concentration in road dust, may be linked to the sampling season (June)
251 marked by scarcity of precipitation, pointed out as a factor that contributes to PTEs accumulation (Men
252 et al., 2023; Davoudi et al., 2021). Samples 3, 4 and 7 PNbkg showed the higher index values, revealing the
253 influence of the dump to dust contamination. Sample 7 PNHB was the highest among all samples
254 suggesting the influence of other anthropogenic sources to the PTEs content, e.g., intense traffic, high
255 degradation of Lurdes Mutola Avenue pavement. Sample 9 presented the lowest values in both indexes,
256 which may be associated to resuspension processes, a factor that disperse contaminants (Valotto et al.,
257 2015).
258 Individual contribution of Cr, Cu, Mn, Ni, Pb, Zn, and Zr to the Nemerov index (PN) are presented in Figure
259 4. Nemerov index calculated using background soils (Pnbkg), in samples 1 to 6 contributions were ranked
260 Zn > Ni > Cr >, Cu > Pb > Zr > Mn while in samples 7 and 9 Ni > Cu > Cr > Mn > Zr > Zn. PnHB, that considered
261 Hulene-B surrounding soils, samples 1 to 6 were ranked Zn > Cr > Ni > Pb > Zr > Mn, and in samples 7 to
262 9, Ni > Cu > Cr > Mn > Zn > Zr. Resulted suggested that samples 1 to 6 have varied sources of enrichment
263 by PTEs. Results suggested that samples 1 to 6 have highly contaminated sources of enrichment which
264 was evident in both Pn methods, e.g., studied PTEs high content can be associated with asphalt
265 degradation, fuel, poor oil burning, tire and brake wear (Sternbeck et al. 2002; Valotto et al. 2015). In
266 samples 7 to 9, Pb content was no significant for the index. Previous studies reported that Pb accumulates
267 on road dust (Binh et al., 2021), revealing higher vulnerability to be mobilized in open areas (Kosheleva et
268 al., 2019).

269 Hazard quotient (HQ) was calculated for fraction < 63 µm due to the higher PTEs Cu, Fe, Ni, Pb, and Zr
270 concentration and size fraction (Fig. 5). Hazard quotient, or non-carcinogenic hazard, suggested that
271 ingestion was the main exposure route, and that HQ was higher for children, mostly due to hands-to-
272 mouth predisposition, being dermal contact and inhalation routes considered negligible. Children have a
273 significantly lower body mass and absorb a higher proportion of ingested Pb, when compared to adults,
274 due to dietary Fe and Ca deficiencies (WHO, 2021). Previous studies suggested that women are more
275 exposed to road dust in African cities (Friberg et al., 2021). Wani et al. (2015) reported that pregnant
276 women exposed to high Pb-containing dust may have high levels in blood, which raises the risk of
277 premature birth or low birth weight newborns. In Maputo urban area, women and children do most of
278 the commercial activities alongside sidewalks, increasing the risk due to greater exposure to contaminated
279 road dust (Rogerson, 2017).

280 Carcinogenic risk was only identified in PTEs Ni and Pb (Fig. 6). A higher risk due to Pb content was found
281 in samples 1, and 3 to 6, collected near populated places, including schools. Kazemi et al. (2021), suggested
282 that Pb exposure can induce seizure disorders, aggressive behavioral disorders, chronic abdominal pain,
283 anemia, and cancer. WHO (2021) suggested that children are particularly vulnerable to dust pollution by
284 Pb and estimated that this exposure is responsible for 30 % of the global burden of idiopathic development
285 and intellectual disability. Severe damage to the brain and kidneys, both in adults and children, were
286 linked to exposure to high levels of Pb, resulting in death (Wani et al., 2015). All samples revealed a risk
287 due to Ni below carcinogenic target. Begum et al. (2022) reported that Ni is emitted from inefficient
288 vehicles with difficulties in the total burning of fuel by engines. Zhao et al. (2021) reported that Ni is mainly
289 inhaled in the form of insoluble compounds, and nickel carbonyl (Ni(CO)4) vapors that can induce chronic
290 bronchitis, decreased lung function and cancer. It has been reported that Ni carbonyl enriched dust
291 inhalation can cause severe lung diseases, such as, diffuse interstitial pneumonia and cerebral
292 hemorrhage (Begum et al., 2022). Samples posing a slightly higher risk were collected in locations with
293 diverse commercial activities, such as street food vendors, and with intense road traffic.

294 Conclusions
295 The study of road dust in Maputo urban area revealed that pH, EC, OM and particle size showed
296 characteristics that can be associated to greater absorption capacity and toxicity. Mineralogical analysis
297 revealed the predominance of quartz (SiO2) > phyllosilicates > feldspars (KAlSi3O8)>, plagioclase
298 ((Na,Ca)((Si,Al)AlSi2)O8)>, Carbonates (calcite (CaCO3)>, dolomite (CaMg(CO3)2) > aluminite sulfates
299 (KAl3(SO4)2(OH)6) > anhydrite (CaSO4). Chemical analyses showed a potential health risk related to PTEs
300 concentration, especially in the thinner fraction. Hazard quotient (HQ) was mostly due to Cu, Fe, Ni, Pb,
301 and Zr of the thinner fraction, suggesting a higher probability of occurrence of non-carcinogenic diseases
302 in children. Ni and Pb represented a carcinogenic risk to both children and adults. Results suggested the
303 need for continuous monitoring and the definition of mitigation measures for the protection of the
304 population. This research requires further investigation to determine the bioaccessibility of the PTEs with
305 higher concentrations in the road dusts studied.

306 Author Contributions: Conceptualization, C.C.; methodology, C.C., F.R.; validation, C.C, F.R.; formal
307 analysis, B.B., C.C.; investigation, B.B., C.C., F.R.; writing—original draft preparation, B.B., C.C.; writing—
308 review and editing, C.C., F.R.; supervision, C.C., F.R.; funding acquisition, F.R. All authors have read and
309 agreed to the published version of the manuscript.

310 Funding: This work was partially supported by GeoBioTec Research Centre (UIDB/04035/2020), funded
311 by FEDER funds through the Operational Program Competitiveness Factors COMPETE and by National
312 funds through FCT. The first author acknowledges grant from the Portuguese Institute Camões and FNI
313 (Investigation National Fund—Mozambique).

314 Conflicts of Interest: The authors declare no conflict of interest.

315 References

316 Alghamdi, A.G., EL-Saeid, M.H., Alzahrani, A.J., Ibrahim, H.M., 2022. Heavy metal pollution and associated
317 health risk assessment of urban dust in Riyadh, Saudi Arabia. PLoS ONE. 17, e0261957.
318 https://doi.org/10.1371/journal.pone.0261957

319 Alghamdi, M.A., Hassan, S.K., Al Sharif, M.Y., Khoder, M.I.,Harrison, R.M., 2023. Pollution characteristics
320 and human health risk of potentially toxic elements associated with deposited dust of sporting
321 walkways during physical activity. Atmosp. Poll. Res. 14(1), 101649.
322 https://doi.org/10.1016/j.apr.2023.101649

323 Ali, M.U., Liu, G., Yousaf, B., Abbas, Q., Ullah, H., Munir, M.A.M., Fu, B., 2017. Pollution characteristics and
324 human health risks of potentially (eco)toxic elements (PTEs) in road dust from metropolitan area of
325 Hefei, China. Chemosphere. 181, 111–121. https://doi.org/10.1016/j.chemosphere.2017.04.061.

326 Alves, C.A., Evtyugina, M., Vicente, A.M.P., Vicente, E.D., Nunes, T.V, Silva, P.M.A., Querol, X., 2018.
327 Chemical profiling of PM 10 from urban road dust. Sci. Total Environ. 634, 41–51.
328 https://doi.org/10.1016/j.scitotenv.2018.03.338.

329 Amato, F., Cassee, F.R., Denier van der Gon, H.A.C., Gehrig, R., Gustafsson, M., Hafner, W., Querol, X.,
330 2014. Urban air quality: The challenge of traffic non-exhaust emissions. J. Hazard. Mat. 275, 31–36.
331 https://doi.org/10.1016/j.jhazmat.2014.04.053.

332 Bai, B., Qingke, N., Yike Z., Xiaolong W., Wei H., 2021. Cotransport of Heavy Metals and SiO2 Particles at
333 Different Temperatures by Seepage. Journal of Hydrology 597: 125771.
334 https://doi.org/10.1016/j.jhydrol.2020.125771.

335 Beji, A., Deboudt, K., Khardi, S., Muresan, B., Flament, P., Fourmentin, M., Lumière, L., 2020. Non-exhaust
336 particle emissions under various driving conditions: Implications for sustainable mobility.
337 Transportation Res. Part D: Transport Environ. 81, 102290.
338 https://doi.org/10.1016/j.trd.2020.102290.

339 Bernardo, B., Candeias, C., Rocha, F., 2022a. Soil Risk Assessment in the Surrounding Area of Hulene-B
340 Waste Dump, Maputo (Mozambique). Geosciences 12, 290.
341 https://doi.org/10.3390/geosciences12080290.

342 Bernardo, B., Candeias, C., Rocha, F., 2022b. Soil properties and environmental risk assessment of soils in
343 the surrounding area of Hulene-B waste dump, Maputo (Mozambique). Environ. Earth Sci. 81, 542
344 https://doi.org/10.1007/s12665-022-10672-7

345 Bernardo, B.; Candeias, C.; Rocha, F., 2023. The Contribution of the Hulene-B Waste Dump (Maputo,
346 Mozambique) to the Contamination of Rhizosphere Soils, Edible Plants, Stream Waters, and
347 Groundwaters. Environments, 10, 45. https://doi.org/10.3390/environments10030045

348 Bisht, L., Gupta, V., Singh, A., Gautam, A.S., Gautam, S., 2022. Heavy metal concentration and its
349 distribution analysis in urban road dust: A case study from most populated city of Indian state of
350 Uttarakhand. Spatio-Temporal Epidem. 40, 100470. https://doi.org/10.1016/j.sste.2021.100470.

351 Campos, C., 2010. Soil attributes and risk of leaching of heavy metals in tropical soils. Ambiência 6(3):547–
352 565.

353 Candeias, C., Ávila, P.F., Ferreira da Silva, E., Rocha, F., 2021. Metal(loids) Bioaccessibility in Road Dust
354 from the Surrounding Villages of an Active Mine. 1–13. Atmosphere. 12, 685.
355 https://doi.org/10.3390/atmos1206068.

356 Candeias, C., Vicente, E., Tomé, M., Rocha, F., Ávila, P., Célia, A., 2020. Geochemical, Mineralogical and
357 Morphological Characterization of Road Dust and Associated Health Risks. Int. J. Environ. Res. Public
358 Health. 17, 1563. https://doi.org/10.3390/ijerph17051563.

359 Candeias, C., Ávila, P.F., Sequeira, C., Manuel, A., Rocha, F., 2022. Potentially toxic elements dynamics in
360 the soil rhizospheric-plant system in the active volcano of Fogo (Cape Verde) and interactions with
361 human health. Catena 209. https://doi.org/10.1016/j.catena.2021.105843

362 CCME, 2015. Canadian Soil Quality Guidelines for the Protection of Environmental and Human Health -
363 Nickel. Canadian Council of Ministers of the Environment. https://ccme.ca/en/res/nickel-canadian-
364 soil-quality-guidelines-for-the-protection-of-environmental-and-human-health-en.pdf (accessed
365 on 30 May 2022).
366 CIAT, 2017. Climate-Smart Agriculture in Mozambique. Center for Tropical Agriculture.
367 https://climateknowledgeportal.worldbank.org/sites/default/files/2019-06/CSA-in
368 Mozambique.pdf (accessed on 14 May 2022).

369 Coker, E., Kizito, S., 2018. A Narrative Review on the Human Health Effects of Ambient Air Pollution in Sub-
370 Saharan Africa: An Urgent Need for Health Effects Studies. Int. J. Environ. Res. Public Health 15,
371 427. https://doi.org/10.3390/ijerph15030427.

372 Davoudi, M., Esmaili-Sari, A., Bahramifar, N., 2022. Spatio-temporal variation and risk assessment of
373 polycyclic aromatic hydrocarbons (PAHs) in surface dust of Qom metropolis, Iran. Environ Sci Pollut
374 Res 28, 9276-9289. https://doi.org/10.1007/s11356-020-08863-5

375 Dehghani, S., Moore, F., Keshavarzi, B., Hale, A., 2017. Health risk implications of potentially toxic metals
376 in street dust and surface soil of Tehran, Iran. Ecotox. Environ. Safety. 136, 92–103.
377 https://doi.org/10.1016/j.ecoenv.2016.10.037.

378 Fayiga, A.O., Ipinmoroti, M.O., Chirenje, T., 2018. Environmental pollution in Africa. Environ. Dev. Sustain.
379 20, 41–73. https://doi.org/10.1007/s10668-016-9894-4.

380 Friberg, J., Abera, A., Friberg, J., Isaxon, C., Jerrett, M., Malmqvist, E., Vargas, A.M., 2021 Air Quality in
381 Africa: Public Health Implications. An. Rev. Public Health. 42, 193–210.
382 https://doi.org/10.1146/annurev-publhealth100119-113802.

383 Gascon, M., Rojas-Rueda, D., Torrico, S., 2016. Urban Policies and Health in Developing Countries: The
384 Case of Maputo (Mozambique) and Cochabamba (Bolivia). Public Health Open. J. 1, 24-31
385 https://doi.org/10.17140/PHOJ-1-106.

386 Grigoratos, T., Martini, G., 2015. Brake wear particle emissions: a review. Environ. Sci. Pollut. Res. 22,
387 2491–2504. https://doi.org/10.1007/s11356-014-3696-8.

388 Han, N.M., Latif, M.T., Othman, M., Dominick, D., Mohamad, N., Juahir, H., Tahir, N.M., 2014. Composition
389 of selected heavy metals in road dust from Kuala Lumpur city center. Environ. Earth Sci. 72, 849–
390 859. https://doi.org/10.1007/s12665-013-3008-5.

391 Heo, S., Kim, D.Y., Kwoun, Y., Lee, T.J., Jo, Y.M., 2021. Characterization and source identification of fine
392 dust in Seoul elementary school classrooms. J. Hazard. Mat. 125531, 414.
393 https://doi.org/10.1016/j.jhazmat.2021.125531.
394 Horwell, C.J., Williamson, B.J., Donaldson, Blond, J.SL., Damby, D.E., Bowen, L. (2012). The structure of
395 volcanic cristobalite in relation to its toxicity; relevance for the variable crystalline silica hazard. Part
396 Fibre Toxicol 9, 44. https://doi.org/10.1186/1743-8977-9-44

397 INE, 2020. Boletim de Estatísticas Demográficas e Sociais, Maputo Cidade 2019. Instituto Nacional de
398 Estatistica. http://www.ine.gov.mz/estatisticas/estatisticas-demograficas-e-indicadores-sociais/
399 (accessed on 01 December 2021).

400 IQAir, 2020. World Air Quality Report Region City PM2.5 Ranking.
401 https://www.greenpeace.org/static/planet4-romania-stateless/2021/03/d8050eab-2020-
402 world_air_quality_report.pdf (accessed 20 December 2021).

403 Johansson, C., Norman, M., Burman, L., 2009. Road traffic emission factors for heavy metals. Atmos.
404 Environ. 43, 4681–4688. https://doi.org/10.1016/j.atmosenv.2008.10.024.

405 Joubert, B.R., Mantooth, S.N., Mcallister, K.A., 2020. Environmental Health Research in Africa: Important
406 Progress and Promising Opportunities. Front. Genet. 10, 1–29.
407 https://doi.org/10.3389/fgene.2019.01166.

408 Kalisa, E., Archer, S., Nagato, E., Bizuru, E., Lee, K., Tang, N., Lacap-Bugler, D., 2019. Chemical and Biological
409 Components of Urban Aerosols in Africa: Current Status and Knowledge Gaps. Int. J. Environ. Res.
410 Public Health. 9, 941. https://doi.org/10.3390/ijerph16060941.

411 Karuppasamy, M.B., Natesan, U., Karuppannan, S., Chandrasekaran, L.N., Hussain, S., Almohamad, H.,
412 2022. Multivariate Urban Air Quality Assessment of Indoor and Outdoor Environments at Chennai
413 Metropolis in South India. Atmosphere. 13, 1627. https://doi.org/10.3390/atmos13101627.

414 Kazemi, Z., Hesami Arani, M., Panahande, M., Kermani, M., Kazemi, Z., 2021. Chemical quality assessment
415 and health risk of heavy metals in groundwater sources around Saravan landfill, the northernmost
416 province of Iran. Int. J. Environ. Anal. Chem. 1–19.
417 https://doi.org/10.1080/03067319.2021.1958800.

418 Khan, R.K., Strand, M.A., 2018. Road dust and its effect on human health: a literature review. Epidem.
419 Health. e2018013. 40, 1–11. https://doi.org/https://doi.org/10.4178/epih.e2018013.

420 Kraidy, N., Armel, B., Bi, B., Emile, B., Daniel, A.K., 2022. Distribution and Characterization of Heavy Metal
421 and Pollution Indices in Landfill Soil for Its Rehabilitation by Phytoremediation. J. Geos. Environ.
422 Protect. 10, 151–172. https://doi.org/10.4236/gep.2022.101011.
423 Kumar, P. G., Lekhana, P., Tejaswi, M., Chandrakala, S., 2020. Effects of vehicular emissions on the urban
424 environment- a state of the art. Materials Today Proceedings. 45, 6314–6320.
425 https://doi.org/10.1016/j.matpr.2020.10.739.

426 Leung, C.C., Yu, I.T.S., Chen, W., 2012. Silicosis. Lancet. 379, 2008–2018. https://doi.org/10.1016/S0140-
427 6736(12)60235-9.

428 Li, D., Chen, J., Zhang, Y., Gao, Z., Ying, N., Gao, J., Zhang, K., 2021. Dust emissions from urban roads using
429 the AP-42 and TRAKER methods: A case study. Atmosp. Pollut. Res. 12, 101051.
430 https://doi.org/10.1016/j.apr.2021.03.014.

431 Mboera, L.E.G., Mfinanga, S.G., Karimuribo, E.D., Rumisha, S.F., Sindato, C., Karimuribo, E., Kikuu, C., 2014.
432 The changing landscape of public health in sub-Saharan Africa: Control and prevention of
433 communicable diseases needs rethinking. J. Vet. Res. 81, 1–6.
434 https://doi.org/10.4102/ojvr.v81i2.734.

435 Men, C., Liu, R., Wang, Y. Cao, L. Jiao, L., Li, L., Wang, Y., 2022 Impact of particle sizes on health risks and
436 source-specific health risks for heavy metals in road dust. Environ. Sci. Pollut. Res. 29, 75471-75486.
437 https://doi.org/10.1007/s11356-022-21060-w

438 Muchangos, A., 1999. Paisagens e Regiões Naturais de Moçambique.


439 https://docplayer.com.br/47220681-Mocambique-paisagens-e-regioes-naturais.html (accessed 01
440 on January 2022).

441 Munsell, 2009. Munsell Soil Color Book - Color Charts. Munsell Color Company, Newburgh.

442 Naidja, L., Ali-khodja, H., Khardi, S., 2018. Sources and levels of particulate matter in North African and
443 Sub-Saharan cities: a literature review. Environ. Sci. Pollut. Res. 25, 12303–12328.
444 https://doi.org/10.1007/s11356-018-1715-x.

445 Murray, H.H. 2000. Traditional and new applications for kaolin, smectite, and palygorskite: A general
446 overview. Ap. Clay Sci. 17(5–6), 207–221. https://doi.org/10.1016/S0169-1317(00)00016-8

447 Nematollahi, M.J., Keshavarzi, B., Zaremoaiedi, F., Rajabzadeh, M.A., Moore, F., 2020. Ecological-health
448 risk assessment and bioavailability of potentially toxic elements (PTEs) in soil and plant around a
449 copper smelter. Environ. Monit. Assess. 192, 10. https://doi.org/10.1007/s10661-020-08589-4.
450 Odukoya, K., Olalemi, A.A. 2023. Mineralogy and Geochemical Characterization of Geophagic Clays
451 Consumed in Parts of Southern Nigeria. J. Trace Eleme. Minerals 100063.
452 https://doi.org/10.1016/j.jtemin.2023.100063

453 Perera, F., 2018. Pollution from fossil-fuel combustion is the leading environmental threat to global
454 pediatric health and equity: Solutions exist. Int. J. Environ. Res. Public Health. 15, 1.
455 https://doi.org/10.3390/ijerph15010016.

456 Pio, C., Alves, C., Nunes, T., Cerqueira, M., Lucarelli, F., Nava, S., Querol, X., 2020. Source apportionment
457 of PM 2.5 and PM 10 by Ionic and Mass Balance (IMB) in a traffic-influenced urban atmosphere, in
458 Portugal. Atm. Environ. 223. https://doi.org/10.1016/j.atmosenv.2019.117217.

459 RAIS, The Risk Assessment Information System. Available online: https://rais.ornl.gov (Accessed on 20
460 March 2022).

461 Rwakarehe, E., 2022. Review of Strategies for Curbing Traffic Congestion in Sub-Saharan Africa Cities:
462 Technical and Policy Perspectives. Tanz. J. Eng. Techn. 40, 24–32.
463 https://doi.org/10.52339/tjet.v40i2.730.

464 Sahu, D., Ramteke, S., Dahariya, N.S., Sahu, B.L., Patel, K.S., Matini, L., Hoinkis, J., 2016. Assessment of
465 Road Dust Contamination in India. Atm. Climate Sci. 77–88
466 http://dx.doi.org/10.4236/acs.2016.61006.

467 Serra, C., 2012. Da problemática Ambiental à mudança: rumo à um mundo melhor. Editora Escolar,
468 Maputo.

469 Shahab, A., Hui, Z., Rad, S., Xiao, H., Siddique, J., Liang, L., 2022. A comprehensive review on pollution
470 status and associated health risk assessment of human exposure to selected heavy metals in road
471 dust across different cities of the world. Environ. Geochem. Health. 0123456789.
472 https://doi.org/10.1007/s10653-022-01255-3.

473 Sharma, N., Chaudhry, K.K., Chalapati Rao, C.V., 2004). Vehicular pollution prediction modelling: A review
474 of highway dispersion models. Transport Rev. 24, 409–435.
475 https://doi.org/10.1080/0144164042000196071.

476 USDA. 1998. Soil Quality Indicators: pH. Managing Soils and Terrestrial Systems.
477 https://doi.org/http://soils.usda.gov.
478 Valotto, G., Rampazzo, G., Visin, F., Gonella, F., Cattaruzza, E., Glisenti, A., Tieppo, P., 2015. Environmental
479 and traffic-related parameters affecting road dust composition: A multi-technique approach
480 applied to Venice area (Italy). At. Environ. 122, 596–608.
481 https://doi.org/10.1016/j.atmosenv.2015.10.006.

482 Wani, A.L., Ara, A., Usmani, J.A., 2015. Lead toxicity: A review. Interd. Toxic. 8, 55–64.
483 https://doi.org/10.1515/intox-2015-0009.

484 WHO, 2020. Global air quality guidelines Particulate matter (PM2.5 and PM10), ozone, nitrogen dioxide,
485 sulfur dioxide and carbon monoxide. World Health Organization.
486 https://doi.org/https://www.who.int/publications/i/item/9789240034228 (accessed on 24 May
487 2022).

488 WHO, 2021. Compendium of WHO and other UN guidance on health and environment Chapter 2. Air
489 pollution. World Health Organization. https://cdn.who.int/media/docs/default-source/who-
490 compendium-on-health-and-
491 environment/who_compendium_chapter2_01092021.pdf?sfvrsn=14f84896_5 (accessed on 20
492 May 2022).

493 Zhang, M., Li, X., Yang, R., Wang, J., Ai, Y., Gao, Y., Yu, H., 2019. Multipotential Toxic Metals Accumulated
494 in Urban Soil and Street Dust from Xining City, NW China: Spatial Occurrences, Sources, and Health
495 Risks. Arch. Environ. Cont. Toxic. 76, 308–330. https://doi.org/10.1007/s00244-018-00592-8.

496 Zhao, G., Zhang, R., Han, Y., Meng, J., Qiao, Q., Li, H., 2021. Pollution characteristics, spatial distribution,
497 and source identification of heavy metals in road dust in a central eastern city in China: a
498 comprehensive survey. Environ. Monit. Asse. 193, 12. https://doi.org/10.1007/s10661-021-09584-
499 z.

500

501 Figures captions

502 Figure 1. Sampling locations (yellow dots) and wind directions (adapt. Google Earth, 2022).

503 Figure 2. Dust samples identified mineral phases of the < 63 and < 2000 µm fractions relative distribution (m-m –
504 magnetite maghemite).

505 Figure 3. Dust samples Nemerov Index using soils background (PN bkg), and soils surrounding Hulene-B dump (PNHB)
506 mean concentration (Bernardo et al., 2022a).
507 Figure 4. Relative distribution of elemental contribution for Nemerov Index using soils background (PN bkg), and
508 soils surrounding Hulene-B dump (PNHB) mean concentration (Bernardo et al., 2022a)

509 Figure 5. Dust samples, fraction < 63 µm, hazard quotient by ingestion (HQ ing) in children and adjusted for children
510 and adults by potentially toxic elements and their sum.

511 Figure 6. Dust samples Risk by Pb, Ni and combination of both.

512

513 Tables captions

514 Table 1. Dust samples < 63 and 2000 µm fractions pH, electrical conductivity (EC; μS/cm) and color.

515 Table 2. Dust samples granulometric distribution (in %).

516 Table 3. Dust samples potentially toxic elements chemical composition of the < 63 and 2000 µm fractions (in
517 mg/kg).
Figure1 Click here to access/download;Figure;Figure 1.docx
Figure2 Click here to access/download;Figure;Figure 2.eps
Figure3 Click here to access/download;Figure;Figure 3.eps
Figure4 Click here to access/download;Figure;Figure 4.eps
Figure5 Click here to access/download;Figure;Figure 5.eps
Figure6 Click here to access/download;Figure;Figure 6.eps
Table1 Click here to access/download;Table (Editable version);Table
1.docx

pH EC color
ID
< 63 µm < 2000 µm < 63 µm < 2000 µm < 63 µm < 2000 µm
1 6.86 7.16 251 425 Light brownish gray Brown
2 6.80 7.09 292 660 Grayish brown Grayish brown
3 6.52 6.35 310 1051 Dark gray Gray
4 7.13 7.48 130 224 Gray Pale brown
5 7.38 6.95 139 218 Dark gray Gray
6 8.62 10.41 252 610 Gray Gray
7 6.97 6.41 128 362 Dark gray Grayish brown
8 7.70 8.82 106 180 Dark gray Gray
9 6.28 6.50 521 930 Gray Grayish brown
Table2 Click here to access/download;Table (Editable version);Table
2.docx

1 2 3 4 5 6 7 8 9
sand 95.54 93.22 94.47 94.95 97.33 97.50 98.34 94.51 94.58
silt 2.67 4.38 3.27 3.04 1.36 1.54 0.80 2.34 3.33
clay 1.79 2.40 2.26 2.02 1.31 0.96 0.86 3.15 2.09
Table3 Click here to access/download;Table (Editable version);Table
3.docx

ID Cr Cu Fe Mn Ni Pb Ti Zn Zr
< 63 µm
1 10400 1210 304020 5030 1040 1360 39450 1080 5860
2 4420 1130 246970 3940 690 nd 46770 1850 4850
3 19660 2750 321950 4160 1320 2240 38680 2320 3500
4 19850 1970 317010 7800 900 2240 42760 2340 7240
5 11090 1620 311600 6340 1410 1320 34540 2160 4860
6 12880 1470 306410 6220 1090 4450 38640 1350 4160
7 5510 1520 385810 8270 3540 nd 36810 420 830
8 2770 800 332150 5130 1260 nd 53440 nd 440
9 2590 840 196420 3280 690 nd 47540 nd 800
< 2000 µm
1 2560 370 102580 1220 740 nd 15410 nd 300
2 2450 450 118330 1600 760 nd 30620 420 2300
3 7820 1070 138060 1270 600 nd 18140 390 1240
4 5510 870 135230 2990 570 720 23250 nd 1370
5 3230 730 84400 1520 620 nd 12630 380 1650
6 3140 510 76700 1720 nd 1360 16520 nd 1150
7 1200 nd 65530 1720 nd nd 15850 nd 870
8 6520 1010 121730 1530 580 nd 18370 620 2940
9 980 430 90660 1490 nd 830 21040 530 3770
GV 75.8a 27.1a 12600a 615.6a 39.1a 19.6b 1618a 82.4a 219.0c
nd – not detected; GV – guide values; aAli et al., 2017; bZhao et al., 2021;
cZhang et al., 2019.
Maputo urban area (Mozambique) road dust characterization and risk
assessment
Bernardino Bernardo1,2, Carla Candeias1, Fernando Rocha1

1
GeoBioTec Research Centre, Department of Geosciences, University of Aveiro, 3810-193 Aveiro, Portugal

2
Faculty of Earth Sciences and Environment, Pedagogic University of Maputo, Mozambique

Figure S1. Informal trade in the vicinity of the sample areas.

(a) (b)
(c ) (d)

(f)
(e )

(g ) (h)

(i)

Figure S2. Sampling surrounding environment: (a) sample 1; (b) sample 2; (c) sample 3; (d) sample 4; (e)
sample 5; (f) sample 6; (g) sample 7; (h) sample 8; (i) sample 9.

Você também pode gostar