You are on page 1of 24

Engineering Geology 275 (2020) 105742

Contents lists available at ScienceDirect

Engineering Geology
journal homepage: www.elsevier.com/locate/enggeo

The effects of vegetation traits and their stability functions in bio-engineered T


slopes: A perspective review

Sanandam Bordoloi, Charles Wang Wai Ng
Department of Civil and Environmental Engineering, the Hong Kong University of Science and Technology, Hong Kong Special Administrative Region

A R T I C LE I N FO A B S T R A C T

Keywords: Bio-engineered slopes use vegetation as “live” protection elements against the triggering forces of landslides,
Landslides erosions and debris flows. In this paper, the effects of basic plant traits (both root and shoot) on their slope
Root reinforcement stability functions in compacted and planned green slopes have been reviewed. This review is based on over 200
Shallow landslides papers on areas primarily related to unsaturated soil mechanics, plant biology, hydrology and ecology. The first
Transpiration induced suction
section of this paper gives an overview of the plant-soil–water continuum, highlighting the key interactions that
Vegetation
Soil erosion, Green infrastructure
determine reinforcing actions of vegetation in sloped soil. The hydraulic pathway (water flow from roots to
atmosphere) and plant response at varying soil stress (suction) is explained. Thereafter, the effects of inherent
plant properties (bio-polymer composition, plant physiological parameters and root morphology) on soil re-
inforcement are discussed. The second section discusses the three stability functions (i.e., mechanical, hydro-
logical and interception reinforcements) of vegetation in slopes. The first two subsections considering me-
chanical and hydrological reinforcement action are catalogued. In the third subsection, the interception
functions of plant (both shoot and root parameters) against erosion and debris flow are illustrated based on
experimental, numerical and field observations. In the last section, the knowledge gaps on the topic of bio-
engineered slope stability are highlighted such as the consideration of plant age effect, plantation strategy and
high suction response of vegetation. The possible negative effects of vegetation towards slope stability resulting
from extreme climate conditions such as extended droughts, forest fires, freeze–thaw cycles and elevated CO2
concentration levels are discussed. The review identifies new research themes for developing futuristic bio-
engineered slopes. The potential of real time monitoring and maintenance of bio-engineered slopes, redefinition
of wilting point and concept of “plant sensors” are also put forth.

1. Introduction (Kim et al., 2017). Debris flow is defined as a very rapid movement or
flow of non-plastic debris (boulder, rock and aggregate) in steep terrain
The effects of global climate change result in cycles of extreme (Hürlimann et al., 2019). Concentrated flow erosion is the gradual
precipitation which bring upon adverse impact on both natural and detachment and displacement of soil particles by concentrated water
man-made slopes (Crozier, 2010; Gariano and Guzzetti, 2016). Ha- flow (Vannoppen et al., 2015). For many centuries, vegetation (both
zardous geological processes related to slopes, such as landslides (Korup living or dead mass) had been utilized to stabilize hazardous geological
et al., 2012), concentrated flow erosion (Knapen et al., 2007) and debris processes as reported for ancient Roman and Chinese civilizations
flow (Prieto et al., 2018) are frequently reported. United Nations En- (Smith and Snow, 2008; Partov et al., 2016). The design and utilization
vironment Program (UNEP) annual report (Seneviratne et al., 2017) of these green infrastructures were basically empirical in nature. With
forecasted that sloped soil mass is susceptible to increased frequency in advent of industrialization in the past century, synthetic reinforcement
shallow landslides, granular material flow and erosion, due to extreme measures such as cementitious cover, metal piling and geo-synthetics
precipitation events. Landslide is the movement of soil mass down a were utilized to mitigate adverse geological processes (Koerner, 2012).
slope, under gravitational force. (Van Westen et al., 2003). Approxi- The mechanical properties of such synthetic reinforcement materials
mately, a fifth of the world’s land surface (Hong et al., 2007) is sus- are highly controllable and predictive. In today’s age, with dwindling
ceptible to rainfall induced shallow landslides, resulting in annual resources, climate change realization and concern for sustainability,
fatality of 4500 and property damages up to 3.2 U.S. billion dollars people are actively seeking eco-friendly and green solutions for


Corresponding author.
E-mail addresses: sanandam@ust.hk (S. Bordoloi), cecwwng@ust.hk (C.W.W. Ng).

https://doi.org/10.1016/j.enggeo.2020.105742
Received 24 April 2020; Received in revised form 19 June 2020; Accepted 20 June 2020
Available online 01 July 2020
0013-7952/ © 2020 Elsevier B.V. All rights reserved.
S. Bordoloi and C.W.W. Ng Engineering Geology 275 (2020) 105742

Fig. 1. Schematic representation on evolution and effect of plant properties, its interaction with soil-atmosphere and resultant effect on vegetated soil properties.

addressing engineering problems. 2015). In fact, the hydrological reinforcement provided by plants has
The current trend towards sustainability has gradually motivated not been comprehensively reviewed, which was explored and quanti-
governments and engineers to rediscover vegetation as an engineering fied in the past decade. The inherent role of plant traits such as root
solution for slope protection. The special report by Intergovernmental architecture, canopy type, bio-polymer composition and microbial in-
Panel on Climate Change (IPCC, 2019) has urged policy makers to in- teraction on slope protection measures have not been reviewed. Plants
crease vegetation plantation in order to combat global CO2 levels. Soil being diverse in species, it is rather prudent to comprehend the effect of
bioengineering using vegetation in compacted soil slopes was explored their basic traits with slope stability functions, especially in the context
in the past decades (Gray and Sotir, 1996; Simon and Steinemann, of compacted and planned green slopes. Furthermore, the limitations of
2000). Soil bioengineering provides an eco-friendly, economic and using vegetation as well as the negative effects of plants in slope sta-
aesthetically pleasant measure for shallow slope stability and surficial bility functions were majorly neglected. This, is especially relevant
erosion. Although, the role of vegetation in ecological restoration, where climate change has recently exposed the slopes to extreme events
forested slopes and riverbank protection has been reviewed in the past such as fires, increase in droughts and acid rain. The complex interac-
(Stokes et al., 2009; Pawlik, 2013; Hubble et al., 2017), these studies tion between plant-soil-atmosphere and their competing effects on
mostly consider natural slopes. Moreover, they address singular func- plant growth make the subject matter highly interdisciplinary (Blight,
tion of vegetation (i.e. root reinforcement) on either slope stability or 1997). This study attempts to provide a state-of-the-art on the role of
erosion protection (De Baets and Poesen, 2010; Vannoppen et al., vegetation in engineered slope stability by reviewing the effect of

2
S. Bordoloi and C.W.W. Ng Engineering Geology 275 (2020) 105742

Fig. 2. Schematic illustration of (a) “Soil-Water-Plant-Atmosphere Continuum” in context of an engineered vegetated slope; (b) Hydraulic pathway through a plant
during transpiration.

inherent plant traits/parameters with that of its slope stability and reinforcement, hydraulic reinforcement and interception. The me-
protection functions. The work provides the reader a simple and holistic chanical reinforcement deals with stress–strain response of rooted soil
understanding of stability functions provided by plants in context of whereas the hydraulic reinforcement deals with increase in negative
engineered vegetated slopes. Majority of the discussion in this review is pore water pressure (i.e. suction) due to transpiration. Interception
from the perspective of unsaturated soil mechanics and material sci- functions include protection given by vegetation against particle de-
ence, while the role of plant parameters is discussed in parallel. tachment, debris damping and canopy interception. Based on the
The discussion in this review paper was divided into three major abovementioned sections, the third section highlights new gap areas
sections. In the first major section, the plant-soil-atmosphere continuum and future scope in engineered vegetated slopes. At first, the negative
was put forward. This section is sub-divided by initially explaining the effects of vegetation towards slope stability accounting extreme climate
hydraulic pathway through which plants survive, grow and induce conditions were put forward as one of the major gap areas in the topic.
changes to soil moisture. The second sub-division identifies basic plant Second, plantation strategy based on plant trait type, spacing effect,
traits (bio-polymer composition, root architecture, canopy, stomatal tree pruning, capillary barrier usage was highlighted. Third, usage of
conductance) and juxtaposes their role in slope stability applications. smart monitoring approaches to maintain vegetation in slopes and
The second major section provides a detailed overview of the major concept of “plant sensors” were put forth. The fourth subsection dis-
stability functions provided by plants in slope. i.e. mechanical cusses biomaterials and nature-based alterations that could boost plants

3
S. Bordoloi and C.W.W. Ng Engineering Geology 275 (2020) 105742

traits for slope protection measures. their corresponding effects on slope stability parameters. Precipitation
The literature about effect of vegetation traits on slope stability is results in buildup of pore water pressure (due to infiltration), increases
broad and interdisciplinary. The authors focus primarily on the pub- particle detachment and initiates debris flow due to concentrated sur-
lications having keywords “vegetated slope”, “bio-engineered slopes”, face runoff (Matsuura et al., 2008). Evapotranspiration is favorable to
“soil suction”, “unsaturated soil”, “root reinforcement”, “transpiration”, slope stability as it builds up negative pore water pressure (suction) and
“hydrological reinforcement”, “erosion”, “landslide”, “debris flow” and consequently decreases the hydraulic conductivity of vegetated soil (Ng
“drought”. More than 200 works on the keywords given above have et al., 2019a). However, based on soil type, excessive evapotranspira-
been reviewed considering different soil types, plant species, plant tion is detrimental, as it leads to the formation of desiccation cracks
parameters, slope angle, analysis type, plant age effect, spacing and (Cheng et al., 2020) which are precursors to slip surface development
root depth profiles. The authors limit the review to peer-reviewed pa- during precipitation events (Hobbs et al., 2002). Freezing and thawing
pers in scientific journals, technical conference proceedings, peer re- are associated with decrease in plant transpiration, formation of surface
viewed book chapters on relevant topic and exclude any grey literature. cracks and change in soil retention due to microstructural arrangement
These studies include laboratory investigation, field monitoring and (Wynn and Mostaghimi, 2006; Zhou et al., 2011). Excessive evapo-
numerical assessment of vegetated soil. The scientific databases utilized transpiration with no natural water recharge for plants and virgin
for the search in this review paper were “Science Direct”, “Science freezing of non-resilient native plants may lead to plant death and result
Citation Index Expanded”, “Google scholar”, “Mendeley” and “Scopus”. in root decay (Johnson, 2018).

2. Plant-soil–water interaction 2.1. Hydraulic pathway of plants

Fig. 1 provides a concentric framework on the evolution and effect Vascular plants (trees, grass and shrubs) use vascular tissues such as
of basic plant properties, its interaction with soil water status and re- xylem and phloem to transport water from the vicinity of the roots to
sultant effect on vegetated soil properties. These properties determine the atmosphere through the plant leaves. In bio-engineered slopes,
the stability of a vegetated slope. Stomatal conductance (SC), root type mostly vascular plants are used as they have water absorbing and
and photosynthesis rate govern the hydraulic pathway of a plant transportation mechanisms, which facilitate both hydrological and
through which water is transferred from the soil to atmosphere (Wong mechanical reinforcement to soil (Ng et al., 2019a). On a global scale,
et al., 1979; Brodribb, 2009). In Section 3, detailed discussion is done vascular plants transport back more than half of the 1,10,000 km3 y−1
on how induced suction due to this hydraulic pathway affects the sta- of precipitation which falls on land each year (Chahine, 1992). Water
bility of vegetated slopes. Bio-polymer distribution and root type de- moves from root hairs in soil through xylem tissues in the plant to at-
termine the mechanical reinforcements given by plants underneath the mosphere through leaf stoma (Fig. 2b), along a continuum of increas-
soil surface (Abdi et al., 2014; Genet et al., 2005). The plant properties ingly negative water potential or suction (ψ) (Aroca et al., 2012). Plants
are also governed by the soil and atmospheric conditions. These con- synthesize carbon-based polymers for growth through a process called
ditions can be categorized as soil density effect, pH and salt sensitivity, “photosynthesis”. Photosynthesis depends on CO2 intake from atmo-
drought tolerance, sunlight exposure, wind effects, CO2 concentration sphere and water up taken from roots. Photosynthesis happens in
and mycorrhiza biota effects (Garg, 2015; Garg and Ng, 2015; Ng et al., leaves, wherein CO2 is ingested from atmosphere, while at the same
2019a,b). Engineered slopes are generally compacted, have limited time, water is lost through their stomata (Lefi et al., 2004). Stomata
water recharge sources and are generally in the vicinity of industry or (Fig. 2b) are attributed to being pressure regulators of a plant; wherein
urban spaces (Chatterjea, 2011; Tardío and Mickovski, 2016). These they limit xylem and tissue water pressure from reaching damaging
urban spaces are exposed to pollutants, high CO2 concentration and values by regulating water flow through soil–plant hydraulic pathway
limited sunlight. For landfill slopes and mining sites, vegetation in (Jackson et al., 2000). Xylem cavitation, xylem anatomy and root ar-
slopes interact with extreme pH conditions, lower mycorrhiza con- chitecture influence leaf water supply and plant water usage (Linton
centration and are exposed to methane gas (Börjesson et al., 2004; Chen and Nobel, 2001). Sequential water movement (soil-root-xylem-leaf)
et al., 2016). Survival and growth of vegetation may be hindered or under increasingly negative pressure happens due to capillary forces as
accelerated, depending upon the exposed condition. All the aforemen- per the cohesion-tension theory (Steudle, 2001; Novick et al., 2016).
tioned plant traits and their interaction affect the mechanical para- Capillarity dominates sap-water flow due to capillary radius (Fig. 3a) in
meters (shear strength, soil aggregation, erodibility, desiccation the xylem being smaller for stem as compared with deeper roots
cracking) and hydraulic parameters (soil water retention, porosity, in- (Jackson et al., 2000). Cavitation may occur in the xylem when water
filtration, gas flow) of rooted soil in slopes. The next sub-section dis- tension exceeds atmospheric pressure, wherein sap water vaporizes
cusses in detail some of the basic plant traits and their relevancy to locally so that the vessel elements are filled with water vapor (Cochard
slope stability. et al., 1996; Poggi et al., 2007). Excessive cavitation resists sap water
The “soil–water–plant–atmosphere” continuum in context of vege- flow and reduces the photosynthesis yield of plants leading to tree
tated slope is presented in Fig. 2a. Vegetated slopes are exposed to the mortality (Fig. 3b). Cavitation and consequent wilting upon drought
allied processes of atmosphere such as evapotranspiration, precipita- stresses are significant in context of vegetated slope stability, as they
tion, radiant energy and wind action. Evapotranspiration is the sum of gradually decrease the plant transpiration. Furthermore, root decay
total loss of water due to evaporation from the soil surface and tran- upon wilting can lead to preferential flow of water in the slope through
spiration through the plant leaves (Allen et al., 1994). Shallow land- decayed roots which can exacerbate the development of concentrated
slides in vegetated slope happen progressively by the formation of slip pore water pressure (Mitchell et al., 1995; Perillo et al., 1999). Al-
surfaces during precipitation events of high intensity and magnitude though unlikely, water logging in localized sections of the vegetated
(Ciurleo et al., 2017). Slip surface is defined as the surface plane along slope reduces transpiration and can even lead to plant death due to
which a soil mass in a slope slides or detaches (Morgenstern and Price, limited dissolved oxygen and oxidized nitrogen (Drew, 1997). Feddes
1965). The slip surfaces can be exacerbated by local faults and dis- (1982) proposed the popular transpiration reduction function (TRF)
continuities such as surface desiccation cracks, root decay, preferential which relates soil matric suction with plant transpiration rate and is
erosion paths and fauna intrusion. Slope failure in vegetated soil can incorporated as a sink term within the soil water transfer equation
also be triggered due to combined effects of tree uprooting upon (Richards, 1931). Four major suction points are defined in the tran-
windstorm along with high precipitation (Abe and Ziemer, 1991). spiration reduction function (TRF), namely, (i) Anaerobiosis point (ψ1),
Fig. 2a showcases the cyclic events (precipitation-evapotranspiration Maximum transpiration point (ψ2 ), Field capacity point (ψ3 ) and Wilting
and freezing-thawing) that a vegetated soil generally undergoes and point (ψ4 ). The schematic relationship of the TRF with SC is shown in

4
S. Bordoloi and C.W.W. Ng Engineering Geology 275 (2020) 105742

Fig. 3. (a) Capillary pore size variation along the xylem tissue (Jackson et al., 2000); (b) Schematic illustration of plant basic trait and plant hydraulic phenomenon
with soil matric suction.

Fig. 3b. Optimum functioning of the hydraulic pathway helps in plant systems are also found in grass species (Ghestem et al., 2011). Re-
growth through photosynthesis and induces suction, conducive to slope gardless of the species type, finer roots assist in imparting shear
stability. strength of the slope by forming a complex fibrous network at shallow
soil depth, while broad roots generally act as anchorage elements
resting upon the hard bedrock (Mao et al., 2013). A higher ratio of finer
2.2. Basic plant traits and their relationship with slope stability roots to broad roots by volume assist in water and nutrient uptake as
reported in literature (Haling et al., 2013). It was generally believed
Fig. 4a presents a “Ying and Yan” presentation of the inter- that smaller the root diameter, higher is the root tensile strength ob-
dependent soil-root interaction that determines the mechanical strength served (Mao et al., 2013, 2018, Mahannopkul and Jotisankasa, 2019).
of rooted soil in slopes. The parameters mentioned and their inter- The variation in tensile strength (Tr) with root diameter (dr) was ex-
dependency are systematically discussed in this section. Vascular plants pressed as per the commonly adopted power decay law (Eq. (1)) to
are composed of shoot biomass above ground and root biomass below measure the so-called “root cohesion” (De Baets et al., 2008).
ground. Both shoot and root architecture are based on interdependent
factors such as plant family, soil type and temperature tolerance Tr = k1dr −k2 (1)
(Lynch, 1995; López-Bucio et al., 2003). Trees develop different root
systems such as tuft root system, taproot system and heart shaped roots. where k1 and k2 are empirical fitting co-efficients, quantitatively posi-
Taproot systems with varied lateral branching or uniform fibrous tive and species dependent. The next major section and Table 1

5
S. Bordoloi and C.W.W. Ng Engineering Geology 275 (2020) 105742

Fig. 4. Schematic illustration of (a) soil-root interaction in terms of basic parameters that govern mechanical strength of rooted soil, (b) Effect of root architecture in
slopes on preferential flow during precipitation (after Ghestem et al., 2011).

extensively discuss different existing models on rooted shear strength composition i.e. proportion of cellulose, hemicellulose and lignin.
which incorporate this power relation. Boldrin et al. (2017a) conducted Cellulose is the primary strength contributing bio-polymer component
uni-axial root tensile strength tests on 10 species with different root and is constituted of polymer chains of glucose units linked together by
diameter, native to temperate Europe. The result highlight that Tr re- highly resistant H-bonds (Khalil et al., 2015). Hemicellulose are poly-
lationship with dr does not necessarily follow the popular power decay saccharides of plant cell walls, having low molecular weight. They
law, and obeys three different fitting relationships, i.e. positive power impart secondary strength to the root fiber by forming cross linkage
decay trend, bi-modal trend and negative power decay trend. This between cellulose fibril aggregate and in turn also form complexes with
variability with root diameter was explained by the differences between lignin (Delmer and Amor, 1995). Cellulose polymer chains group to-
root primary and secondary structures in the tested samples. Primary gether in a hemicellulose matrix to collectively form a micro fibril,
structures are composed of high volume of young cortex tissues (which which then arrange in a helical structure to form a macro fibril. The
have less amount of cellulose and hemicellulose biopolymer) (Gregory angle at which these micro fibrils arrange with respect to the lumen is
et al., 2009). On the other hand, Tr was observed to be higher in roots called micro fibrillary angle (MFA). In plant science, lumen is defined as
(regardless of age) which had higher secondary xylem tissues (thick a continuous aqueous phase enclosed by the thylakoid membrane
lignified cell walls) and cortex was relatively lower. which plays an essential role for photophosphorylation during photo-
Root tensile strength is directly related to the inherent bio-polymer synthesis (Kieselbach and Schröder, 2003). Lower MFA leads to higher

6
Table 1
Chronological overview on mechanical reinforcement of vegetated slope.
Reference Species* Soil type** Analysis type* Root parameter Slope Consideration for ΔS FS
considered* gradient
Root Root Tensile Vegetation Mucilage/
branching strength spacing Mycorrhizae effects
S. Bordoloi and C.W.W. Ng

Chok et al. (2004) NA Sand Numerical (FEM, LEA) NA 26° x x x x x ✓


Pollen and Simon (2005) Multiple species (12) Silt Field and lab RAR NA x ✓ x x ✓ ✓
Baets et al. (2008) Multiple species (25) Multiple Soil Theoretical and lab test RAR x ✓ ✓ x x ✓ x
types
Danjon et al. (2008) Quercus alba Sand Lab and numerical RAR 12-18° ✓ ✓ ✓ x ✓ ✓
(LEA)
Fan and Su (2008) Sesbania cannabina Sand Field (in-situ shear test) RAR, RER x x ✓ x x ✓ x
Genet et al. (2008) Cryptomeria japonica Clay soil Field, Lab, Numerical RLD, RAR (33-35°) x ✓ ✓ x ✓ ✓
Hales et al. (2009) Multiple species (15) Sandy-silty loam In-situ and lab RAR, Root cellulose NA x ✓ x x ✓ x
content
Pollen-Pollen-Bankhead and Multiple species (16) Silt Field, lab and numerical RAR NA x ✓ x x ✓ ✓
Simon (2009)
Preti and Giadrossich (2009) Spartium junceum Clay-silty Field RAR 50° x ✓ x x ✓ ✓
Fan and Chen (2010) Multiple species (5) Sand and clay Field (in-situ shear test) RAR, RER x ✓ ✓ x x ✓ x
Lin et al. (2010) Makino bamboo Silty clay, Clayey In-situ test and x 50-70° ✓ ✓ x x ✓ ✓
silt numerical modelling
Sonnenberg et al. (2010) Willow Silty sand Centrifuge modelling x 330 x ✓ x x ✓ ✓
Thomas and Pollen-Bankhead Multiple species (22) NA Sensitivity analysis Root orientation factor NA ✓ ✓ x x ✓ x
(2010)
Zhang et al. (2010) Robinia pseucdoacacia Loess Lab RAR NA x ✓ x x ✓ x
Burylo et al. (2011) Multiple species (6) Weathered Lab test (Coulomb RAR x x ✓ x x ✓ x
regolith envelope)

7
Fattet et al. (2011) Vernicia fordii, Artemisia Clay loam Field (in-situ shear test) RLD x x ✓ x x ✓ x
codonocephala
Ali et al. (2012) Mature lime tree Boulder clay Numerical (LEA) NA 20° x ✓ x x x ✓
Ghestem et al. (2014) Ricinus communis, Jatropha curcas, Alluvial silty clay Lab Root number, NA ✓ ✓ x x ✓ x
Rhus chinensis Branching density
Haling et al. (2014) Hordeum vulgare Sand Lab Shoot dry mass, Root NA x ✓ x ✓ x x
hair length
Mao et al. (2014) Multiple species Coarse sand 3-D numerical x 25-45° x ✓ ✓ x x ✓
modelling
Yildiz et al. (2015) Alnus incana, Trifolium pratense, Poorly graded Laboratory x 30° x x x ✓ ✓ x
Poa pratensis sand
Liang et al. (2015) NA Silica sand Dynamic centrifuge RAR, Root geometry NA x ✓ x x ✓ x
modelling
Veylon et al. (2015) Jatropha curcas, Rhus chinensis, Silty clay Lab RAR NA x ✓ x x ✓ x
Ricinus communis
Chiaradia et al. (2016) Spruce, chestnut, beech tree Umbrisols, Lab and numerical RAR, RLD 20-50° ✓ ✓ x x ✓ ✓
Cambisols (LEA)
Leung et al. (2017) NA CDG Centrifuge and RAR, Root geometry 45-60° ✓ ✓ x x x ✓
numerical
Boldrin et al. (2017a) Multiple species (10) Sandy loam Lab test RSR, RLD x x x x x x x
Das et al. (2017a) Cynadon Dactylon (CD), Schefflera CDG Field and Probabilistic NA 33° x x x x x ✓
heptaphylla (SH) Modelling
Das et al. (2017b) CD, SH CDG Field and Probabilistic NA 33° x x x x x ✓
Modelling
Demenois et al. (2017) Costularia arundinacea, Tristaniopsis Sandy loam LAB RLD, RMD, Specific x x x x ✓ ✓ x
glauca, Arillastrum gummiferum root length
Kim et al. (2017) Multiple species Multiple soils Meta-analysis and NA NA x ✓ x x x ✓
Numerical
(continued on next page)
Engineering Geology 275 (2020) 105742
Table 1 (continued)

Reference Species* Soil type** Analysis type* Root parameter Slope Consideration for ΔS FS
considered* gradient
Root Root Tensile Vegetation Mucilage/
branching strength spacing Mycorrhizae effects

Liang and Knappett (2017a) NA Sand Dynamic centrifuge Root geometry NA x ✓ x x x ✓


S. Bordoloi and C.W.W. Ng

modelling
Liang and Knappett (2017b) NA Sand Dynamic centrifuge RAR, Root geometry 57° x ✓ x x x ✓
modelling
Liang et al. (2017) Willow, Gorse and Festulolium Sandy loam Centrifuge modelling RAR, RLD, Root NA x ✓ x x ✓ x
grass biomass
0
Zhu et al. (2017) NA Sand Probabilistic RLD 37 x x x x x ✓
Boldrin et al. (2018) Corylus avellana, Ulex europaeus Sandy loam Lab test (Penetration Root biomass, RLD x x ✓ x x ✓ x
strength)
Chen et al. (2018) Arabidopsis thaliana CDG Lab test Plant biomass, fungal x x ✓ x ✓ ✓ x
colonization
Das et al. (2018) CD, SH CDG Field and Probabilistic RAI 30-50° x x x x x ✓
Modelling
Demenois et al. (2018) Costularia arundinacea, Tristaniopsis Sandy loam Field RLD, RMD, Percentage x x x x ✓ ✓ x
glauca, Arillastrum gummiferum finer
Liang et al. (2019) NA Silica sand Dynamic centrifuge RAR NA ✓ ✓ x x x ✓
modelling
Ni et al. (2019b) Schefflera arboricola CDG Field and lab RAR NA x ✓ ✓ x ✓ x
Yildiz et al. (2020) Multiple species Clayey sand Lab Root biomass NA x x x ✓ x x

* Not applicable (NA), Root area ratio (RAR), Root cohesion (Cr), Leaf area Index (LAI), Root-shoot ratio (RSR), Root length density (RLD), Completely decomposed granite (CDG), Limit Equilibrium Approach (LEA),
Root mass density (RMD), Root efficiency ratio (RER), Root area Index (RAI), ΔS = Increase in shear strength for vegetated soil, Factor of safety (Fs).

8
** Soil classification as per Unified Soil Classification System (USCS) or U.S. Department of Agriculture (USDA) system.
Engineering Geology 275 (2020) 105742
S. Bordoloi and C.W.W. Ng Engineering Geology 275 (2020) 105742

root tensile strength and stiffness, while a higher MFA results in higher reinforcements against rain-induced landslides. Root length density
root ductility (Hearle and Morton, 2008), regardless of cellulose con- (RLD) is defined as the total length of roots per unit volume of soil. RLD
tent. Lignin, a phenolic polymer, with high molecular weight imparts in indicative of the soil volume explored by a root system in search of
function for mechanical support, water uptake through the xylem and soil water and nutrients (Aziz et al., 2017). Specific leaf area is defined
resists biodegradation (Reddy and Yang, 2005; Khalil et al., 2015). as the ratio of leaf area a plant builds with a given amount of dry leaf
Lignocellulose materials experience volume change with respect to biomass (Boldrin et al., 2017a). Its unit is in m2.kg−1. The induced
their moisture status (Bordoloi et al., 2017). During a drying cycle, suction with increase in the aforementioned plant traits are shown in
when plant roots are water stressed (refer Fig. 3b) they shrink in size Fig. 5b. Ng et al. (2019a) have shown that increase in leaf area index
(within the rhizosheath) and result in preferential flow during sub- (LAI) and root area index (RAI) of S. heptaphylla resulted in an almost
sequent rainfall events (Ghestem et al., 2010). The rhizosheath can be linear increase in retained suction (Fig. 5c) which restrict water ingress
defined as the soil layer that adheres to the plant roots due to the for- during rainfall events. The transpiration efficiency increases 6-fold with
mation of root hairs and mucilage. It essentially helps in improving higher SC (25–175 mmol m−2 s−1) of the species, based on measure-
water and nutrient uptake (Basirat et al., 2019). The orientation of roots ments of 10 different species (Fig. 5d). However, very high SC of Vigna
along the slope and root architecture affect the stability of the slope unguiculata (400 mmol m−2 s−1) could be detrimental for slope stabi-
during rainfall events (Fig. 4b). Orientation of roots downslope is lity application as it results in desiccation cracks due to excessive
beneficial in draining out water, whereas upslope alignment cause transpiration (Gadi et al., 2017). It is to be noted that root tensile
formation of localized zones of high pore-water pressure. Depending on strength of this species was very less than fibrous or woody plants used
the position and orientation regarding the tree trunk, fork formation in for bio-engineered slopes (Ng et al., 2019c).
root system either exacerbate or dissipate zones of high pore-water
pressure. Conditions in which root curvature converges, high pore- 3. Slope stabilization function of plants
water pressure zones are formed within the slope. Herbaceous or grass
roots on the other hand uniformly distribute water flux into the soil, Fig. 6a presents a schematic representation of the slope stabilization
which minimizes concentrated pore-water pressure zones and even function of vegetation, which can be discussed under three headings.
bridges detrimental desiccation cracks (Bordoloi et al., 2015). They are, (i) Mechanical or root reinforcement, (ii) Hydrological re-
Soil stress history and soil density also effect the tensile effect of inforcement, and (iii) Interception function. Both the root and shoot
roots by altering the proportion of cellulose and bio-mass growth architecture provide these functions. Mechanical reinforcement in-
(Wilson et al., 1977; Correa et al., 2019). Over-consolidated soil char- cludes tensile force mobilization, anchorage upon hard stratum, soil
acterized with low level of O2 (condition known as hypoxia/anoxia), aggregation by root growth and secretion of cohesive enzymes called
plant available water and nutrients can cause reduction in root growth. mucilage. The hydrological reinforcement happens two-fold by re-
Bingham et al., (2010) investigated the effect of soil density and ni- moving water from soil by transpiration and by storage in woody and
trogen availability on the proportion of root cellulose-hemicellulose, deciduous trees. The interception function primarily relates to soil
canopy and root biomass growth of Hordeum vulgare. Soil density from erosion and debris damping as well as interception. All three functions
0.9 g/cc to 1.1 g/cc resulted in a root and shoot biomass decrease by are systematically discussed in detail below.
42% and 47%, respectively. In fact, the increase in density resulted in a
decrease in the cellulose-hemicellulose concentration by 30%, thus 3.1. Mechanical reinforcement
decreasing root tensile strength. However, increase in soil density at
lower available N content, increases the lignin concentration con- The shear strength of root permeated soil was generally explored
siderably, reducing chances of root decay by microbes. Chen et al. (refer Table 1) by estimating the so-called additional root cohesion (ΔS )
(2018) found that arbuscular mycorrhizal fungi application to rooted and incorporating them in slope stability works. The complex nature of
soil (in vetiver grass) increased the tensile strength of the root, as their root–soil interaction makes modelling and quantification of root re-
fungi-root symbiosis resulted in higher cellulose content. inforcement in soil, a challenging endeavor. Wu et al. (1979) and
Plants induce additional negative pore water pressure (or suction) Waldron and Dakessian (1981) are credited for initiating one of the
within the root depth by transpiration (or root water uptake). This foremost shear strength models for root permeated soil. The perpen-
provides additional shear strength based on governing shear strength dicular shear strength model considers the fundamental Mohr-coulomb
equation later explained in Section 3.2. Several plant traits that can shear strength model (Eq. (2)) and extends it for root permeated soil
potentially help an engineer to induce greater hydrological reinforce- (Eq. (3)):
ment to sloped soil are showcased in Fig. 5a. The principal-component
plot based on 14 plant traits carried out on 10 different species, seg- S = C + σN . tan (ϕ) (2)
regate the traits associated with hydrological reinforcement, plant
S = C + (ΔS ) + σN . tan (ϕ) (3)
growth and transpiration (Boldrin et al., 2017a). Principal component
analysis is a technique used to emphasize variation and bring out strong where S is shear strength of soil, C is cohesion due to soil matrix, σN is
patterns in a dataset. The technique basically reduces the dimension- the acting stress normal to shear plane and ϕ is the soil friction angle.
ality of a dataset, while preserving as much ‘variability’ (statistical in- The additional root cohesion component (ΔS ) is the summation of
formation) as possible (Jolliffe and Cadima, 2016). In Fig. 5a, a prin- mobilized tensile stress due to root fibers per unit area of soil. ΔS is
cipal-component (PC) biplot (Boldrin et al., 2017a) is shown from the dependent on the calculation of root tensile strength (Tr) and the soil
projection of plant traits and soil hydromechanical characteristics on cross-section fraction occupied by the root architecture (i.e. RAR).
the plane composed by the two first explanatory axes (Component 1: Mathematically it is represented in Eq. (4).
48% of variation; Component 2: 24% of variation). The small angles
ΔS = K. Tr . RAR (4)
between soil hydromechanical characteristics (i.e. matric suction and
penetration resistance) and plant traits (specific leaf area; root length where K is an empirical factor (1–1.3) accounting for the decomposition
density; root: shoot ratio) have strong correlations among these para- of Tr along tangential and normal component on the shear plane.
meters. Thus, plant properties such as root length density (RLD), spe- However, this model tends to over-estimate root reinforcement based
cific leaf area (SLA) and root to shoot ratio (RSR) should be selected to on the inherent assumption that full tensile strength of each individual
provide reinforcements against rain-induced landslides. Thus, plant root was mobilized during shearing. Subsequently, Pollen et al. (2004),
properties such as root length density (RLD), specific leaf area (SLA) came up with the fiber-bundle model of estimating shear strength,
and root to shoot ratio (RSR) should be selected to provide which accounts for progressive breakage of roots and redistribution of

9
S. Bordoloi and C.W.W. Ng Engineering Geology 275 (2020) 105742

Fig. 5. (a) Biplot projection of plant traits and soil hydro-mechanical parameters; (b) SLA, RLD, RSR relation with matric suction of soil; (c) Effect of LAI, RAI on
suction increment for S. heptaphylla (Ng et al., 2019a); (c) SC relation with transpiration efficiency (a, b and d data considering 10 different species from (Boldrin
et al., 2017a).

loads along the root system. Comparative analysis, using a multitude of Limit equilibrium theory was generally used for calculating shallow
species showed that the developed fiber bundle model (RipRoot algo- slope stability using factor of safety (Fs) approach (Eq. (5)).
rithm) provided better root reinforcement estimations (Bischetti et al.,
2009; Hales et al., 2009; Mickovski et al., 2009; Ji et al., 2012). Acting force
FS =
Shallow landslides in vegetated slopes usually occur in colluvium Resisting force (5)
soil layer (Milledge et al., 2014). In slope failure modelling, colluvium
soil mass can be considered as a rigid volume of thin soil (Fig. 6) sliding If the slope is vegetated with trees (Chiaradia et al., 2016), Fs can be
on a planar shear surface (Casadei et al., 2003; Dietrich et al., 2007). written as shown in Eq. (6).

10
S. Bordoloi and C.W.W. Ng Engineering Geology 275 (2020) 105742

Fig. 6. Schematic representation of (a) Slope stabilization function of vegetation; (b) Idealized representation of hillslope stability parameters considered for pre-
dicting factor of safety (Fs) based on limit equilibrium approach (modified after (Chiaradia et al., 2016)).

FS cohesion and shear strength) to estimate Fs, was discussed under a


CS + Cr _b +
AL
. Cs +
AL
. CR _L + [(D . γS − Dw . γw + q0). cos ∝)] tan∅/ probabilistic framework (Das et al., 2017a, b, 2018).
AB AB
= In-situ measurement of rooted shear strength for shallow slope
(D . γS + Dw . γsat + q0 ) sin∝ stability was done by various instruments. The most common approach
(6) was by taking undisturbed rooted soil sample of known dimension from
field and conducting laboratory direct shear tests on the sample (Fattet
The notations for Eq. (6) are given in Fig. 6. Based on literature et al., 2011; Das et al., 2017a,b; Das et al., 2018). This approach re-
(Schmidt et al., 2001; Rickli and Graf, 2009), D can be assumed as equal quires careful sampling wherein the moisture status and pore size dis-
to 1 and the ratio AL was approximated with a value of 0.5 (Milledge tribution of the sample need to be preserved to get representative shear
AB
et al., 2014). strength parameters. Das et al. (2017b) based on copula distribution
In literature, the ΔS and Fs approach for estimating slope stability analysis observed that vegetation induces more uncertainties in shear
was incorporated in conjunction or separately and tabulated in Table 1. parameters than hydraulic parameters. Hence, in-situ shear strength
The species selected in these studies incorporating mechanical effects of measurement is advisable to reduce the uncertainties that involve
plant reinforcement, mostly were of mature medium trees and grass. sample handling. In-situ-direct shear box test approach, wherein the
The change in mechanical reinforcements with plant age has been shear box is embedded in the vegetated soil (Fan and Su, 2008; Fan and
hardly incorporated in majority of these studies. Plant age changes the Chen, 2010; Yildiz et al., 2018; Yildiz et al., 2020) was used for me-
bio-polymer composition ratio in the roots affecting root tensile chanical shear strength estimation. Yildiz et al. (2018) developed an
strength as previously discussed in Section 2.2. The aforementioned inclinable large-scale direct shear apparatus that can be used in inclined
approaches of root tensile strength and Fs do not incorporate inherent bio-engineered slopes for soils exhibiting dilatancy under saturated
variations in root tensile strength that arise due to root decay (Ni et al., condition. In bio-engineered slopes, wherein mature tree species are
2019a) and root moisture status (Methacanon et al., 2010). Further- planted, the shear strength provided by the mature root system was
more, majority of the studies listed in Table 1 measure the additional estimated by conducting pull-out resistance test (Nilaweera and
root reinforcement based on point measurement or in-situ shear test Nutalaya, 1999; Lin et al., 2010). The penetrometer method which
and do not incorporate the spatial variation of root strength from tree measures the penetration resistance by a loading gauge was also used to
trunk. Plant spacing effects on root tensile strength (Genet et al., 2008; measure soil aggregation and soil shear strength in vegetated plots
Ng et al., 2019c) due to interplant competition and root diameter (Boldrin et al., 2016; Boldrin et al., 2017a,b; Boldrin et al., 2018). The
variability was majorly neglected. Detrimental local soil slippage oc- “corkscrew method” of measuring shear strength of rooted soil (Meijer
currence was observed by Genet et al. (2008) if spacing was too high. et al., 2018; Meijer et al., 2019) in field conditions was found to be ideal
Loades et al. (2010) based on laboratory tests on Hordeum vulgare, due to its ease of use and short test duration. However, it is only feasible
found that the spacing effects on soil shear strength ceases with growth for small trees and grass species with shallow root depth.
period (from 5 to 20 weeks in this case). The incorporation of slope Implementing a full-scale field monitoring is usually expensive, time
angle was majorly accounted for numerical studies wherein the addi- consuming and often difficult to trigger a real time failure to identify
tional root reinforcement by ΔS approach was considered. In the recent the failure mechanism for safety reasons. In the past decade, effect of
past, incorporation of vegetated shear strength parameters (apparent

11
S. Bordoloi and C.W.W. Ng Engineering Geology 275 (2020) 105742

Fig. 7. Geotechnical centrifuge modelling using (a) Live roots (Liang et al., 2017) (b) Root analogues (Liang and Knappett, 2017b).

vegetation on slope stability in terms of mechanical strength has been u w is pore-water pressure; ∅' is effective friction angle; θs is saturated
explored by conducting geotechnical centrifuge tests. The scaled water content and θr is residual water content. (ua − u w ) is re-
models using root reinforcements are spun to attain higher gravity presentation of the negative pore water pressure or matric suction in
forces on the model such that stresses in the model are equal to those the soil. Fig. 8a shows the soil water retention curve (SWRC) of bare soil
expected in field conditions during failure. Roots have previously been and vegetated soil with Schefflera heptaphylla grown in field conditions
modelled by either using live plants (Sonnenberg et al., 2010; with same initial compaction state. Similar SWRC response was also
Askarinejad and Springman, 2014; Ng et al., 2016a; Liang et al., 2017) seen for bare soil and soil with grass species (Axonopus compressus)
or root analogues (Sonnenberg et al., 2012; Eab et al., 2014; Liang grown in laboratory conditions. Both studies indicate that at the same
et al., 2015). Live plants (Fig. 7a) can model root mechanical effects but water content or saturation, root permeated soil retains more matric
are non-repeatable and time consuming to grow plants in the centrifuge suction in comparison to bare soil, thus increase τf as per Eq. (7).
box. Root analogues (Fig. 7b) can be used to model mechanical Hydraulic conductivity (k) in unsaturated soil depend on the ψ as
(Sonnenberg et al., 2010) and hydro-mechanical effects (Ng et al., well as void ratio (e) of the soil. Fig. 8b is a representative relationship
2016a; Leung et al., 2017). They have advantage of high repeatability, of k with ψ and is popularly known as “hydraulic conductivity function
accurate root architecture and can be quickly produced. The seismic (HCF)” in the field of soil mechanics. Mathematically, Eq. (8) represents
performance of rooted slopes using dynamic finite-element analysis, this function, wherein it is seen that after a particular suction, hydraulic
validated against centrifuge test data has gained traction in the recent conductivity is inversely related with ψ.
past (Liang et al., 2015; Liang and Knappett, 2017a,b; Liang et al.,
b θ (e y ) − e(ψ) ' y
2019). ⎡ ∫ln(ψ) ey
θ (e )dy ⎤
k(ψ) = k s ⎢ b θ (e y ) − θs ' y

⎢∫ θ (e )dy ⎥
⎣ ln(ψAEV ) ey ⎦ (8)
3.2. Hydrological reinforcement
Here, b = ln(106), ks is saturated hydraulic conductivity, y is a
Plant transpiration (hydraulic pathway described in Section 2.1) dummy variable for integration of ψ and θ' is the first derivative of
initiates additional lowering down of the water content with corre- function θ. AEV also called air–entry value is the suction at which air
sponding increase in matric suction within the root zone. These “hy- protrudes in the soil matrix upon drying. Based on Eqs. (7) and (8), it is
drologic effects” due to root water uptake induce change in soil shear evident that plant induced suction increases soil shear strength and
strength and hydraulic conductivity, which govern shallow slope sta- reduces water permeability. Unsaturated vegetated slope under rainfall
bility (Ng and Menzies, 2007). The shear strength of an unsaturated soil can thus preserve additional amount of soil suction enhancing shear
(τf ) in terms of volumetric water content (θ) and matric suction (ψ) is strength and reducing water infiltration (delayed pore water pressure
expressed as in Eq. (7): development in rhizosphere). The hydrological effects of vegetated
θ − θr ⎞ ⎤ slope incorporating SWRC and hydraulic conductivity was recently
τf = c ' + (σn − ua)tan∅' + (ua − u w ) ⎡ (tan∅') ⎛
⎜ ⎟
studied for field and lab conditions as shown in Table 2.
⎢ ⎝ θs − θr ⎠ ⎥ (7)
⎣ ⎦
Ng et al. (2016b) developed a novel void ratio dependent theoretical
'
where c is effective cohesion; σn is normal stress; ua is pore-air pressure; model to estimate the SWRC and HCF of root permeated soil. It

12
S. Bordoloi and C.W.W. Ng Engineering Geology 275 (2020) 105742

Fig. 8. (a) Soil-water retention curve (Ng et al., 2016c; Bordoloi et al., 2018), (b) typical hydraulic conductivity function, and (c) phase diagram of root permeated
soil (after Ng et al., 2016c).

considers void ratio change in soil through volume reduction of air ranged between 0 to less than (e0/(1 + e0)) . These limiting values are
voids due to root permeation (Fig. 8c). This modified void ratio (e) of corresponding to bare soil and total soil pore size, respectively.
rooted soil (Eq. (9)) was then fed into a void ratio dependent SWRC Transpiration induced suction in engineered vegetated slopes was
model (Gallipoli et al., 2003) and HCF model (van Genuchten, 1980) to influenced by the soil compaction and plant spacing (Ng et al., 2014; Ng
predict any resultant change in property. The developed model was et al., 2016b). Root growth will be affected by high degree of com-
verified (Ng et al., 2016b,c) to replicate both field and laboratory paction that is prerequisite of urban engineered slopes. Fig. 9a shows
measurements of ψ and unsaturated hydraulic conductivity. that both root and shoot length (grass species) was lowest at 95% de-
gree of compaction (DOC), whereas at lower densities, growth was
e0 − R v (1 + e0) relatively higher. However, based on suction response (Fig. 9a) in ve-
e=
1 + R v (1 + e0) (9) getated soil after 1-hour duration of high intensity rainfall event, it was
observed that lower densities (70% and 80%) lose highest suction. Such
Here, e0 is the void ratio of bare soil (dimensionless); Rv is the root suction loss at low compaction throughout the root depth, coupled with
volume ratio (mm3/ mm3) and is species-dependent with limiting value

13
Table 2
Chronological overview of hydraulic reinforcement in vegetated slope.
Reference Species Soil type Analysis type Hydraulic Consideration for Suction range Key conclusions and contribution#
conductivity studied (kPa)
Plant age Root Spacing
effect depth effects
profile
S. Bordoloi and C.W.W. Ng

Simon and Collison (2002) Multiple species (6) Sand Lab and numerical x x ✓ x 0–88 Root reinforcement and hydro mechanical
modelling reinforcement by tree cover increased FS by 32% and
71%
Doussan et al. (2006) Lupin Sandy clay Lab x x ✓ ✓ 0–95 Taproot architecture induced higher spatially
concentrated uptake zone
Indraratna et al. (2006) Lime tree, Eucalyptus Clay Numerical x x ✓ ✓ 1–1500 Maximum settlement occurs near point with highest
root water uptake rate
Pollen-Bankhead and Simon Multiple species (6) Silty loam Lab and numerical x x ✓ x 0–4 For riverbanks, evapotranspiration provided potential
(2010) modelling benefit to FS only during summer
Leung and Ng (2013) Dicranopteris pedata, Colluvium, coarse Field and Seepage ✓ x ✓ x 0.1–90 During dry season, peak suction induced by plant Etr
Rhodomyrtus tomentosa, ash tuff analysis was up to 200 kPa, Depth of influence < 2 times of
Baeckea frutescens root depth
Ng et al. (2014) Cynadon Dactylon (CD) Silty sand Lab ✓ x ✓ x 0–80 At 0.95 relative compaction, suction retained was
double than bare soil
Ng et al. (2013) CD Silty sand Lab x x ✓ x 0–80 Vertical suction influence zone was 4-times the root
depth
Garg and Ng (2015) Schefflera heptaphylla (SH) CDG Numerical x x ✓ x 0–830 Computed suction decreased by 21%, when coupled
effects of soil density and RAI were considered
Garg et al. (2015a) SH CDG Lab x x ✓ x 0–80 Evaporation rate reduced when LAI was higher
Garg et al. (2015b) SH CDG Lab x x ✓ x 0–80 SH with higher LAI has lower water stress tolerance
Garg et al. (2015c) CD, SH CDG Field ✓ x ✓ x 0.8–900 SH induce max. suction due to high shoot mass and

14
RAI
Leung et al. (2015a) SH Clayey sand Lab and numerical x x ✓ x 0.1–90 Contribution of root-water uptake was less compared
with root-induced change to suction
Leung et al. (2015b) CD, SH Silty sand Field ✓ x ✓ x 1–90 No significant difference (< 10%) of infiltration, ks
and suction retained for both species
Ng et al. (2015) NA CDG Analytical analysis x x ✓ x 0–90 Maximum suction was induced by parabolic root
architecture
Boldrin et al. (2016) Corylus avellana, Ilex Sandy loam Lab test x x x x 0–1500 Mean suction induced by Corylus A was 2.7 times than
Aquifolium Ilex A
Ng et al. (2016a) NA CDG Centrifuge modelling ✓ x ✓ x 0–90 FS of heart shaped roots was highest among tested root
system
Ng et al. (2016b) SH CDG Lab ✓ x ✓ ✓ 0–90 364% increase in peak suction for spacing of 60 mm as
compared to 180 mm
Ni et al. (2017) CD, SH CDG Field ✓ x ✓ ✓ 0–130 Highest matric suction preserved for the 240 mm
spacing
Leung et al. (2017) NA CDG Centrifuge modelling ✓ x ✓ x 0.1–200 Heart-shaped roots induce higher suction, leading to
14% reduction of infiltration
Li et al. (2016) Tall Fescue Low plastic clay Lab ✓ x ✓ x 10–77 Roots restrict crack formation in early stages of
drying–wetting cycles.
Liu et al. (2016) NA CDG Analytical analysis x x ✓ x 5–49 FS Exponential > FS parabolic in terms of root architecture
Shao et al. (2016) SH CDG Numerical modelling ✓ x ✓ ✓ 0–100 New dual-permeability model developed for
preferential flow in vegetated slope
Boldrin et al. (2017a) Multiple species (10) Sandy loam Lab test x x x x 0–70 LAI, RLD and the RSR have positive relation with
suction
Das et al. (2017a) CD, SH CDG Field and Modelling x x x x 0–105 Species type did not affect dependence structure of ψ
and θ.
Das et al. (2017b) CD, SH CDG Field and Probabilistic x x x x 0–600 Hydrological correlation (ψ-θ) is stronger in treed soil
Modelling than grass.
(continued on next page)
Engineering Geology 275 (2020) 105742
Table 2 (continued)

Reference Species Soil type Analysis type Hydraulic Consideration for Suction range Key conclusions and contribution#
conductivity studied (kPa)
Plant age Root Spacing
effect depth effects
profile
S. Bordoloi and C.W.W. Ng

Gadi et al. (2017) Pongamia pinnata, Cyperus, Sand Field ✓ ✓ x ✓ x Spatial heterogeneity in both vegetation cover and k
Poaceae was more significant during drought periods
Jotisankasa and Chrysopogon zizanioides Clayey sand, low Lab ✓ x x x 0–100 Influence of roots on permeability was less significant.
Sirirattanachat (2017) plastic silt
Leung et al. (2018) Salix viminalis tora, Silty sand Lab and numerical ✓ ✓ ✓ x 0–20 Age effect was more prominent for willow than grass
Festulolium grass species
Song et al. (2017) CD, Vetiver grass Low plastic clay In-situ and lab ✓ x ✓ x 9–32 Hydraulic conductivity (CD < Bare < Vetiver)
Boldrin et al. (2018) Corylus avellana, Ulex Sandy loam Lab test x x x x 0–70 Ulex europaeus induced higher suction up to 7 m depth
europaeus
Das et al. (2018) CD, SH CDG Field and Probabilistic x x x x 0–900 Probabilistic model developed for univariate ψ, under
Modelling varying non-linear c − ϕ correlation
Feng et al. (2019) CD, SH Sand Analytical ✓ x ✓ x 0–90 CN, RN and WR govern unsaturated seepage
Gadi et al. (2018) SH CDG Numerical ✓ x ✓ x 0–1500 EDZ depth decreased with increase in LAI
Guo et al. (2018) CD CDG Field ✓ x ✓ x 0–45 CD was effective in minimizing percolation
Liu et al. (2018) SH CDG, Silt Analytical analysis ✓ x ✓ x 0–40 Root water induced suction increased with a
desaturation coefficient
Ni et al. (2018a) NA CDG Numerical modelling ✓ x ✓ x 0.1–100 Tree root-water uptake induced suction in a mixed
species did not increase further when LAI exceeds 5
Ni et al. (2018b) SH CDG Numerical modelling ✓ x ✓ x 0.1–100 Hydrological reinforcement benefits was seen for
twice the depth of only mechanical reinforcement.
Singh et al. (2018) CD, SH CDG Numerical and ✓ x x x 0–600 Permeability was high in treed than grassed cover,
probabilistic since tree induced higher suction

15
modelling
Świtała et al. (2017) Avena sativa Silty sand Constitutive x x x x 0–90 New coupled hydro-mechanical model for vegetated
modelling and lab soil was validated
Zhu et al. (2018) NA CDG Numerical modelling x x ✓ x 20–100 Exponential and triangular root architecture enhance
soil suctions more than the uniform and parabolic
Gadi et al. (2019) Axonopus Compressus, CD Silty loam Lab x x x x 0–2500 SC indicates plant wilting point
Garg et al. (2019) SH CDG Probabilistic ✓ ✓ x x 1–1500 Probability of wilting for CD was lower than SH
modelling
Ng et al. (2019d) SH CDG Lab x x ✓ x 0–90 Plant water uptake intensity positively correlated with
root length density
Ng et al. (2019c) CD, SH CDG Field and numerical ✓ ✓ ✓ ✓ 0–120 Plant spacing effects was predominant only up to
9 months
Ni et al. (2019a) CD, SH, Schefflera arboricola CDG Theoretical, Lab and ✓ x x x 0.7–100 Void ratio-based model could capture effects of root
field growth and root decay on soil hydraulic properties

#
Capillary effect number (CN), root water uptake number (RN), water transfer-storage ratio (WR), Stomatal conductance (SC), Depth of evaporation dominant zone (EDZ), Evapotranspiration (Etr).
Engineering Geology 275 (2020) 105742
S. Bordoloi and C.W.W. Ng Engineering Geology 275 (2020) 105742

Fig. 9. Effect of (a) degree of compaction (Ng et al., 2013; Bordoloi et al., 2018) and (b) plant spacing (Ng et al., 2016b) on vegetation growth and induced root zone
suction.

resultant increase on rainfall infiltration (refer Eq. (8)) can potentially vegetated soil has been majorly neglected in literature. Plant metabo-
result in shallow slope failure. High soil compaction state limits easy lism (especially for trees) changes with growth stage (initial growth,
growth of grass but showcases lower drop of suction as compared to vegetative and mature) (Unterscheutz et al., 1974; Aslam et al., 1977)
bare soil in shallow depths, with no change in suction at deeper profiles. and thus, there is high probability of induced suction being dependent
Ng et al. (2016b) investigated the effect of plant spacing on suction on plant age. Studies on vegetated soil with monitored suction have
induced in highly compacted soil (90% DOC) in controlled laboratory recently considered plant age effects on SWRC (Leung et al., 2018). The
conditions (Fig. 9b). The non-linear and parabolic RAI for the species monitoring has been done at maximum for 8 weeks in case of labora-
(Schefflera heptaphylla) indicated that root growth was conducive for tory tests (Leung et al., 2018; Ni et al., 2020) and 13 months in case of
highest spacing (180 mm) with lowest root decay. On the contrary, the field tests (Ng et al., 2019b).
roots of plants with lowest spacing (60 mm) were shorter, more loca- The usage of centrifuge modelling to replicate the hydrological ef-
lized and exhibited root decaying. The effect of plant spacing was fects of vegetation on slope stability was explored in the past decade.
evident for suction profiles, wherein lowest spacing resulted in highest Artificial roots usage in geotechnical centrifuge test specimen was first
induced suction at all depths. Tree saplings placed at lower spacing developed (Ng et al., 2014; Ng and Yu, 2014) for modelling both plant
experience higher suction as there is high competition for water from hydrological and mechanical effects. The artificial roots consisting of a
the neighboring roots. Similar trends were observed for the same tree porous filter (in contact with soil) with high air entry value were con-
species in field tests wherein spacing of 120, 180 and 240 mm were nected to a vacuum system to simulate artificial negative pore water
considered (Ni et al., 2019b). In this study, in-situ hydraulic con- pressure. Transpiration was simulated at filter tip as water within soil
ductivity rates gradually decreased by one order as plant spacing in- seeps into the artificial root upon lowering of total water head by at-
creased. This maybe plausible due to soil pore blockage by roots and tached vacuum system. The axial rigidity of the artificial roots was
mucilage secretion (Bengough, 2012). However, for non-compacted maintained by using suitable material for filter. This was done such that
soil, presence of tree roots apparently increases rainwater infiltration mechanical reinforcement provided by actual roots was additionally
due to preferential flow along the roots as reported from dye-tracer mimicked during centrifuge tests. Kamchoom et al. (2014) and Ng et al.
experiments (Mitchell et al., 1995; Johnson and Lehmann, 2006; Luo (2016a,b,c) investigated three artificial root models based on root
et al., 2019). geometries (tap-shaped, heart shaped and plate shaped root) in cen-
Table 2 summarizes work done on the hydrological reinforcing ef- trifuge. Tap and heart shaped roots were identified to be better at re-
fects of vegetation incorporating both laboratory, field and numerical sisting pull-out stress as compared to plate shaped root. Ng et al.
investigation. The brief and specific outcome of each individual study (2016a) found that slope supported by heart-shaped roots retained
was incorporated in the remarks section of the table. Most of the ta- highest suction after rainfall events and provided higher Fs as compared
bulated studies incorporate depth profile to investigate the hydraulic to other tested root geometries.
response of rooted soil. However, detailed response of plant spacing Numerical, analytical, and constitutive modelling of compacted
effects on hydraulic properties was only considered in the recent past vegetated soil for engineered slope stability applications were explored
(Ng et al., 2016b; Ni et al., 2017;Shao et al., 2017; Gadi et al., 2017; Ng in the past decades. Riverbank slope stability for vegetated soil was
et al., 2019a,c). The effect of plant age on hydraulic properties of investigated by considering both mechanical and hydraulic

16
S. Bordoloi and C.W.W. Ng Engineering Geology 275 (2020) 105742

reinforcement together using the limit equilibrium and Mohr-column (RMD in kg/m3) which is defined as the ratio of dry mass of roots to
failure approach (Simon and Collison, 2002). Indraratna et al. (2006) total dry mass of soil considered. Fig. 11a presents the possible ranges
proposed a mathematical model to represent root water uptake by as- of normalized values of τc coeff and K e coeff with RMD based on field
suming that tree root zone geometry was conical in nature. Based on the measurements in vegetated soil. τc coeff was seen to increase linearly with
assumption, numerical analysis by finite element method was used to higher RMD regardless of the species. However, the rate of increase
predict root water uptake, pore water pressure distribution and soil depends on the species considered for individual studies. Along the
deformation along tree vicinity. Numerical models considering only same lines, K e decreases logarithmically with higher RMD, thus redu-
hydrological reinforcements were developed in recent past (Garg and cing surficial erosion. Shen et al. (2017) conducted parametric study to
Ng, 2015; Leung et al., 2015a; Singh et al., 2018, Zhu et al., 2018). evaluate the effect of variable vegetation cover on various sites within
Analytical solutions of water flow, SWRC and pore water pressure Xiaojiagou Ravine (Sichuan province, China), where extreme erosion
profiles considering hydrological reinforcement provided by con- and debris flow had taken place after the 2008 Wenchuan earthquake.
trasting root architecture were also validated in recent times (Ng et al., Based on forensic tests and numerical modelling, including in-situ
2015; Liu et al., 2016; Liu et al., 2018; Feng et al., 2020). The modelling sampling and satellite imagery to gauge vegetation growth for a two-
approaches used in compacted vegetated slope stability were provided year period, the relative area susceptible to erosion was modelled.
in Table 2. The “modified Richard’s equation” which obeys mass con- Based on the simulation results, the relationship of relative erosion area
servation principle and Darcy's law was majorly used in these studies to (i.e. the ratio of the erosion area at τc coeff to the erosion area at
model root-water uptake by incorporating a sink term (S(z)). The (τc coeff = 1) ) and the variable τc coeff of vegetated soil was presented
modified partial differential equation with time-dependency was pre- (Fig. 11b). After τc coeff reaches beyond 2, surface erosion was almost
sented as Eq. (10). diminished by the erosion resistance provided by the fibrous roots.
Deciduous trees were reported to have physical effects on the dy-
∂θ ⎛ ∂h ⎞ ∂θ ⎛ ∂h ⎞ ∂θ
k + k = − S (z ) namics of low volume debris flow triggered after heavy rainfall. The
∂x ⎝ ∂x ⎠ ∂z ⎝ ∂z ⎠ ∂t (10) physical effects namely are (i) kinetic energy absorption through direct
where k represents co-efficient of hydraulic conductivity at a given impact of boulder with tree trunk (Stokes et al., 2005; Lundström et al.,
matric suction, h represents hydraulic head, (x, z) are Cartesian co-or- 2009); and (ii) energy dissipation of debris flow by coppice structures
dinates in horizontal and vertical directions, t represents time elapsed (Ciabocco et al., 2009). Winching tests were conducted upon trees on
and θ is the volumetric water content. Recently, constitutive models actual slopes to gauge the energy that trees can absorb before failure
were developed that incorporate both hydrological and mechanical (Dorren and Berger, 2006; Peterson and Claassen, 2012). Based on
effects of plants (Świtała et al., 2018; Świtała and Wu, 2018). However, winching tests, it was found that the energy that a tree can absorb
the developed models could not account for the complex root archi- before toppling (Ecap) or uproot failure increased exponentially with
tectures and temporal changes induced by root growth. trunk or stem diameter, regardless of the species (Fig. 11b). Jonsson
(2007) developed a numerical model which can be used to measure Ecap
3.3. Interception of a single tree during debris flow. The numerical model was based on a
series of full-scale impact tests on Picea abies using high-speed cameras
The interception function of vegetation for slope stability can be and accelerometers. This developed numerical single tree model was
majorly discussed in terms of erosion resistance and debris or rock fall validated for higher energy levels based on full-scale rock fall experi-
interception. Vegetation intercepts erosion through three ways: (i) ments conducted on Abies alba plantation (Stokes et al., 2005). Based on
provides protection to the soil particles against raindrop impact the simulated full-scale test data, it was revealed that tree stem ab-
(Herwitz, 1987); (ii) attenuates runoff volume through increased sur- sorbed 2/3rd of the energy, while rest was dissipated into the root
face infiltration through roots, i.e. preferential flow (Blanco and Lal, system. However, it should be noted that the relationship is variable
2008); and (iii) reduces sediment transportation by entrapment (Burylo with impact velocity and shape of the projectile. Mature trees have been
et al., 2012). Vannoppen et al. (2015) conducted a meta-analysis review used as anchor elements to flexible barriers for protection against debris
of the mechanical effects of roots on concentrated flow erosion. Based flow (Bourrier et al., 2015). Radtke et al. (2014) compared 40 coppice
on their study, a schematic illustration comparing the erosion-reducing plantation on two sloped forest sites to study its protective effect
potential of roots and canopy cover is shown in Fig. 10a. It can be in- against rockfall using the simulation model Rocky for 3D. Based on
ferred that plant roots have higher potential against rill and gully simulations, it was observed that random stem distribution in coppice
erosion, whereas canopy growth was more effective in mitigating splash plantations provided better protective effect than the clumped stem
and inter-rill erosion (Gyssels et al., 2005; Zuazo and Pleguezuelo, distribution. Stoffel and Wilford (2012) show that tree-rock fall inter-
2009). Soil loss due to erosion (A) in hillslope, was generally estimated action leads to change in annual ring pattern and width of willow
using the predictive Revised Universal Soil Loss Equation (RUSLE) as species, which can be used to trace debris flow intervals, especially for
per Eq. (11) (IECA, 2008). forensic studies. For instance, Stoffel et al. (2005, 2008)) used the intra-
seasonal position of debris-flow damage in willows, local meteor-
A= R. K. LS. C. P (11)
ological data to reconstruct 400 years of debris flow history at Riti-
where R represents rainfall erosivity factor, K represents soil erodibility graben terrain, Swiss Alps.
factor; LS is a topographic factor derived from slope length and gra-
dient; C is the vegetative cover and management factor; and P is erosion 4. Present gap areas for future research
control practice factor. The general methodology adopted to under-
stand the erosion hazard of a vegetated hill slope is shown in Fig. 10b. The future gap areas emanating from the discussion in the previous
The erosion rate in soil is generally expressed as given in Eq. (12). sections are detailed below in four categories. These include suscept-
ibility to failure considering extreme climate scenarios; plantation
i = K e coeff (τ − τc coeff ) (12)
strategy; vegetation maintenance and real time monitoring; and nature-
where i = erosion rate (m/s), K e is the erodibility co-efficient (m3/
coeff based alteration to root properties.
N-s), τ = shear stress at air–water interface (kPa), and τc coeff = critical
erosive shear stress at initiation of soil erosion (kPa). The smaller the 4.1. Extreme climate scenarios
magnitude of K e coeff and larger the τc , higher will be the erosion re-
sistance. Zhu and Zhang (2016) investigated the effects of grass roots on Slope stability on vegetated slopes have been extensively studied for
surficial erosion, using the vegetation parameter- root mass density high precipitation (including 100-year return period) events in the past

17
S. Bordoloi and C.W.W. Ng Engineering Geology 275 (2020) 105742

Fig. 10. (a) Idealized role of canopy and root area ratio on resistance against soil erosion considering different erosion patterns (Vannoppen et al., 2015); (b)
Schematic representation to assess erosion potential of vegetated hillslope through RUSLE approach.

Fig. 11. (a) Possible ranges of τccoeff and K ecoeff with root mass density, (b) Simulated relationship between relative erosion area and τccoeff for a site near the epicenter of
2008 Wenchuan earthquake based on variable vegetation cover (after Shen et al., 2017), (c) relation between energy adsorbed before failure and stem diameter of
tree (after Stokes et al., 2005;).

18
S. Bordoloi and C.W.W. Ng Engineering Geology 275 (2020) 105742

decade considering climate change effects (Ng et al., 2019a). Climate include plants with fibrous roots and trees with greater root depth for
change has also increased the frequency of droughts leading to tree anchorage. Canopy area should also be high for minimizing erosion. In
mortality, forest fires, freeze–thaw cycles. Atmospheric CO2 emission is hill slopes which are vulnerable to debris flow, random stem distribu-
also increasing exponentially and projected to reach 700–1000 ppm by tion in coppice plantations will provide better protective functions than
2100 due to climate change (Keeling and Whorf, 2005; IPCC, 2013). the clumped stem distribution (refer Section 3.3). For landfill slopes,
Extended droughts coupled with a drop in ground water table will make species selection should be based on shallow root morphology (against
vegetation in slopes more susceptible to wilting and mortality. Plant barrier intrusion), high transpiration and drought resistance (no natural
wilting will lead to termination of the hydrological reinforcement ac- water resource). Plant spacing also affects transpiration induced suc-
tion and gradually decrease mechanical reinforcement through de- tion, root decaying, desiccation cracks and spatial variation in hy-
caying roots. A decayed root network following wilting will lead to draulic conductivity due to canopy distribution (Indraratna et al., 2006;
multiple regions of concentrated pore water pressure within the slope, Genet et al., 2008; Ni et al., 2019b; Gadi et al., 2019). These con-
which increases the probability of slope failure upon subsequent rain- siderations should be studied and accounted for, while designing bio-
fall. The past year witnessed an outbreak of extreme forest fires like the engineered slopes based on application. Wind induced tree uprooting
2019 California fire (Williams et al., 2019), the 2019 Amazon fires can be modelled based on computational fluid dynamics (Gan and
(Fonseca et al., 2019), and the 2019–2020 Australian fire (Nolan et al., Salim, 2014). Tree uprooting, especially root uprooting can lead to
2020). Post-fire debris flow has been extensively reported in literature progressive slope failure during a cyclone as seen for 2018-Typhoon
(Jakob et al., 2005). The root zone is essentially turned into a hydro- Mangkhut (Sassa, 2019). Usage of geotechnical centrifuge with a si-
phobic layer with decayed or degraded root network which essentially mulated wind setup could potentially reveal failure mechanisms with
form a “cocktail” of causal agents triggering shallow landslides and different slope geometry, tree location and height. The use of capillary
debris flow (Doerr et al., 2009). Few case studies exist that measure barrier in embankment slopes and mining slopes with herbaceous
change in material properties post-fire and its implication on debris species can be explored as they can retain water in the root zone and
flow (Kean et al., 2011; De Graff, 2014). Building upon those pre- minimize water percolation to deeper layers. The use of capillary bar-
liminary literature, the mechanisms that induce post fire landslide need rier has been reported (Ityel et al., 2014) to increase dissolved oxygen
to be investigated in future. Centrifuge modelling of prototype slopes concentration in the root zone and improved biomass by 5% and 18%,
with a hydrophobic topsoil layer up to the root depth along with dif- respectively.
ferent slope geometry, rainfall intensities, root architecture and soil
type could unravel failure and trigger mechanisms related to post fire 4.3. Vegetation maintenance and real time monitoring
landslides. A transition from perennially frozen to seasonally frozen soil
and vice versa due to climate change will accelerate the effect of free- From an engineering perspective, a lot of existing knowledge por-
ze–thaw processes on vegetated slopes (Patton et al., 2019). Till date, tray the beneficial effects of vegetation on slope stability and protection
the effect of freeze–thaw cycles on slope stability due to the changes in measures. Most of urban engineered slopes such as road embankment
hydraulic (Konrad and Samson, 2000) and physical properties (Qi et al., and landfill, do not have a natural source of water and are dependent on
2006; Yamamoto and Springman, 2019) of soil have been discussed rainfall events or regular irrigation. Climate change has resulted in
with only bare soil mechanics. These cycles increase the degree of periods of extended droughts wherein plants are susceptible to wilting
fracturing, pore water pressure and hydraulic conductivity; while re- and may result in tree mortality. The recent case studies of extended
ducing the physical properties such as apparent cohesion and friction droughts and tree mortality in previously lush regions were reflected in
angle (Guo et al., 2014). Change in rainfall patterns and rapid snow/ice the recent Zimbabwe drought (Frischen et al., 2020), El Nino drought in
melting are likely to increase the frequency and magnitude of landslides South Africa (Baudoin et al., 2017) and the 2013–2019 Australian
based on field observations (Patton et al., 2019). Particularly hillslopes drought (Gallant et al., 2019). Over irrigation to maintain bio-en-
which experience post-disturbance change in vegetation cover due to gineered slopes is also not cost-effective wherein water is not a readily
forest fire are susceptible to catastrophic failure. Recently, Archer and available resource. In the past, irrigation scheduling in urban spaces
Ng (2017) developed a geotechnical centrifuge setup that could simu- have been basically developed based on soil moisture status, neglecting
late freeze–thaw cycles. The usage of such setup in simulating slope plant health status. One of the key parameters that governs irrigation
failure incorporating varied root analogues, artificial cracks, hydro- schemes and vegetation survival is the permanent wilting point (PWP).
phobic soil cover and precipitation patterns could reveal new findings In the past decades, soil science practitioners, agriculturists and ecol-
and failure mechanisms associated with post-fire and frozen soil con- ogists have determined PWP from the SWRC of vegetated soil (Garg
ditions. Rise in CO2 concentrations is generally estimated to increase et al., 2017) as the water content corresponding to a soil matric po-
plant biomass (Keenan et al., 2013; Reef et al., 2016), which should tential of 4.2 pF (i.e., 1500 kPa). The approach of determining PWP at
facilitate higher mechanical and hydrological reinforcement in vege- 1500 kPa as an invariant index representing the lower limit of plant
tated slopes (Ng et al., 2018; Ng et al., 2019e). However, increase in water availability may be misleading (Philip, 1957; Raats et al., 2002).
plant water uptake to satiate the higher biomass will lead to lowering Kirkham (2014) has criticized that this overestimates the water avail-
down of ground water table, increase susceptibility to wilting and ne- ability for plants in case of fine-textured (clay) soils where sufficient
cessitate higher irrigation supply. Thus, the change in plant water up- bound water may be present, which cannot be extracted by plant roots.
take efficiency, expected change in biomass, transpiration induced On the contrary, for coarse-textured soils, the soil reaches residual
suction and irrigation requirement for different species need to be as- water content (2–3%) even at (200–500) kPa (Dekker et al., 2001).
certained at higher CO2 concentrations. Hence, the consideration of PWP at 1500 kPa in coarse grained soils is
dubious. Research is needed to quantify the actual PWP of a plant
4.2. Plantation strategy species based on soil type and CO2 concentrations, which will help
practitioners to optimize irrigation needs. Recently, Boldrin et al.
Species type and plant spacing selected for constructing bio-en- (2019) were able to apply thermal imaging technique to evaluate
gineered slopes such as road embankment slope, hill slope, landfill plants’ ability to induce suction through transpiration (hydrological
slope and mining slope need to be different based on the infrastructure reinforcement). Gadi et al. (2019) were able to segregate plant canopy
(Fig. 1). In case of embankment slopes wherein erosion protection and based on color to co-relate plant stomatal conductance which is an
strength are of paramount interest, species selection should be based on essential indicator of plant wilting. Similar smart monitoring techni-
fibrous root system, high LAI and greater cellulose-hemicellulose con- ques can be used wherein plants act as “live sensors”. Based on imaging
tent. In case of steep hill slope, vegetation planning ideally should techniques of plant traits such as canopy temperature and leaf color,

19
S. Bordoloi and C.W.W. Ng Engineering Geology 275 (2020) 105742

real-time models can be developed to indirectly measure plant health studies on young trees. The suction range considered for majority of the
parameters and soil suction for predicting tree mortality. The usage of studies on hydrological reinforcements range from 1 to 90 kPa, which is
unmanned aerial vehicles can help monitor remote vegetated slopes representative of scenarios relating to rainfall induced landslides and
susceptible to slope failure. The usage of machine learning technique to optimal plant growth (refer Fig. 3b). Vegetation possess a “love-hate
identify and predict landslide occurrence in slopes has gained traction relationship” towards slope stability as it gradually reaches higher
recently (Wang et al., 2020). Adopting the spatial and temporal effects suction range near residual water content or wilting point. Of course, it
of vegetation cover can make these models more robust, as they directly imparts higher matric suction resulting in low water infiltration and
affect the suction profile of slopes. inducing additional shear strength (refer Section 3.2). However, it also
makes the slope vulnerable to future catastrophic failure due to ex-
4.4. Nature based alteration to improve root growth and root cohesion acerbated desiccation cracks and root decay upon wilting. To mitigate
such adverse events, plantation strategy and real time monitoring of
The usage of synthetic materials such as fibers, geotextiles and geo- vegetated slopes is important, especially in the early-plant establish-
grids in vegetated soil have been in practice (Cazzuffi et al., 2014). ment period. The end of early-plant establishment period also includes
However, they are non-biodegradable, do not facilitate nutrient reten- smooth functioning of the microbial diversity, ecological restoration
tion, are aesthetically unpleasant and possess threat to local biota. and maintenance of reinforcement functions. Thus, natural amend-
Nature based amendment material and organisms which do not hamper ments that promote microbial diversity and do not harm native flora-
the biota, rather, help in vegetation growth especially during early fauna need to be taken under consideration.
plant establishment period can be incorporated in bio-engineered The review points out gap areas which require urgent study due to
slopes. The use of natural fibers has been shown to decrease cracks as extreme climate scenarios. Bio-engineered slopes are now exposed to
well as retain higher moisture for high biomass growth (Kurugodu frequent and extreme scenarios of prolonged droughts, forest fires and
et al., 2018). Biochar, another bio-amendment, can be explored in slope high atmospheric CO2 concentrations. Such extreme cases can be stu-
stability as it restricts volume expansion, desiccation cracks, increases died in laboratory by building prototype slopes in geotechnical cen-
plant available water content, facilitates nutrition and has higher ve- trifuge. Accounting for extreme droughts and maintenance of vegetated
getation yield (Atkinson et al., 2010). Natural biopolymers such as slopes, key hydrological terms related to plants i.e. threshold suction
Xantham gum have been explored to increase shear strength of coarse- (where plants feel drought stress), tipping point suction (beyond which
grained soil (Chen and Harbottle, 2019; Chen et al., 2019a,b) and can plants cannot survive) and permanent wilting point (where plant dies)
be extended for usage in rooted soil. Recently, the prospect of using need to be revisited. All these terms in the field of geotechnical en-
microbial induced calcite precipitation (MICP) was explored for re- gineering are arbitrarily established based on soil status, rather than
sisting cracks in expansive soil (Liu et al., 2019). The use of mycorrhiza accounting plant hydraulic pathway (refer Fig. 3b). These terms should
fungi and symbiosis of certain root species results in increase in root be plant specific and the knowledge can help in optimal irrigation and
tensile strength, thus enhancing the mechanical reinforcement (Zhang maintenance of bio-engineered slopes, especially where water resources
et al., 2016; Chen et al., 2018). The use of mycorrhiza (ectomycorrhiza) are scarce and limited. Smart and real time monitoring of vegetated
in the rhizosphere usually enhances root branching density, root yield infrastructure could be explored using thermal imaging, machine
and increase RLD (Cui and Caldwell, 1996; Hodge et al., 2000; Stokes learning tools and unmanned aerial vehicles. Use of nature-based al-
et al., 2009). Plants can access a larger surface area of nutrient and teration as well as suitable plantation schemes, keeping in mind the
phosphate rich soil via the mycelial network of the mycorrhiza infrastructural needs would make it easier to build futuristic and sus-
(Lambers et al., 2008). The use of genetically modified plants for in- tainable bio-engineered slopes.
creasing resistance against drought and salinity have been extensively
used in the field of agriculture (Wang et al., 2003). Genetically mod- Declaration of Competing Interest
ified herbaceous and non-crop species could be used to stabilize mining
slopes such as fly ash deposits, tailing dams and salt mines; wherein The authors declare that they have no known competing financial
natural water source is absent and presents conditions of high alkalinity interests or personal relationships that could have appeared to influ-
and salinity. ence the work reported in this paper.

5. Concluding remarks Acknowledgements

The current review encompasses theoretical understanding, la- The authors thank the Environment Conservation Fund
boratory experiments, prototype modelling, field observations and (ECWW19EG01) provided by the Government of Hong Kong SAR,
long-term monitoring studies relating to soil–plant-water atmosphere China and the Research Grant No. AoE/E-603/18 awarded by the Areas
interaction. In literature, the mechanical reinforcement provided by of Excellence (AoE) Scheme of the Hong Kong Research Grants Council
roots have been extensively studied and understood from the perspec- of the Government of Hong Kong SAR, China.
tive of geotechnical engineering (Table 1). Even though root branching
(forking), root bio-polymer concentration, vegetation spacing and rhi- Appendix A. Supplementary data
zosheath biota greatly affect the adhesive forces between soil and root,
majority of the studies consider singularly the root tensile strength as a Supplementary data to this article can be found online at https://
direct measure of apparent root cohesion in sloped soil. The hydro- doi.org/10.1016/j.enggeo.2020.105742.
logical reinforcement effects provided by plants have gained grounds in
the past decade based on recent advancements in physical and geo- References
technical centrifuge modelling (Table 2). The past decade has seen the
usage of conventional lab experiments to quantify transpiration in- Abdi, E., Azhdari, F., Abdulkhani, A., Mariv, H.S., 2014. Tensile strength and cellulose
duced suction and its relationship with vegetation traits. It is to be content of Persian ironwood (Parrotia persica) roots as bioengineering material. J.
mentioned that most of these studies include young trees and limited For. Sci. 60 (10), 425–430.
Abe, K., Ziemer, R.R., 1991. Effect of tree roots on shallow-seated landslides. In: USDA
grass species. Future studies need to account the temporal effects of forest Service Gen Tech. Rep. PSW-GT, pp. 11–20.
vegetation growth and consequent hydrological reinforcement pro- Ali, N., Farshchi, I., Múazu, M.A., Rees, S.W., 2012. Soil-root interaction and effects on
vided in slopes. Studies on spacing effect of vegetation and age effect on slope stability analysis. Electron. J. Geotech. Eng. 17, 319–328.
Allen, R.G., Smith, M., Perrier, A., Pereira, L.S., 1994. An update for the definition of
hydrological benefits are still at a nascent stage, with only a couple of

20
S. Bordoloi and C.W.W. Ng Engineering Geology 275 (2020) 105742

reference evapotranspiration. ICID Bull. 43 (2), 1–34. Chen, C., Wu, L., Harbottle, M., 2019b. Influence of biopolymer gel-coated fibres on sand
Archer, A., Ng, C.W.W., 2017. Hong Kong University of Science and Technology (HKUST). reinforcement as a model of plant root behaviour. Plant Soil 438 (1–2), 361–375.
Centrifuge environmental chamber. U.S. Patent Application 15/425,501. Cheng, Q., Tang, C.S., Zeng, H., Zhu, C., An, N., Shi, B., 2020. Effects of microstructure on
Aroca, R., Porcel, R., Ruiz-Lozano, J.M., 2012. Regulation of root water uptake under desiccation cracking of a compacted soil. Eng. Geol. 265, 105418.
abiotic stress conditions. J. Exp. Bot. 63 (1), 43–57. Chiaradia, E.A., Vergani, C., Bischetti, G.B., 2016. Evaluation of the effects of three
Askarinejad, A., Springman, S.M., 2015. Centrifuge modelling of the effects of vegetation European forest types on slope stability by field and probabilistic analyses and their
on the response of a silty sand slope subjected to rainfall. In: Computer Methods and implications for forest management. For. Ecol. Manage. 370, 114–129.
Recent Advances in Geomechanics: Proceedings of the 14th International Conference Chok, Y., Kaggwa, G., Jaksa, M., Griffiths, D., 2004. Modelling the effects of vegetation on
of International Association for Computer Methods and Recent Advances in stability of slopes.
Geomechanics, 2014 (IACMAG 2014). Taylor & Francis Books Ltd, pp. 1339–1344. Ciabocco, G., Boccia, L., Ripa, M.N., 2009. Energy dissipation of rockfalls by coppice
Aslam, M., Lowe, S.B., Hunt, L.A., 1977. Effect of leaf age on photosynthesis and tran- structures. Nat. Hazard. Earth. Sys. 9 (3), 993.
spiration of cassava (Manihot esculenta). Can. J. Bot. 55 (17), 2288–2295. Ciurleo, M., Cascini, L., Calvello, M., 2017. A comparison of statistical and deterministic
Atkinson, C.J., Fitzgerald, J.D., Hipps, N.A., 2010. Potential mechanisms for achieving methods for shallow landslide susceptibility zoning in clayey soils. Eng. Geol. 223,
agricultural benefits from biochar application to temperate soils: a review. Plant Soil 71–81.
337 (1–2), 1–18. Cochard, H., Bréda, N., Granier, A., 1996. Whole tree hydraulic conductance and water
Aziz, M.M., Palta, J.A., Siddique, K.H., Sadras, V.O., 2017. Five decades of selection for loss regulation in Quercus during drought: evidence for stomatal control of embo-
yield reduced root length density and increased nitrogen uptake per unit root length lism? In: Annales des Sciences Forestieres. EDP Sciences., pp. 197–206.
in Australian wheat varieties. Plant Soil 413 (1–2), 181–192. Correa, J., Postma, J.A., Watt, M., Wojciechowski, T., 2019. Soil compaction and the
Basirat, M., Mousavi, S.M., Abbaszadeh, S., Ebrahimi, M., Zarebanadkouki, M., 2019. The architectural plasticity of root systems. J. Exp. Bot. 70 (21), 6019–6034.
rhizosheath: a potential root trait helping plants to tolerate drought stress. Plant Soil Crozier, M.J., 2010. Deciphering the effect of climate change on landslide activity: a
445 (1–2), 565–575. review. Geomorph. 124 (3–4), 260–267.
Baudoin, M.A., Vogel, C., Nortje, K., Naik, M., 2017. Living with drought in South Africa: Cui, M.U.Y.I., Caldwell, M.M., 1996. Facilitation of plant phosphate acquisition by ar-
lessons learnt from the recent El Niño drought period. Int. J. Disast. Risk. Re. 23, buscular mycotrhizas from enriched soil patches: I. Roots and hyphae exploiting the
128–137. same soil volume. New phytol. 133 (3), 453–460.
Bengough, A.G., 2012. Water dynamics of the root zone: rhizosphere biophysics and its Danjon, F., Barker, D.H., Drexhage, M., Stokes, A., 2008. Using three-dimensional plant
control on soil hydrology. Vadose Zone J. 11 (2). root architecture in models of shallow-slope stability. Ann. Bot. 101 (8), 1281–1293.
Bingham, I.J., Bengough, A.G., Rees, R.M., 2010. Soil compaction–N interactions in Das, G.K., Hazra, B., Garg, A., Ng, C.W., 2017a. Impact of hydrological and mechanical
barley: root growth and tissue composition. Soil Tillage Res. 106 (2), 241–246. correlations on the reliability of vegetated slopes. ASCE ASME J. Risk Uncertain. Eng.
Bischetti, G.B., Chiaradia, E.A., Epis, T., Morlotti, E., 2009. Root cohesion of forest species Syst. A Civ. Eng., Part A Civ. Eng. 3 (4), 04017029.
in the Italian Alps. Plant Soil 324 (1–2), 71–89. Das, G.K., Hazra, B., Garg, A., Ng, C.W.W., Avani, N., Lateh, H., 2017b. Bivariate prob-
Blanco, H., Lal, R., 2008. Principles of Soil Conservation and Management (167169). abilistic modelling of hydro-mechanical properties of vegetated soils. Adv. Civ. Eng.
Springer, New York. Mater. 6 (1), 235–257.
Blight, G.E., 1997. 37th Rankine Lecture: interactions between the atmosphere and the Das, G.K., Hazra, B., Garg, A., Ng, C.W.W., 2018. Stochastic hydro-mechanical stability of
Earth. Geotechnique 47 (4), 715–766. vegetated slopes: an integrated copula based framework. Catena 160, 124–133.
Boldrin, D., Leung, A.K., Bengough, A.G., Delage, P., Cui, Y.-J., Ghabezloo, S., Pereira, J.- De Baets, S., Poesen, J., 2010. Empirical models for predicting the erosion-reducing ef-
M., Tang, A.-M., 2016. Desirable leaf traits for hydrological reinforcement of soil. In: fects of plant roots during concentrated flow erosion. Geomorph. 118 (3–4), 425–432.
E3S Web Conferences. EDP Sciences, pp. 12006. De Baets, S., Poesen, J., Reubens, B., Wemans, K., De Baerdemaeker, J., Muys, B., 2008.
Boldrin, D., Leung, A.K., Bengough, A.G., 2017a. Correlating hydrologic reinforcement of Root tensile strength and root distribution of typical Mediterranean plant species and
vegetated soil with plant traits during establishment of woody perennials. Plant Soil their contribution to soil shear strength. Plant Soil 305 (1–2), 207–226.
416 (1–2), 437–451. De Graff, J.V., 2014. Improvement in quantifying debris flow risk for post-wildfire
Boldrin, D., Leung, A.K., Bengough, A.G., 2017b. Root biomechanical properties during emergency response. Geoenviron. Disast. 1 (1), 5.
establishment of woody perennials. Ecol. Eng. 109, 196–206. Dekker, L.W., Doerr, S.H., Oostindie, K., Ziogas, A.K., Ritsema, C.J., 2001. Water re-
Boldrin, D., Leung, A.K., Bengough, A.G., 2018. Hydrologic reinforcement induced by pellency and critical soil water content in a dune sand. Soil Sci. Soc. Am. J. 65 (6),
contrasting woody species during summer and winter. Plant Soil 427 (1–2), 369–390. 1667–1674.
Boldrin, D., Leung, A.K., Bengough, A.G., Jones, H.G., 2019. Potential of thermal imaging Delmer, D.P., Amor, Y., 1995. Cellulose biosynthesis. Plant Cell 7 (7), 987.
in soil bioengineering to assess plant ability for soil water removal and air cooling. Demenois, J., Rey, F., Stokes, A., Carriconde, F., 2017. Does arbuscular and ectomycor-
Ecol. Eng. 141, 105599. rhizal fungal inoculation improve soil aggregate stability? A case study on three
Bordoloi, S., Yamsani, S.K., Garg, A., Sreedeep, S., Borah, S., 2015. Study on the efficacy tropical species growing in ultramafic Ferralsols. Pedobiologia 64, 8–14.
of harmful weed species Eicchornia crassipes for soil reinforcement. Ecol. Eng. 85, Demenois, J., Carriconde, F., Bonaventure, P., Maeght, J.L., Stokes, A., Rey, F., 2018.
218–222. Impact of plant root functional traits and associated mycorrhizas on the aggregate
Bordoloi, S., Hussain, R., Garg, A., Sreedeep, S., Zhou, W.H., 2017. Infiltration char- stability of a tropical Ferralsol. Geoderma 312, 6–16.
acteristics of natural fiber reinforced soil. Transp. Geotech. 12, 37–44. Dietrich, W.E., McKean, J., Bellugi, D., Perron, T., 2007. The prediction of shallow
Bordoloi, S., Gadi, V.K., Hussain, R., Sahoo, L., Garg, A., Sreedeep, S., Mei, G., Poulsen, landslide location and size using a multidimensional landslide analysis in a digital
T.G., 2018. Influence of Eichhornia crassipes fibre on water retention and cracking of terrain model. In: Chen, C.L., Major, J.J. (Eds.), Proceedings of the Fourth
vegetated soils. Géotech. Lett. 8 (2), 130–137. International Conference on Debris-Flow Hazards Mitigation: Mechanics, Prediction,
Börjesson, G., Sundh, I., Svensson, B., 2004. Microbial oxidation of CH4 at different and Assessment (DFHM-4); Chengdu, China, September 10-13, 2007. IOS Press, The
temperatures in landfill cover soils. FEMS Microbiol. Ecol. 48 (3), 305–312. Netherlands, Amsterdam, pp. 12.
Bourrier, F., Lambert, S., Baroth, J., 2015. A reliability-based approach for the design of Doerr, S.H., Shakesby, R.A., MacDonald, L.H., 2009. Soil water repellency: a key factor in
rockfall protection fences. Rock Mech. Rock Eng. 48 (1), 247–259. post-fire erosion. In: Fire Effects on Soils and Restoration Strategies. CRC Press, pp.
Brodribb, T.J., 2009. Xylem hydraulic physiology: the functional backbone of terrestrial 213–240.
plant productivity. Plant Sci. 177 (4), 245–251. Dorren, L.K., Berger, F., 2006. Stem breakage of trees and energy dissipation during
Burylo, M., Hudek, C., Rey, F., 2011. Soil reinforcement by the roots of six dominant rockfall impacts. Tree Physiol. 26 (1), 63–71.
species on eroded mountainous marly slopes (Southern Alps, France). Catena 84 Doussan, C., Pierret, A., Garrigues, E., Pagès, L., 2006. Water uptake by plant roots:
(1–2), 70–78. II–modelling of water transfer in the soil root-system with explicit account of flow
Burylo, M., Rey, F., Bochet, E., Dutoit, T., 2012. Plant functional traits and species ability within the root system–comparison with experiments. Plant Soil 283 (1–2), 99–117.
for sediment retention during concentrated flow erosion. Plant Soil 353 (1–2), Drew, M.C., 1997. Oxygen deficiency and root metabolism: injury and acclimation under
135–144. hypoxia and anoxia. Annu. Rev. Plant Biol. 48 (1), 223–250.
Casadei, M., Dietrich, W.E., Miller, N.L., 2003. Controls on shallow landslide size. In: Eab, K.H., Takahashi, A., Likitlersuang, S., 2014. Centrifuge modelling of root-reinforced
Debris-Flow Hazards Mitigation: Mechanics, Prediction, and Assessment, pp. 91–101. soil slope subjected to rainfall infiltration. Géotech. Lett. 4 (3), 211–216.
Cazzuffi, D., Cardile, G., Gioffrè, D., 2014. Geosynthetic engineering and vegetation Fan, C.C., Chen, Y.W., 2010. The effect of root architecture on the shearing resistance of
growth in soil reinforcement applications. Transp. Infra. Geotechnol. 1 (3–4), root-permeated soils. Ecol. Eng. 36 (6), 813–826.
262–300. Fan, C.C., Su, C.F., 2008. Role of roots in the shear strength of root-reinforced soils with
Chahine, M.T., 1992. The hydrological cycle and its influence on climate. Nature 359 high moisture content. Ecol. Eng. 33 (2), 157–166.
(6394), 373. Fattet, M., Fu, Y., Ghestem, M., Ma, W., Foulonneau, M., Nespoulous, J., Le Bissonnais, Y.,
Chatterjea, K., 2011. Severe wet spells and vulnerability of urban slopes: case of Stokes, A., 2011. Effects of vegetation type on soil resistance to erosion: relationship
Singapore. Nat. Hazards 56 (1), 1–18. between aggregate stability and shear strength. Catena 87 (1), 60–69.
Chen, C., Wu, L., Harbottle, M., 2019. Exploring the effect of biopolymers in near-surface Feddes, R.A., 1982. Simulation of Field Water Use and Crop Yield (194–209). Pudoc.
soils using xanthan gum–modified sand under shear. Can. Geotech. J. 999, 1–10. Feng, S., Liu, H.W., Ng, C.W.W., 2019. Dimensional analysis of pore-water pressure re-
Chen, X.W., Wong, J.T.F., Ng, C.W.W., Wong, M.H., 2016. Feasibility of biochar appli- sponse in a vegetated infinite slope. Can. Geotech. J. 56 (8), 1119–1133.
cation on a landfill final cover—a review on balancing ecology and shallow slope Feng, S., Liu, H.W., Ng, C.W.W., 2020. Analytical analysis of the mechanical and hy-
stability. Environ. Sci. Pollut. R. 23 (8), 7111–7125. drological effects of vegetation on shallow slope stability. Comput. Geotech. 118,
Chen, X.W., Kang, Y., San So, P., Ng, C.W.W., Wong, M.H., 2018. Arbuscular mycorrhizal 103335.
fungi increase the proportion of cellulose and hemicellulose in the root stele of ve- Fonseca, M.G., Alves, L.M., Aguiar, A.P.D., Arai, E., Anderson, L.O., Rosan, T.M.,
tiver grass. Plant Soil 425 (1–2), 309–319. Shimabukuro, Y.E., de Aragão, L.E.O.E.C., 2019. Effects of climate and land-use
Chen, C., Wu, L., Harbottle, M., 2019a. Exploring the effect of biopolymers in near-surface change scenarios on fire probability during the 21st century in the Brazilian Amazon.
soils using xanthan gum-modified sand under shear. Can. Geotech. J (In Press). Glob. Change Biol. 25 (9), 2931–2946.

21
S. Bordoloi and C.W.W. Ng Engineering Geology 275 (2020) 105742

Frischen, J., Meza, I., Rupp, D., Wietler, K., Hagenlocher, M., 2020. Drought risk to examples. Earth-Sci. Rev. 102981.
agricultural systems in zimbabwe: a spatial analysis of hazard, exposure, and vul- Indraratna, B., Fatahi, B., Khabbaz, H., 2006. Numerical analysis of matric suction effects
nerability. Sustainability 12 (3), 752. of tree roots. Proc. Inst. Civ. Eng.-Geotech. Eng. 159 (2), 77–90.
Gadi, V.K., Bordoloi, S., Garg, A., Sahoo, L., Berretta, C., Sekharan, S., 2017. Effect of International Erosion Control Association (IECA), 2008. Best Practice Erosion and
shoot parameters on cracking in vegetated soil. Environ. Geotech. 5 (2), 123–130. Sediment Control.
Gadi, V.K., Hussain, R., Bordoloi, S., Hossain, S., Singh, S.R., Garg, A., Sekharan, S., IPCC, 2013. Climate Change, 2013. The Physical Science Basis. In: Contribution of
Karangat, R., Lingaraj, S., 2019. Relating stomatal conductance and surface area with Working Group I to the Fifth Assessment Report of the Intergovernmental Panel on
evapotranspiration induced suction in a heterogeneous grass cover. J. Hydrol. 568, Climate Change, pp. 1535.
867–876. IPCC, I.P.O.C.C., 2019. Special report on global warming of 1.5 C (SR15).
Gadi, V., Singh, S., Singhariya, M., Garg, A., Sreedeep, S., Ravi, K., 2018. Modeling soil- Ityel, E., Ben-Gal, A., Silberbush, M., Lazarovitch, N., 2014. Increased root zone oxygen
plant-water interaction: effects of canopy and root parameters on soil suction and by a capillary barrier is beneficial to bell pepper irrigated with brackish water in an
stability of green infrastructure. Eng. Comput. 35 (3), 1543–1566. arid region. Agri. Water Manage. 131, 108–114.
Gallant, A., Lewis, S., Adamu, M., McGregor, S., 2019. January. The connection between Jackson, R.B., Sperry, J.S., Dawson, T.E., 2000. Root water uptake and transport: using
protracted drought and extreme precipitation during recent decadal-scale drought physiological processes in global predictions. Trends Plant Sci. 5 (11), 482–488.
episodes. Geophys. Res. Abstr. Vol, 21). Jakob, M., Hungr, O., Jakob, D.M., 2005. Debris-flow Hazards and Related Phenomena.
Gallipoli, D., Wheeler, S.J., Karstunen, M., 2003. Modelling the variation of degree of Springer, Berlin, pp. 739.
saturation in a deformable unsaturated soil. Géotechnique 53 (1), 105–112. Ji, J., Kokutse, N., Genet, M., Fourcaud, T., Zhang, Z., 2012. Effect of spatial variation of
Gan, C.J., Salim, S.M., 2014. Numerical analysis of Fluid-Structure Interaction between tree root characteristics on slope stability. A case study on Black Locust (Robinia
wind flow and trees. In: Proceedings of the World Congress on Engineering, pp. 1218. pseudoacacia) and Arborvitae (Platycladus orientalis) stands on the Loess Plateau,
Garg, A., 2015. Effects of vegetation types and characteristics on induced soil suction. China. Catena 92, 139–154.
Doctoral dissertation. Hong Kong University of Science and Technology. Johnson, K.L., 2018. Baby, It’s Cold Inside: Maintaining Membrane Integrity during
Garg, A., Coo, J.L., Ng, C.W.W., 2015a. Field study on influence of root characteristics on Freezing. Plant Physiol. 177 (4), 1350–1351.
soil suction distribution in slopes vegetated with Cynodon dactylon and Schefflera Johnson, M.S., Lehmann, J., 2006. Double-funneling of trees: Stemflow and root-induced
heptaphylla. Earth Surf. Process. Landf. 40 (12), 1631–1643. preferential flow. Ecosci. 13 (3), 324–333.
Garg, A., Ng, C.W.W., 2015. Investigation of soil density effect on suction induced due to Jolliffe, I.T., Cadima, J., 2016. Principal component analysis: a review and recent de-
root water uptake by Schefflera heptaphylla. J. Plant Nutr. Soil Sci. 178 (4), 586–591. velopments. Philos. Trans. R. Soc. A 374 (2065), 20150202.
Garg, A., Leung, A.K., Ng, C.W.W., 2015b. Comparisons of soil suction induced by eva- Jonsson, M.J., 2007. Energy absorption of trees in a rockfall protection forest. Doctoral
potranspiration and transpiration of S. heptaphylla. Can. Geotech. J. 52 (12), dissertation. ETH Zurich.
2149–2155. Jotisankasa, A., Sirirattanachat, T., 2017. Effects of grass roots on soil-water retention
Garg, A., Leung, A.K., Ng, C.W.W., 2015c. Transpiration reduction and root distribution curve and permeability function. Can. Geotech. J. 54 (11), 1612–1622.
functions for a non-crop species Schefflera heptaphylla. Catena 135, 78–82. Kamchoom, V., Leung, A.K., Ng, C.W.W., 2014. Effects of root geometry and transpiration
Garg, A., Li, J., Hou, J., Berretta, C., Garg, A., 2017. A new computational approach for on pull-out resistance. Géotech. Lett. 4 (4), 330–336.
estimation of wilting point for green infrastructure. Measurement 111, 351–358. Kean, J.W., Staley, D.M., Cannon, S.H., 2011. In situ measurements of post-fire debris
Garg, A., Hazra, B., Zhu, H., Wen, Y., 2019. A simplified probabilistic analysis of water flows in southern California: comparisons of the timing and magnitude of 24 debris-
content and wilting in soil vegetated with non-crop species. Catena 175, 123–131. flow events with rainfall and soil moisture conditions. J. Geophysic. Res. Earth Surf.
Gariano, S.L., Guzzetti, F., 2016. Landslides in a changing climate. Earth-Sci. Rev. 162, 116 (F4).
227–252. Keeling, C.D., Whorf, T.P., 2005. Atmospheric carbon dioxide record from Mauna Loa.
Genet, M., Stokes, A., Salin, F., Mickovski, S.B., Fourcaud, T., Dumail, J.F., Van Beek, R., Carbon Dioxide Research Group, Scripps Institution of Oceanography, University of
2005. The influence of cellulose content on tensile strength in tree roots. Plant Soil California La Jolla, California 92093-0444.
278 (1–2), 1–9. Keenan, T.F., Hollinger, D.Y., Bohrer, G., Dragoni, D., Munger, J.W., Schmid, H.P.,
Genet, M., Kokutse, N., Stokes, A., Fourcaud, T., Cai, X., Ji, J., Mickovski, S., 2008. Root Richardson, A.D., 2013. Increase in forest water-use efficiency as atmospheric carbon
reinforcement in plantations of Cryptomeria japonica D. Don: effect of tree age and dioxide concentrations rise. Nature 499 (7458), 324–327.
stand structure on slope stability. For. Ecol. Manag. 256 (8), 1517–1526. Khalil, H.A., Hossain, M.S., Rosamah, E., Azli, N.A., Saddon, N., Davoudpoura, Y., Islam,
Ghestem, M., Sidle, R.C., Stokes, A., 2011. The influence of plant root systems on sub- M.N., Dungani, R., 2015. The role of soil properties and its interaction towards
surface flow: implications for slope stability. Biosci. 61 (11), 869–879. quality plant fiber: a review. Renewable Sustainable Energy Rev. 43, 1006–1015.
Ghestem, M., Veylon, G., Bernard, A., Vanel, Q., Stokes, A., 2014. Influence of plant root Kieselbach, T., Schröder, W.P., 2003. The proteome of the chloroplast lumen of higher
system morphology and architectural traits on soil shear resistance. Plant Soil 377 plants. Photosynth. Res. 78 (3), 249–264.
(1–2), 43–61. Kim, J.H., Fourcaud, T., Jourdan, C., Maeght, J.L., Mao, Z., Metayer, J., Meylan, L.,
Gray, D.H., Sotir, R.B., 1996. Biotechnical and Soil Bioengineering Slope Stabilization: A Pierret, A., Rapidel, B., Roupsard, O., De Rouw, A., 2017. Vegetation as a driver of
Practical Guide for Erosion Control. John Wiley & Sons. temporal variations in slope stability: the impact of hydrological processes. Geophys.
Gregory, P.J., Bengough, A.G., Grinev, D., Schmidt, S., Thomas, W.B.T., Wojciechowski, Res. Lett. 44 (10), 4897–4907.
T., Young, I.M., 2009. Root phenomics of crops: opportunities and challenges. Funct. Kirkham, M.B., 2014. Principles of Soil and Plant Water Relations. Academic Press.
Plant Biol. 36 (11), 922–929. Knapen, A., Poesen, J., Govers, G., Gyssels, G., Nachtergaele, J., 2007. Resistance of soils
Guo, H.W., Ng, C.W.W., Coo, J.L., Ni, J.J., 2018. Field study of water infiltration into a to concentrated flow erosion: a review. Earth-Sci. Rev. 80 (1–2), 75–109.
vegetated sustainable three-layer landfill cover system. In: Unsaturated Soil 2018, Koerner, R.M., 2012. Designing with Geosynthetics. Xlibris Corporation.
Hong Kong. Conference Proceedings, pp. 1–6. Konrad, J.M., Samson, M., 2000. Hydraulic conductivity of kaolinite-silt mixtures sub-
Guo, Y., Shan, W., Jiang, H., Sun, Y., Zhang, C., 2014. The impact of freeze–thaw on the jected to closed-system freezing and thaw consolidation. Can. Geotech. J. 37 (4),
stability of soil cut slope in high-latitude frozen regions. In: Landslides in Cold 857–869.
Regions in the Context of Climate Change. Springer, Cham, pp. 85–98. Korup, O., Görüm, T., Hayakawa, Y., 2012. Without power? Landslide inventories in the
Gyssels, G., Poesen, J., Bochet, E., Li, Y., 2005. Impact of plant roots on the resistance of face of climate change. Earth Surf. Process. Landf. 37 (1), 92–99.
soils to erosion by water: a review. Prog. Phys. Geogr. 29 (2), 189–217. Kurugodu, H.V., Bordoloi, S., Hong, Y., Garg, A., Garg, A., Sreedeep, S., Gandomi, A.H.,
Hales, T.C., Ford, C.R., Hwang, T., Vose, J.M., Band, L.E., 2009. Topographic and ecologic 2018. Genetic programming for soil-fiber composite assessment. Adv. Eng. Softw.
controls on root reinforcement. J. Geophys. Res. Earth Surf. 114 (F3). 122, 50–61.
Haling, R.E., Brown, L.K., Bengough, A.G., Young, I.M., Hallett, P.D., White, P.J., George, Lambers, H., Chapin III, F.S., Pons, T.L., 2008. Plant Physiological Ecology. Springer
T.S., 2013. Root hairs improve root penetration, root–soil contact, and phosphorus Science & Business Media.
acquisition in soils of different strength. J. Exp. Bot. 64 (12), 3711–3721. Lefi, E., Medrano, H., Cifre, J., 2004. Water uptake dynamics, photosynthesis and water
Haling, R.E., Brown, L.K., Bengough, A.G., Valentine, T.A., White, P.J., Young, I.M., use efficiency in field-grown Medicago arborea and Medicago citrina under pro-
George, T.S., 2014. Root hair length and rhizosheath mass depend on soil porosity, longed Mediterranean drought conditions. Ann. Appl. Biol. 144 (3), 299–307.
strength and water content in barley genotypes. Planta 239 (3), 643–651. Leung, A.K., Ng, C.W.W., 2013. Analyses of groundwater flow and plant evapo-
Hearle, J.W., Morton, W.E., 2008. Physical Properties of Textile Fibres. Elsevier. transpiration in a vegetated soil slope. Can. Geotech. J. 50 (12), 1204–1218.
Herwitz, S.R., 1987. Raindrop impact and water flow on the vegetative surfaces of trees Leung, A.K., Garg, A., Coo, J.L., Ng, C.W.W., Hau, B.C.H., 2015a. Effects of the roots of
and the effects on stemflow and throughfall generation. Earth Surf. Process. Landf. 12 Cynodon dactylon and Schefflera heptaphylla on water infiltration rate and soil hy-
(4), 425–432. draulic conductivity. Hydrol. Process. 29 (15), 3342–3354.
Hobbs, P.R.N., Humphreys, B., Rees, J.G., Tragheim, D.G., Jones, L.D., Gibson, A., Leung, A.K., Garg, A., Ng, C.W.W., 2015b. Effects of plant roots on soil-water retention
Rowlands, K., Hunter, G., Airey, R., 2002. Monitoring the role of landslides in ‘soft and induced suction in vegetated soil. Eng. Geol. 193, 183–197.
cliff’coastal recession. In: Instability-Planning and Management, pp. 589–600. Leung, A.K., Kamchoom, V., Ng, C.W.W., 2017. Influences of root-induced soil suction
Hodge, A., Robinson, D., Fitter, A.H., 2000. An arbuscular mycorrhizal inoculum en- and root geometry on slope stability: a centrifuge study. Can. Geotech. J. 54 (3),
hances root proliferation in, but not nitrogen capture from, nutrient-rich patches in 291–303.
soil. New Phytol. 145 (3), 575–584. Leung, A.K., Boldrin, D., Liang, T., Wu, Z.Y., Kamchoom, V., Bengough, A.G., 2018. Plant
Hong, Y., Adler, R., Huffman, G., 2007. Use of satellite remote sensing data in the age effects on soil infiltration rate during early plant establishment. Géotechnique 68
mapping of global landslide susceptibility. Nat. Hazards 43 (2), 245–256. (7), 646–652.
Hubble, T., Clarke, S., Stokes, A., Phillips, C., 2017. 4th International Conference on soil Li, J.H., Li, L., Chen, R., Li, D.Q., 2016. Cracking and vertical preferential flow through
bio-and eco-engineering (SBEE-2016) ‘The Use of Vegetation to Improve Slope landfill clay liners. Eng. Geol. 206, 33–41.
Stability’. Ecol. Eng. 109, 141–144. Liang, T., Knappett, J.A., 2017b. Newmark sliding block model for predicting the seismic
Hürlimann, M., Coviello, V., Bel, C., Guo, X., Berti, M., Graf, C., Hübl, J., Miyata, S., performance of vegetated slopes. Soil Dyn. Earthq. Eng. 101, 27–40.
Smith, J.B., Yin, H.Y., 2019. Debris-flow monitoring and warning: review and Liang, T., Knappett, J.A., 2017a. Centrifuge modelling of the influence of slope height on

22
S. Bordoloi and C.W.W. Ng Engineering Geology 275 (2020) 105742

the seismic performance of rooted slopes. Géotechnique 67 (10), 855–869. Ng, C.W.W., Ni, J.J., Leung, A.K., Zhou, C., Wang, Z.J., 2016c. Effects of planting density
Liang, T., Knappett, J.A., Duckett, N., 2015. Modelling the seismic performance of rooted on tree growth and induced soil suction. Géotechnique 66 (9), 711–724.
slopes from individual root–soil interaction to global slope behaviour. Géotechnique Ng, C.W.W., Tasnim, R., Coo, J.L., 2018. Effects of atmospheric CO2 concentration on
65 (12), 995–1009. soil-water retention and induced suction in vegetated soil. Eng. Geol. 242, 108–120.
Liang, T., Bengough, A.G., Knappett, J.A., MuirWood, D., Loades, K.W., Hallett, P.D., Ng, C.W.W., Leung, A., Ni, J., 2019a. Plant-soil Slope Interaction. CRC Press.
Boldrin, D., Leung, A.K., Meijer, G.J., 2017. Scaling of the reinforcement of soil slopes Ng, C.W.W., Tasnim, R., Wong, J.T.F., 2019b. Coupled effects of atmospheric CO2 con-
by living plants in a geotechnical centrifuge. Ecol. Eng. 109, 207–227. centration and nutrients on plant-induced soil suction. Plant Soil 439 (1–2), 393–404.
Liang, T., Knappett, J.A., Leung, A.K., Bengough, A.G., 2019. Modelling the seismic Ng, C.W.W., Ni, J.J., Leung, A.K., 2019b. Effects of plant growth and spacing on soil
performance of root-reinforced slopes using the finite-element method. Géotechnique hydrological changes: a field study. Géotechnique 1–15.
1–17. Ng, C.W.W., Wang, Z.J., Leung, A.K., Ni, J.J., 2019e. A study on effects of leaf and root
Lin, D.G., Huang, B.S., Lin, S.H., 2010. 3-D numerical investigations into the shear characteristics on plant root water uptake. Géotechnique 69 (2), 151–157.
strength of the soil–root system of Makino bamboo and its effect on slope stability. Ni, J.J., Leung, A.K., Ng, C.W.W., So, P.S., 2017. Investigation of plant growth and
Ecol. Eng. 36 (8), 992–1006. transpiration-induced matric suction under mixed grass–tree conditions. Can.
Linton, M.J., Nobel, P.S., 2001. Hydraulic conductivity, xylem cavitation, and water Geotech. J. 54 (4), 561–573.
potential for succulent leaves of Agave deserti and Agave tequilana. Inter. J. Plant Sci. Ni, J.J., Leung, A.K., Ng, C.W.W., 2018a. Modelling soil suction changes due to mixed
162 (4), 747–754. species planting. Ecol. Eng. 117, 1–17.
Liu, H.W., Feng, S., Ng, C.W.W., 2016. Analytical analysis of hydraulic effect of vegeta- Ni, J.J., Leung, A.K., Ng, C.W.W., Shao, W., 2018b. Modelling hydro-mechanical re-
tion on shallow slope stability with different root architectures. Comput. Geotech. 80, inforcements of plants to slope stability. Comput. Geotech. 95, 99–109.
115–120. Ni, J.J., Leung, A.K., Ng, C.W.W., 2019a. Modelling effects of root growth and decay on
Liu, H.W., Feng, S., Garg, A., Ng, C.W.W., 2018. Analytical solutions of pore-water soil water retention and permeability. Can. Geotech. J. 56 (7), 1049–1055.
pressure distributions in a vegetated multi-layered slope considering the effects of Ni, J.J., Leung, A.K., Ng, C.W.W., 2019b. Influences of plant spacing on root tensile
roots on water permeability. Comput. Geotech. 102, 252–261. strength of Schefflera arboricola and soil shear strength. Landscape Ecol. Eng. 15 (2),
Liu, B., Zhu, C., Tang, C.S., Xie, Y.H., Yin, L.Y., Cheng, Q., Shi, B., 2019. Bio-remediation 223–230.
of desiccation cracking in clayey soils through microbially induced calcite pre- Ni, J., Ng, C.W.W., Gao, Y., 2020. Modelling root growth and soil suction due to plant
cipitation (MICP). Eng. Geol. 105389. competition. J. Theor. Biol. 484, 110019.
Loades, K.W., Bengough, A.G., Bransby, M.F., Hallett, P.D., 2010. Planting density in- Nilaweera, N.S., Nutalaya, P., 1999. Role of tree roots in slope stabilization. Bull. Eng.
fluence on fibrous root reinforcement of soils. Ecol. Eng. 36 (3), 276–284. Geol. Environ. 57 (4), 337–342.
López-Bucio, J., Cruz-Ramırez, A., Herrera-Estrella, L., 2003. The role of nutrient avail- Nolan, R.H., Boer, M.M., Collins, L., Resco de Dios, V., Clarke, H., Jenkins, M., Kenny, B.,
ability in regulating root architecture. Curr. Opin. Plant Biol. 6 (3), 280–287. Bradstock, R.A., 2020. Causes and consequences of eastern Australia’s 2019–20
Lundström, T., Jonsson, M.J., Volkwein, A., Stoffel, M., 2009. Reactions and energy ab- season of mega-fires. Global Change Biol.
sorption of trees subject to rockfall: a detailed assessment using a new experimental Novick, K.A., Miniat, C.F., Vose, J.M., 2016. Drought limitations to leaf-level gas ex-
method. Tree Physiol. 29 (3), 345–359. change: results from a model linking stomatal optimization and cohesion–tension
Luo, Z., Niu, J., Zhang, L., Chen, X., Zhang, W., Xie, B., Du, J., Zhu, Z., Wu, S., Li, X., 2019. theory. Plant, Cell Environ. 39 (3), 583–596.
Roots-enhanced preferential flows in deciduous and coniferous forest soils revealed Partov, D., Maślak, M., Ivanov, R., Petkov, M., Sergeev, D., Dimitrova, A., 2016. The
by dual-tracer experiments. J. Environ. Qual. 48 (1), 136–146. development of wooden bridges through the ages–a review of selected examples of
Lynch, J., 1995. Root architecture and plant productivity. Plant Physiol. 109 (1), 7. heritage objects. Part 1–the milestones. Czasopismo Techniczne, 2016(Budownictwo
Mahannopkul, K., Jotisankasa, A., 2019. Influence of root suction on tensile strength of Zeszyt 2-B 2016), 93–105.
Chrysopogon zizanioides roots and its implication on bioslope stabilization. J. Mount. Patton, A.I., Rathburn, S.L., Capps, D.M., 2019. Landslide response to climate change in
Sci. 16 (2), 275–284. permafrost regions. Geomorphology.
Mao, Z., Jourdan, C., Bonis, M.L., Pailler, F., Rey, H., Saint-André, L., Stokes, A., 2013. Pawlik, Ł., 2013. The role of trees in the geomorphic system of forested hillslopes—a
Modelling root demography in heterogeneous mountain forests and applications for review. Earth-Sci. Rev. 126, 250–265.
slope stability analysis. Plant Soil 363 (1–2), 357–382. Perillo, C.A., Gupta, S.C., Nater, E.A., Moncrief, J.F., 1999. Prevalence and initiation of
Mao, Z., Bourrier, F., Stokes, A., Fourcaud, T., 2014. Three-dimensional modelling of preferential flow paths in a sandy loam with argillic horizon. Geoderma 89 (3–4),
slope stability in heterogeneous montane forest ecosystems. Ecol. Model. 273, 11–22. 307–331.
Mao, Z., Wang, Y., McCormack, M.L., Rowe, N., Deng, X., Yang, X., Xia, S., Nespoulous, J., Peterson, C.J., Claassen, V., 2012. An evaluation of the stability of Quercus lobata and
Sidle, R.C., Guo, D., Stokes, A., 2018. Mechanical traits of fine roots as a function of Populus fremontii on river levees assessed using static winching tests. Forestry 86 (2),
topology and anatomy. Ann. Bot. 122 (7), 1103–1116. 201–209.
Matsuura, S., Asano, S., Okamoto, T., 2008. Relationship between rain and/or meltwater, Philip, J.R., 1957. The physical principles of soil water movement during the irrigation
pore-water pressure and displacement of a reactivated landslide. Eng. Geol. 101 cycle.
(1–2), 49–59. Poggi, I., Polidori, J.J., Gandoin, J.M., Paolacci, V., Battini, M., Albertini, M., Ameglio, T.,
Meijer, G.J., Bengough, A.G., Knappett, J.A., Loades, K.W., Nicoll, B.C., 2018. In situ Cochard, H., 2007. Stomatal regulation and xylem cavitation in Clementine (Citrus
measurement of root reinforcement using corkscrew extraction method. Can. clementina Hort) under drought conditions. J. Hortic. Sci. Biotechnol. 82 (6),
Geotech. J. 55 (10), 1372–1390. 845–848.
Meijer, G., Bengough, G., Knappett, J., Loades, K., Nicoll, B., 2019. Measuring the Pollen, N., Simon, A., 2005. Estimating the mechanical effects of riparian vegetation on
strength of root-reinforced soil on steep natural slopes using the corkscrew extraction stream bank stability using a fiber bundle model. Water Resour. Res. 41 (7).
method. Forests 10 (12), 1135. Pollen, N., Simon, A., Collison, A., 2004. Advances in assessing the mechanical and hy-
Methacanon, P., Weerawatsophon, U., Sumransin, N., Prahsarn, C., Bergado, D.T., 2010. drologic effects of riparian vegetation on streambank stability. Riparian Veg. Fluvial
Properties and potential application of the selected natural fibers as limited life Geomorphol. 8, 125–139.
geotextiles. Carbohyd. Polym. 82 (4), 1090–1096. Pollen-Bankhead, N., Simon, A., 2009. Enhanced application of root-reinforcement al-
Mickovski, S.B., Hallett, P.D., Bransby, M.F., Davies, M.C., Sonnenberg, R., Bengough, gorithms for bank-stability modeling. Earth Surf. Process. Landf. 34 (4), 471–480.
A.G., 2009. Mechanical reinforcement of soil by willow roots: impacts of root prop- Pollen-Bankhead, N., Simon, A., 2010. Hydrologic and hydraulic effects of riparian root
erties and root failure mechanism. Soil Sci. Soc. Am. J. 73 (4), 1276–1285. networks on streambank stability: Is mechanical root-reinforcement the whole story?
Milledge, D.G., Bellugi, D., McKean, J.A., Densmore, A.L., Dietrich, W.E., 2014. A mul- Geomorphology 116 (3–4), 353–362.
tidimensional stability model for predicting shallow landslide size and shape across Preti, F., Giadrossich, F., 2009. Root reinforcement and slope bioengineering stabilization
landscapes. J. Geophys. Res. Earth Surf. 119 (11), 2481–2504. by Spanish Broom (Spartium junceum L.). Hydrol Earth. Syst Sci. 13 (9), 1713–1726.
Mitchell, A.R., Ellsworth, T.R., Meek, B.D., 1995. Effect of root systems on preferential Prieto, J.A., Journeay, M., Acevedo, A.B., Arbelaez, J.D., Ulmi, M., 2018. Development of
flow in swelling soil. Commun. in Soil Sci. Plant Anal. 26 (15–16), 2655–2666. structural debris flow fragility curves (debris flow buildings resistance) using mo-
Morgenstern, N.U., Price, V.E., 1965. The analysis of the stability of general slip surfaces. mentum flux rate as a hazard parameter. Eng. Geol. 239, 144–157.
Geotechnique 15 (1), 79–93. Qi, J., Vermeer, P.A., Cheng, G., 2006. A review of the influence of freeze-thaw cycles on
Ng, C.W.W., Leung, A.K., Woon, K.X., 2014. Effects of soil density on grass-induced soil geotechnical properties. Permafrost Periglac. Process. 17 (3), 245–252.
suction distributions in compacted soil subjected to rainfall. Can. Geotech. J. 51 (3), Raats, P.A., Smiles, D.E., Warrick, A.W., 2002. Contributions to environmental me-
311–321. chanics: Introduction. In: Washington DC American Geophysical Union Geophysical
Ng, C.W.W., Liu, H.W., Feng, S., 2015. Analytical solutions for calculating pore-water Monograph Series, pp. 1–28.
pressure in an infinite unsaturated slope with different root architectures. Can. Radtke, A., Toe, D., Berger, F., Zerbe, S., Bourrier, F., 2014. Managing coppice forests for
Geotech. J. 52 (12), 1981–1992. rockfall protection: lessons from modeling. Ann. For. Sci. 71 (4), 485–494.
Ng, C.W.W., Kamchoom, V., Leung, A.K., 2016a. Centrifuge modelling of the effects of Reddy, N., Yang, Y., 2005. Biofibers from agricultural byproducts for industrial appli-
root geometry on transpiration-induced suction and stability of vegetated slopes. cations. Trends Biotechnol. 23 (1), 22–27.
Landslides 13 (5), 925–938. Reef, R., Slot, M., Motro, U., Motro, M., Motro, Y., Adame, M.F., Garcia, M., Aranda, J.,
Ng, C.W., Menzies, B.K., 2007. Unsaturated Soil Mechanics and Engineering. Taylor & Lovelock, C.E., Winter, K., 2016. The effects of CO 2 and nutrient fertilisation on the
Francis. growth and temperature response of the mangrove Avicennia germinans. Photosynth.
Ng, C.W.W., Woon, K.X., Leung, A.K., Chu, L.M., 2013. Experimental investigation of Res. 129 (2), 159–170.
induced suction distribution in a grass-covered soil. Ecol. Eng. 52, 219–223. Richards, L.A., 1931. Capillary conduction of liquids through porous mediums. Physics 1
Ng, C.W.W., Yu, R., 2014. A novel technique to model water uptake by plants in geo- (5), 318–333.
technical centrifuge. Géotech. Lett. 4 (4), 244–249. Rickli, C., Graf, F., 2009. Effects of forests on shallow landslides–case studies in
Ng, C.W.W., Ni, J.J., Leung, A.K., Wang, Z.J., 2016b. A new and simple water retention Switzerland. For. Snow Landscape Res. 82 (1), 33–44.
model for root-permeated soils. Géotech. Lett. 6 (1), 106–111. Sassa, K., 2019. Foreword by Flavia schlegel for the journal of the international

23
S. Bordoloi and C.W.W. Ng Engineering Geology 275 (2020) 105742

consortium on landslides. Landslides 16 (1) 1-1. Unterscheutz, P., Ruetz, W.F., Geppert, R.R., Ferrell, W.K., 1974. The effect of age, pre-
Schmidt, K.M., Roering, J.J., Stock, J.D., Dietrich, W.E., Montgomery, D.R., Schaub, T., conditioning, and water stress on the transpiration rates of Douglas-Fir (Pseudotsuga
2001. The variability of root cohesion as an influence on shallow landslide suscept- menziesii) seedlings of several ecotypes. Physiol. Plant. 32 (3), 214–221.
ibility in the Oregon Coast Range. Can. Geotech. J. 38 (5), 995–1024. van Genuchten, M.T., 1980. A closed-formequation for predicting the hydraulic con-
Seneviratne, S.I., Nicholls, N., Easterling, D., Goodess, C.M., Kanae, S., Kossin, J., Luo, Y., ductivity of unsaturated soils. Soil Sci. Soc. Am. J. 44 (5), 892–898.
Marengo, J., McInnes, K., Rahimi, M., Reichstein, M., 2017. Changes in climate ex- Van Westen, C.J., Rengers, N., Soeters, R., 2003. Use of geomorphological information in
tremes and their impacts on the natural physical environment. indirect landslide susceptibility assessment. Nat. Hazards 30 (3), 399–419.
Shao, W., Ni, J., Leung, A.K., Su, Y., Ng, C.W.W., 2017. Analysis of plant root–induced Vannoppen, W., Vanmaercke, M., De Baets, S., Poesen, J., 2015. A review of the me-
preferential flow and pore-water pressure variation by a dual-permeability model. chanical effects of plant roots on concentrated flow erosion rates. Earth-Sci. Rev. 150,
Can. Geotech. J. 54 (11), 1537–1552. 666–678.
Shen, P., Zhang, L.M., Chen, H.X., Gao, L., 2017. Role of vegetation restoration in miti- Veylon, G., Ghestem, M., Stokes, A., Bernard, A., 2015. Quantification of mechanical and
gating hillslope erosion and debris flows. Eng. Geol. 216, 122–133. hydric components of soil reinforcement by plant roots. Can. Geotech. J. 52 (11),
Simon, A., Collison, A.J., 2002. Quantifying the mechanical and hydrologic effects of 1839–1849.
riparian vegetation on streambank stability. Earth Surf. Process. Landforms 27 (5), Waldron, L.J., Dakessian, S., 1981. Soil reinforcement by roots: calculation of increased
527–546. soil shear resistance from root properties. Soil Sci. 132 (6), 427–435.
Simon, K., Steinemann, A., 2000. Soil bioengineering: challenges for planning and en- Wang, W., Vinocur, B., Altman, A., 2003. Plant responses to drought, salinity and extreme
gineering. J. Urban Plan. Dev. 126 (2), 89–102. temperatures: towards genetic engineering for stress tolerance. Planta 218 (1), 1–14.
Singh, S., Prakash, A., Garg, A., Hazra, B., Kumar Das, G., 2018. Stochastic modeling of Wang, H., Zhang, L., Yin, K., Luo, H., Li, J., 2020. Landslide identification using machine
relative permeability for vegetated covers. Int. J. Geomech. 18 (9), 06018020. learning. Geosci. Front.
Smith, I., Snow, M.A., 2008. Timber: an ancient construction material with a bright fu- Williams, A.P., Abatzoglou, J.T., Gershunov, A., Guzman-Morales, J., Bishop, D.A., Balch,
ture. For. Chron. 84 (4), 504–510. J.K., Lettenmaier, D.P., 2019. Observed impacts of anthropogenic climate change on
Song, L., Li, J.H., Zhou, T., Fredlund, D.G., 2017. Experimental study on unsaturated wildfire in California. Earth's Future 7 (8), 892–910.
hydraulic properties of vegetated soil. Ecol. Eng. 103, 207–216. Wilson, A.J., Robards, A.W., Goss, M.J., 1977. Effects of mechanical impedance onroot
Sonnenberg, R., Bransby, M.F., Hallett, P.D., Bengough, A.G., Mickovski, S.B., Davies, growth in barley, Hordeum vulgare L. II. Effects on cell development in seminal roots.
M.C.R., 2010. Centrifuge modelling of soil slopes reinforced with vegetation. Can. J. Exp. Bot. 28, 1216–1227.
Geotech. J. 47 (12), 1415–1430. Wong, S.C., Cowan, I.R., Farquhar, G.D., 1979. Stomatal conductance correlates with
Sonnenberg, R., Bransby, M.F., Bengough, A.G., Hallett, P.D., Davies, M.C.R., 2012. photosynthetic capacity. Nature 282 (5737), 424–426.
Centrifuge modelling of soil slopes containing model plant roots. Can. Geotech. J. 49 Wu, T.H., McKinnell III, W.P., Swanston, D.N., 1979. Strength of tree roots and landslides
(1), 1–17. on Prince of Wales Island, Alaska. Can. Geotech. J. 16 (1), 19–33.
Steudle, E., 2001. The cohesion-tension mechanism and the acquisition of water by plant Wynn, T.M., Mostaghimi, S., 2006. Effects of riparian vegetation on stream bank subaerial
roots. Ann. Rev. Plant Biol. 52 (1), 847–875. processes in southwestern Virginia, USA. Earth Surf. Process. Landforms 31 (4),
Stoffel, M., Wilford, D.J., 2012. Hydrogeomorphic processes and vegetation: disturbance, 399–413.
process histories, dependencies and interactions. Earth Surf. Process. Landforms 37 Yamamoto, Y., Springman, S.M., 2019. Triaxial stress path tests on artificially prepared
(1), 9–22. analogue alpine permafrost soil. Can. Geotech. J. 56 (10), 1448–1460.
Stoffel, M., Lièvre, I., Conus, D., Grichting, M.A., Raetzo, H., Gärtner, H.W., Monbaron, Yildiz, A., Askarinejad, A., Graf, F., Rickli, C., Springman, S.M., 2015. Effects of roots and
M., 2005. 400 years of debris-flow activity and triggering weather conditions: mycorrhizal fungi on the stability of slopes.
Ritigraben, Valais, Switzerland. Arct. Antarct. Alp. Res. 37 (3), 387–395. Yildiz, A., Graf, F., Rickli, C., Springman, S.M., 2018. Determination of the shearing be-
Stoffel, M., Conus, D., Grichting, M.A., Lièvre, I., Maître, G., 2008. Unraveling the pat- haviour of root-permeated soils with a large-scale direct shear apparatus. Catena 166,
terns of late Holocene debris-flow activity on a cone in the Swiss Alps: chronology, 98–113.
environment and implications for the future. Glob. Planet. Change 60 (3–4), Yildiz, A., Graf, F., Springman, S.M., 2020. On the dilatancy of root-permeated soils under
222–234. partially saturated conditions. Géotechnique Letters 10, 1–4.
Stokes, A., Salin, F., Kokutse, A.D., Berthier, S., Jeannin, H., Mochan, S., Dorren, L., Zhang, C.B., Chen, L.H., Liu, Y.P., Ji, X.D., Liu, X.P., 2010. Triaxial compression test of
Kokutse, N., Ghani, M.A., Fourcaud, T., 2005. Mechanical resistance of different tree soil–root composites to evaluate influence of roots on soil shear strength. Ecol. Eng.
species to rockfall in the French Alps. Plant Soil 278 (1–2), 107–117. 36 (1), 19–26.
Stokes, A., Atger, C., Bengough, A.G., Fourcaud, T., Sidle, R.C., 2009. Desirable plant root Zhang, H., Liu, Z., Chen, H., Tang, M., 2016. Symbiosis of arbuscular mycorrhizal fungi
traits for protecting natural and engineered slopes against landslides. Plant Soil 324 and Robinia pseudoacacia L. improves root tensile strength and soil aggregate sta-
(1–2), 1–30. bility. PLoS ONE 11 (4), 0153378.
Świtała, B.M., Wu, W., 2018. Numerical modelling of rainfall-induced instability of ve- Zhou, W., Chen, H., Zhou, L., Lewis, B.J., Ye, Y., Tian, J., Li, G., Dai, L., 2011. Effect of
getated slopes. Géotechnique 68 (6), 481–491. freezing-thawing on nitrogen mineralization in vegetation soils of four landscape
Świtała, B.M., Askarinejad, A., Wu, W., Springman, S.M., 2017. Experimental validation zones of Changbai Mountain. Ann. For. Sci. 68 (5), 943–951.
of a coupled hydro-mechanical model for vegetated soil. Géotechnique 68 (5), Zhu, H., Zhang, L.M., 2016. Field investigation of erosion resistance of common grass
375–385. species for soil bioengineering in Hong Kong. Acta Geotechnica 11 (5), 1047–1059.
Świtała, B.M., Askarinejad, A., Wu, W., Springman, S.M., 2018. Experimental validation Zhu, H., Zhang, L.M., Xiao, T., Li, X.Y., 2017. Enhancement of slope stability by vege-
of a coupled hydro-mechanical model for vegetated soil. Géotechnique 68 (5), tation considering uncertainties in root distribution. Comput. Geotech. 85, 84–89.
375–385. Zhu, H., Zhang, L.M., Garg, A., 2018. Investigating plant transpiration-induced soil suc-
Tardío, G., Mickovski, S.B., 2016. Implementation of eco-engineering design into existing tion affected by root morphology and root depth. Comput. Geotech. 103, 26–31.
slope stability design practices. Ecol. Eng. 92, 138–147. Zuazo, V.H.D., Pleguezuelo, C.R.R., 2009. Soil-erosion and runoff prevention by plant
Thomas, R.E., Pollen-Bankhead, N., 2010. Modeling root-reinforcement with a fiber- covers: a review. In: Sustainable Agriculture. Springer, Dordrecht, pp. 785–811.
bundle model and Monte Carlo simulation. Ecolo. Eng. 36 (1), 47–61.

24

You might also like