You are on page 1of 17

Engineering Fracture Mechanics 247 (2021) 107675

Contents lists available at ScienceDirect

Engineering Fracture Mechanics


journal homepage: www.elsevier.com/locate/engfracmech

Crackling noise and bio-cementation


Lei Wang a, Xiang Jiang a, b, *, Xiang He a, Jian Chu c, Yang Xiao a, Hanlong Liu a,
Ekhard K.H. Salje b
a
School of Civil Engineering, Chongqing University, 400045 Chongqing, People’s Republic of China
b
Department of Earth Sciences, Cambridge University, Cambridge CB2 3EQ, UK
c
School of Civil and Environmental Engineering, Nanyang Technological University, Singapore

A R T I C L E I N F O A B S T R A C T

Keywords: Microbially induced calcite precipitation (MICP) is used for bio-cementation of calcareous sand
Bio-cemented using microbial metabolism to generated calcite precipitation. This technology has been proposed
Acoustic emission for sealing damage in geological reservoirs, repair of cracks in stone buildings, and strengthening
Calcareous sands
of foundations in coastal engineering. To test the stability and crash dynamics, we compress bio-
Mean filed theory
Probability distribution
cementation samples under uniaxial stress and investigate their collapse mechanism by
measuring their acoustic emission, AE. AE of the bio-cementation sample shows that the collapse
commences even under very modest stress. The collapse proceeds via avalanches where a local
collapse triggers other damage in the sample. All local collapse processes are highly correlated.
The characteristic avalanche parameters (energy, amplitude, interevent time, etc.) are power law
distributed, e.g. the energy exponents are ε = 1.35–1.6, depending on the degree of the collapse.
Sands without cementation show equally avalanches with energy exponents near ε = 1.7 and very
low strength. In contrast, compressed sand grains have high strength and energy exponents near ε
= 1.4. Correlations between AE spectra show that compressed bio-cementation samples imme­
diately reconvert to sand with some larger grains resisting. The distribution of low AE energies in
bio-cementation samples has the same avalanche exponents as sand. This demonstrates that some
parts of the bio-cementation sample behave exactly as sand with some consolidating, bio-
cemented structures in between. These structures are destroyed and local grains collapse under
further compression. The distribution of other avalanche characteristics like inter-event times, the
Omori’s law, and the Båth’s law, are identical for bio-cementation samples, sands and grain.

1. Introduction

In the natural environment, the production of calcite through bacteria from soil is common [1,2] and dramatically alters the
physical properties of porous media [3]. On the surface of Earth, a large portion of the insoluble calcite is of biogenic origin [4], and
microbial carbonate precipitation, MCP, is significant in soil, caves, sediments, and open-water areas [5,6]. Promising aspects of bio-
mineralization were showed in several applications, including metal stabilization, and hazardous materials encapsulation in natural
soil [7,8]. Plugging with biogenic minerals is a potential way to reduce the permeability of a porous stratum [9] or remediate con­
struction cracks [10]. Microbially induced calcite precipitation, MICP, has received the primary attention in bio-geotechnical engi­
neering in recent years due to its ability to enhance the mechanical properties (e.g. strength, stiffness) and bearing capacity of granular

* Corresponding author at: School of Civil Engineering, Chongqing University, 400045 Chongqing, People’s Republic of China.
E-mail address: cqujiangxiang@163.com (X. Jiang).

https://doi.org/10.1016/j.engfracmech.2021.107675
Received 22 November 2020; Received in revised form 5 March 2021; Accepted 10 March 2021
Available online 16 March 2021
0013-7944/© 2021 Elsevier Ltd. All rights reserved.
L. Wang et al. Engineering Fracture Mechanics 247 (2021) 107675

a b
100μm

5mm 400μm

Fig. 1. (a) Calcareous sands; and (b) SEM of a calcareous sand grain with porous structure.

soils by binding soil particles together by nucleation and growth of calcite minerals [11,12]. Laboratory tests have demonstrated that
MICP treatment of granular soils improves their strength by three orders of magnitude [9,13], increases small-strain stiffness signif­
icantly (>103 times) [14–16], decreases the hydraulic conductivity of soil by two orders of magnitude [12,13,15], enhances the
liquefaction resistance [17–19], and reduces foundation slip [9,20]. Current in-situ applications of MICP treatment include large-scale
ground improvements [21–24], protection and consolidation of heritage buildings and sculptures [6], stabilization and erosion
mitigation of slopes [25,26], construction of aquaculture ponds [13], and as durable and self-healing construction materials [10,27].
One of the most promising applications for MICP is to cement calcareous sand in ocean engineering. Calcareous sand from porous
marine sediments is abundant in low-latitude tropical areas. Major enrichment areas are the South China Sea, the Persian Gulf, Red
Sea, coastlines of Australia and Hawaiian Islands [28]. To exploit marine resources significant ocean and coastal infrastructure are
being constructed on coral reefs and other calcareous soil deposits [29–31]. In ocean engineering, calcareous sands generally play a
role as the foundation or backfill materials for buildings and coastal infrastructures such as road embankments and oil platforms
[29,32]. The porosity and grain crushing characteristics of calcareous sand [33,34] require ground reinforcements for the safety of
constructions [28]. A key element for the safety of this bio-cementation depends largely on its stability and crushing dynamics. It is
required to measure the durability and performance during the construction lifetime. In this paper we show that surveillance by
acoustic emission (AE) measurements is a highly sensitive and non-destructive method for the detection of the collapse of bio-
cemented sand, and can reflect the collapse degree of cementation matrix.
Crackling noise under compression has been observed in a large number of different physical systems. Deterioration and collapse
proceeds by discrete events, or jerks, under stress or other applied external fields. Jerk energies and amplitudes are typically power law
distributed and are hence scale invariant over broad ranges of energies between lower and upper cut-offs [35]. Known ‘crackling noise’
includes magnetization processes [36], ferroelectric domain switching [37], martensitic phase transitions [38], dislocation avalanches
in plastic [39], gravitational wave detections [40], avalanches during stellar evolution [41], material failure [35], and earthquakes
[42], to name just a few examples. Crackling noise appears under the absence of characteristic size scales between cut-offs [43]. The
power law exponents are sometimes linked to universality classes and are used as diagnostic parameter to identify underlying physical
processes. Crackling noise appears often to follow mean field behavior with two characteristic energy exponents, 4/3 and 5/3 [35],
while other values where were reported in alloys [44,45]. Even quasi-continuous changes of exponents have been observed in
compression experiments of coal [46,47] and in model simulations [48,49].
The study of failure of porous materials under compression has become a major research activity because it relates to early warning
events before collapse [35]. Typical examples are Earthquakes, where the magnitude distribution is the Gutenberg-Richter (GR)
scaling [50]. The experimental method of choice in laboratory-based experiments (~labquakes) is the detection of acoustic emission
(AE) under applied stress fields [51–54]. AE measurements are very sensitive and detect even weak signals with an AE absolute energy
near 1 aJ (=10− 18 J) [45,55]. A standard laboratory material is mesoporous silica [56–58], coal [59] and other porous minerals like
goethite [60], berlinite [61], corundum [62], and alloys [45,63]. Despite widely different structures between the compression of
minerals and large-scale Earthquakes, homogenous mesoporous quartz mimics earth quakes rather well [64] provided that their
energy exponents (or Gutenberg–Richter b-values) match. In these studies, two energy exponents are prominent. The lowest value is ε
= 4/3 or b = 1/2. This value is well represented in synthetic Vycor, Gelsil and a multitude of minerals. This value is much smaller than
b = 1 as accepted as dominant value in Earthquakes. This value corresponds to ε = 5/3 which is close to values in coal, sandstone,
ferroelectrics and many alloys. A continuous shift between ε = 5/3 and ε = 4/3 during the compression event was observed in coal [65]
and, most recently, during ferroelectric switching of BaTiO3. These two values of ε were predicted by mean field theory [35] on
whether the distribution of avalanche energies corresponds to a small region near the critical point or to a cumulate distribution over
the whole range of the driving field. In this study, we identify AE for the detection of crackling noise as a preferred, highly sensitive
surveillance method also for bio-cemented calcareous sand. We argue that the collapse of grain and sands without cementation
corresponding to ε = 4/3 and ε = 5/3 respectively, and the energy exponent of bio-cemented sample fluctuates between them
depending on the degree of the collapse. Through the AE avalanche distribution and AE spectra correlations, we also demonstrated that
partial bio-cemented samples reconvert to sand under stress, and its distribution of low AE energies share avalanche exponents as sand.

2
L. Wang et al. Engineering Fracture Mechanics 247 (2021) 107675

Fig. 2. (a) Compression arrangement for grain crushing (plane view); (b) partial enlarged drawing (plane view); and (c) sketch for grain crushing
(elevation view).

2. Experimental methods

Calcareous sands (Fig. 1a) were collected from Yongxing Island in the South China Sea, P.R. China. Their composition was mainly
aragonite and Mg-bearing calcite, the carbonate component was above 97%. The porosity was 23%, similar to Vycor [66]. The SEM
shows the sand grain with porous structure, see Fig. 1b. Three types of tests were carried out, uniaxial compression was applied to
calcareous sand grains, sands without cementation, and bio-cemented sand samples treated by MICP. For the compression tests, we
used the Deben micro-test loading system (Fig. 2) for grain crushing, with a loading rate of 0.01 mm/min. The AE sensor was mounted
on the surface of the static part of the steel frame. The main background noise of the test came from the motor driving the loading
station. We avoided the noise signal by setting an appropriate threshold value. To determine the threshold, we compressed rubber with
no appreciable acoustic emission signals. We increased the threshold value until no background noise could be detected. In this way a
threshold value of 40 dB was determined. The AE signals were recorded by the DISP acoustic emission workstation from the American
Physical Acoustics Company (PAC). Compression of calcareous sands without cementation (Fig. 3) was also investigated using the
Deben micro-test loading system and the same AE detection arrangement as for grain crushing was used. A bio-cemented sand sample
was prepared by the MICP process using natural calcareous sands as shown in Fig. 4a-d. The microbe used in our study was the urease
strain of sporosarcina pasteurii. The MICP process consisted of three steps: (1) synthesis of the enzyme by metabolism of microbe [67],
(2) the production of ammonia and carbonate ions though microbial hydrolyses urea, and (3) the generation of insoluble calcite by
binding of carbonate ions with calcium [68]. The accumulated insoluble calcite bridges calcareous sand grains. The SEM image of a
bio-cemented sand sample is shown in Fig. 4e and 4f and the local meso-structure indicates the bridging of sand particles by microbial
induced calcite. This MICP process was adopted to prepare bio-cemented sand cylindrical samples with diameter 50 mm and length
100 mm as shown in Fig. 5a and b. For details of the MICP process, see [29]. For the bio-cemented sand samples, uniaxial compression
experiments were carried by an electronic servo pressure testing machine shown in Fig. 5c. The maximum compression capacity of this
machine was 250 kN, and the precision was ±0.5%. Uniaxial compression was applied to a bio-cemented sand sample with a rate of
0.1 mm/min. Acoustic emission sensors were fixed onto the specimen’s surface directly by plastic bands. All experiments showed
excellent reproducibility.

3
L. Wang et al. Engineering Fracture Mechanics 247 (2021) 107675

Fig. 3. (a) Sketch of the compression of sands without cementation (elevation view); and (b) photo of compression arrangement for loose sands
without cementation (plane view).

3. Results

3.1. AE spectra and energy distribution

Stress–strain relationships and AE spectra for calcareous grains (Fig. 6a), bio-cemented sand sample (Fig. 6b), and sands without
cementation (Fig. 6c) are shown in Fig. 6. The compressive strength of sands is very weak and cannot be detected by our loading
system. As a consequence, only the AE spectra for sands without cementation are shown in Fig. 6c. The AE parameters such as energies
E, amplitudes A, and waiting time δ were extracted from the original AE spectra as described by Salje and Dahmen [35]. The acoustic
emission energy spectra of the three samples all show the typical characteristics of crackling noise——discrete events with a variety of
sizes and energies. Grain and bio-cemented samples generate many events with energies larger than 105 aJ, while no energy larger than
that value happened in the compression of sands without cementation. For grain crushing, there were few AE signals in the initial
elastic compression. ‘Silent’ episodes in the AE energy were found in the spectra of bio-cemented samples. The strain–stress curve of
sand grains is almost stationary besides the initial incubation time of the elastic compression regime. This phenomenon is consistent
with results of the grain crushing of calcareous sand particles [69]. This is not the case for the bio-cemented sand sample. Here we see
no incubation time. Instead we observe silent intervals with no AE activity (Fig. 6b). The silent intervals are related to the displacement
loading control method. When the sample is damaged, cracks are generated, and elastic energy is released. At the end of the stress drop
the crack is fully established. When the displacement increases further the crack will be gradually closed and no further energy is
transferred. No AE signal is generated at this silent interval. This gradual opening and closing of pores are very strong in bio-cemented
samples where the silent period is most obvious. This effect is not expected in force driven experiments where the control parameter is
the external stress and not the external strain (or displacement). Unlike grain and bio-cemented sample, AE energy spectra of sands
without cementation appear evenly during the loading process.
The probability distribution function (PDF) of AE energies is shown in Fig. 7a. This function describes the probability of finding a
signal with energy between E and E + dE. It follows a power law P(E) ~ E− ε, where ε is the energy exponent. Fig. 7a shows the

4
L. Wang et al. Engineering Fracture Mechanics 247 (2021) 107675

Fig. 4. (a)-(d) Sketch of the MICP (microbial induced calcite precipitation) process; (e) SEM photograph of bio-cemented sand sample treated by
MICP method; and (f) local enlargement shows the microbial induced calcite on the surface of calcareous sand grains forming calcite bridges
connecting sand grains.

distribution of AE energy from grains, bio-cemented sand samples, and sands with exponents of 1.37, 1.46, and 1.7, respectively. To
investigate the exponents in more detail, we use the maximum likelihood method, ML [70]. The maximum likelihood method avoids
the choice of bins and the construction of histograms. If the data follow a power law, the ML function shows a plateau at the exponent ε.
Fig. 7b shows the relevant ML curves. The horizontal dashed lines indicate the PDF curve’s slope, which agree well with the plateau of
ML curves. The ML curves of bio-cemented sand samples show an increasing trend with increasing Emin instead of a flat plateau. This
change of ML-shapes indicates whether the captured signals correspond to pure crackling noise or whether it is affected by damping or
a mixture of several crackling noise mechanisms [71]. The grain shows flat plateaus with a single fix point ε = 1.4 over more than four
decades. This result is identical with previous results on Vycor and very close to the MF value 4/3 (b = 0.5). The energy exponent for
sands is 1.7, which agrees with another MF prediction value 5/3.
For grain crushing and compacted sands, the exponents from different time windows are identical (Fig. 8a). This result contrasts
with the ML analysis of bio-cemented sand samples. The overall energy exponent is ε = 1.46 while different time windows show a
broad range of values (Fig. 8b). The estimated exponents lay between 1.35 and 1.6. Such values are identified with the crushing of
grains and friction of mesoscopic arrangements. Crack propagation leads typically to energy exponents near the MF value 1.33 while
the friction of grains (sand particles) leads to much greater values. Curves with ε < 1.46 are shown in red. The AE spectra of these low
exponents occur often near compressive stress drops confirming fracture events, Fig. 8c. Near large stress drops an even lower energy
exponent of 1.33 = 4/3 occurs, which equates with the theoretical MF value. Yellow curves indicate high exponent values in the
friction limit [72]. Blue curves exhibit a mixed crossover behavior [71].

5
L. Wang et al. Engineering Fracture Mechanics 247 (2021) 107675

Fig. 5. (a) Process diagram of bio-cementation; (b) photograph of cylindrical bio-cemented sand sample with 50 mm diameter and 100 mm length;
and (c) compression arrangement for bio-cemented sand sample.

3.2. Mean field exponents

The meso-structure affects mainly the energy distribution exponent ε, which, in turn, changes other exponents via the scaling τ-1 =
x(ε-1)=(α-1)/ χ = 0.66 in mean field theory, and τ-1 = x(ε-1)=(α-1)/ χ = 1.33 in a force integrated mean field model. τ is the exponent
for the power law distribution of AE amplitude, P(A) ~ A− τ. For the distribution of AE event duration, P(D) ~ D− α, the exponent is α. ×
is the exponent relating AE amplitudes to energies E ~ Ax, and χ relates amplitude to duration A ~ Dχ. For the MF model, τ = 1.67, ε =
1.33, α = 2.0, x = 2.0, and χ = 1.5. For force integrated mean field model, τ = 2.33, ε = 1.67, α = 3.0, x = 2.0, and χ = 1.5. Fig. 9 show
these parameters. For the relation between amplitude and duration, conditional average amplitude <A|D> has been adopted, which is
performed over different duration sub-sets and different samples. We find that the result for grains agrees with the predictions of MF
while bio-cemented sand sample and compacted sands agree better with the force integrated MF. The correlation <A|D>f ollows the
predictions of the ‘half-moon’ shaped probability function proposed by Casals et al [73] and confirm the overall mean field scaling.

3.3. Robust distributions

Unlike the energy distribution, previous studies showed that some distributions, like waiting times, Omori’s law, and the Båth’s law
were little affected by the structural details of materials [74,75]. Fig. 10 shows the distribution of waiting times normalized by their
average <δ>. After normalization, waiting time distributions collapse into a double power-law, with a universal scaling law [56]. The
waiting time results of the all samples are almost identical and agree with the previous study of SiO2 based materials [57]. This
observation confirms that the waiting time distribution is extremely robust and does not change much with the collapse mechanism
[74]. Similarly, aftershocks (AS) sequences are described by Omori’s law [76], which states that the reduction of aftershocks follow a
power-law in time measured after the occurrence of the mainshock (MS). The sequence of aftershocks continues until another
mainshock is found, which ends the previous aftershock sequence. The relation between the aftershock rate and the lapse time since
the mainshock is also power law distributed. The power law exponent p, with the value near to 1 (Fig. 11), is similar to sandstone
collapse under uniaxial or confining stress [77]. The Båth’s law states that the average logarithmic ratio of main shock energy and its
the largest aftershock energy, AS*, is around 1.2. The ratio or ‘relative magnitude’ is defined as <ΔM> = lg(EMS/EAS*). Fig. 12 shows
the logarithmic ratio of energy, the result agrees well with the Båth’s law with no significant deviations.

4. Discussion

The probabilistic description reflects some fundamental aspects of complex systems, usually described in terms of power-laws. In
our study, the bio-cemented sand sample show different exponents because it combines crack propagation and friction effects with 1.3
and 1.6 as typical values for these two mechanisms [35,64,72]. The bio-cemented sand sample used in this study is prepared by bio-
cementation with the same chemical composition as the sand particles. The main difference between these two samples is the degree of
heterogeneity. Porous sand particles follow the pathway of crack propagation and ML curves similar to those of Vycor. They are part of
the bio-cemented sand sample material, which contains in addition heterogeneities at a larger length scale, namely the ‘glue’ between
the sand grains. During the compression of bio-cemented sand sample, the fixed point of the crack propagation and friction effect can
affect each other. Fig. 13 shows the probability density function of bio-cemented sample for different energy intervals. When energies
smaller than 1000 aJ, the slope is about 1.6 which near the result of sand. While the slope is similar with grain crushing in large energy
area, which is about 1.4. In order to study if the small energy works in the whole compression process or just comes from the beginning

6
L. Wang et al. Engineering Fracture Mechanics 247 (2021) 107675

Fig. 6. (a), and (b) compression stress–strain curve and AE energy spectrum for grains and a bio-cemented sand sample. (c) AE energy spectrum of
sand. Note that crack propagation in (b) is interrupted by ‘silent’ episodes.

friction, the activity (rate) of small AE energy has been shown in Fig. 14. We find that the small energies happen in the whole
compression process. The avalanche of bio-cemented sample is recognized as a combination of the AE signals of sands and grain. Under
stress, bio-cemented samples crumble and turn to sand again. Fig. 15 shows the evolution of bio-cemented sample collapse. It shows
that the breakage zone has changed into sands and even generates a small sandpile. The cumulative size distribution of fragment
smaller than 1 mm after the collapse of bio-cemented sample, and size distribution of initial natural sands have been compared in
Fig. 16. The size distribution of bio-cemented sample’s fragment is similar with that of natural sands, and this supports the AE

7
L. Wang et al. Engineering Fracture Mechanics 247 (2021) 107675

Fig. 7. (a) Distribution of AE energies, (b) shows the ML-fitted exponent as a function of the lower threshold Emin for three experiments during the
full duration of the experiment.

prediction for the stability and collapse evolution of bio-cementation. In a typical mixing situation, the ML curve shows the hori­
zontally arranged S shape, the maximum exponent value associates with the lower bound of the high exponent, and the gradual
minimum value is near the low exponent resource [71]. The mixing ML curve is controlled by the two exponents values and the ratio of
data points for the two subsystems. In the ML estimation of bio-cemented sample, we do not obtain the typical mixing curves, namely
horizontally arranged S shape. This is because even though the bio-cemented sample crumbles and becomes sand again in the
compression, there are not enough small-energy data to produce the mixing ML. This can be explained as the cementation effect limits
the rotation and rolling of particles in bio-cemented sample, and making them not as free as the particles in the sands. Nevertheless, if
we increase the amount of small energy data artificially, we can get the typical mixing ML estimation. The red points in Fig. 17 is the
ML estimate for the original bio-sample data with a doubled signal for energies smaller than 100aJ. This combination exhibits the
typical mixing ML shape. On the other hand, the energies come from sands and grain can produce the result of bio-cemented sample.
Fig. 17 shows the ML estimate for the superposition of sands data and grain data (green points). This new data set reproduces the bio-
cemented sample very well. Our bio-cemented sand sample shows a low exponent (red curves in Fig. 8b) from force drop intervals,
where crack propagation dominates. The effective ML curves from different time intervals is located between the two typical fracture
and friction fix points (1.33 to 1.66). The sandstone creep compression shows that with the increasing number of low exponent data,
the superposition ML curve shifts towards the low exponent [78]. Similar results are found in compression of granular Mg-Ho alloys
[63], porous stainless [79], and acoustic emission from confined sandstone under compression [77].

5. Implications

The calcareous sand used in this paper is widely distributed in the low altitude areas of the world. The microbial reinforcement
method is directly relevant for the engineering design of constructions on calcareous sand. The MICP method is widely used to reduce

8
L. Wang et al. Engineering Fracture Mechanics 247 (2021) 107675

Fig. 8. (a) ML curves of a grain, and calcareous sands in different time windows; (b) the ML curves of bio-cemented sand sample in different time
windows; and (c) AE spectra of a bio-cemented sand sample. We use the same color scheme for ML curves and AE data.

9
L. Wang et al. Engineering Fracture Mechanics 247 (2021) 107675

Fig. 9. (a) Correlation between the energies and the amplitudes shows power law with E ~ Ax; (b) PDF of amplitude A; (c) PDF of duration D; and
(d) Correlation between the conditional average amplitude and the duration shows power law with <A|D> ~ Dχ.

the permeability of rock and repair the cracks in the surrounding rock of underground energy storage and nuclear waste storage [80].
The uniaxial compressive strength was improved by the MICP [81]. With the increase of microbial reinforcement technology and
engineering applications, the new demand is safety monitoring.
Safety monitoring can be based on our results. The AE method operates at real-time and is non-destructive. It was shown [59] that
crackling noise shifts its critical power law exponents when approaching a critical point, such as the collapse of the material. This result
agrees with computer simulations [48,49] and risk assessment of nuclear waste deposits [82]. Furthermore, the relationships between
crackling noise and force drops was confirmed by Zreihan et al. [83]. Here we have shown that risk assessment of bio-engineered sand
is possible using the following methodology. First one continuously measures the emission of acoustic noise of the construction. The
noise is then continuously analyzed via the determination of its energy exponent. If the energy exponent is near 1.4 the failure
mechanism is crack propagation including the fraction of sand particles and the structural damage is limited. If the signal becomes
larger and shows large damping, then the overall MICP construction is failing. If the exponent increases to values near 1.77, the
structure is in immediate danger of global failure.

Declaration of Competing Interest

The authors declare that they have no known competing financial interests or personal relationships that could have appeared to
influence the work reported in this paper.

10
L. Wang et al. Engineering Fracture Mechanics 247 (2021) 107675

Fig. 10. Distributions of waiting times after normalized by its average 〈δ〉 for different values of Emin for (a) grain; (b) sands; and (c) bio-
cemented Sample.

11
L. Wang et al. Engineering Fracture Mechanics 247 (2021) 107675

Fig. 11. Distributions of aftershocks for (a) grain; (b) sands; and (c) bio-cemented Sample. Number of aftershocks (AS) per unit time, rAS , as a
function of the time distance to the main shock (t − tms ). The main shock is locateed in the energy range described in the legend.

12
L. Wang et al. Engineering Fracture Mechanics 247 (2021) 107675

Fig. 12. Båth’s law for different samples. The theoretical relative magnitude of the Båth’s law is 1.2 which is rather well reproduced.

Fig. 13. Probability density function, P(E), of bio-cemented sample for different energy intervals, N indicates the AE data number in that energy set.

Fig. 14. AE activity (rate) for different energy sets. Dashed curve indicates the accumulate rate.

13
L. Wang et al. Engineering Fracture Mechanics 247 (2021) 107675

Fig. 15. Evolution of bio-cemented sample collapse, the breakage zone has crumbled into sands.

Fig. 16. Particle size distribution of initial natural sands and fragments of bio-cemented sample after loading.

14
L. Wang et al. Engineering Fracture Mechanics 247 (2021) 107675

Fig. 17. ML estimation for original bio-cemented sample data (blue points), and man-made data sets. Green points come from the ML estimation for
Union of sands energy data and grain energy data. Red points show the ML estimation for original bio-sample data add to double signal of energy
smaller than 100aJ. (For interpretation of the references to color in this figure legend, the reader is referred to the web version of this article.)

Acknowledgments

We thank the financial support from the Natural Science Foundation of China (Nos. 51908088, 41831282, and 51922024), and
Natural Science Foundation of Chongqing (cstc2019jcyj-xfkxX0001, cstc2019jcyjjq-X0014). E.K.H. Salje is grateful to EPSRC (No. EP/
P024904/1) for support.

References

[1] Boquet E, Boronat A, Ramoscor A. Production of calcite (calcium carbonate) crystals by soil bacteria is a general phenomenon. Nature 1973;246:527–9.
[2] Han Z, Li D, Zhao H, Yan H, Li P. Precipitation of carbonate minerals induced by the halophilic chromohalobacter israelensis under high salt concentrations:
implications for natural environments. Minerals 2017;7:95.
[3] Armstrong R, Ajo-Franklin J. Investigating biomineralization using synchrotron based X-ray computed microtomography. Geophys Res Lett 2011;38.
[4] Gadd GM. Metals, minerals and microbes: geomicrobiology and bioremediation. Microbiol-sgm 2010;156:609–43.
[5] Achal V, Pan XL, Fu QL, Zhang DY. Biomineralization based remediation of As(III) contaminated soil by Sporosarcina ginsengisoli. J Hazard Mater 2012;201:
178–84.
[6] Jroundi F, Schiro M, Ruiz-Agudo E, Elert K, Martin-Sanchez I, Teresa Gonzalez-Munoz M, et al. Protection and consolidation of stone heritage by self-inoculation
with indigenous carbonatogenic bacterial communities. Nat Commun 2017;8:279.
[7] Etemadi O, Petrisor IG, Kim D, Wan MW, Yen TF. Stabilization of metals in subsurface by biopolymers: Laboratory drainage flow studies. Soil Sediment Contam
2003;12:647–61.
[8] Khachatoorian R, Petrisor IG, Kwan CC, Yen TF. Biopolymer plugging effect: laboratory-pressurized pumping flow studies. J Pet Sci Eng 2003;38:13–21.
[9] DeJong JT, Fritzges MB, Nüsslein K. Microbially induced cementation to control sand response to undrained shear. J Geotech Geoenviron 2006;132:1381–92.
[10] De Muynck W, De Belie N, Verstraete W. Microbial carbonate precipitation in construction materials: A review. Ecol Eng 2010;36:118–36.
[11] Terzis D, Laloui L. 3-D micro-architecture and mechanical response of soil cemented via microbial-induced calcite precipitation. Sci Rep 2018;8:1416.
[12] Dejong JT, Soga K, Kavazanjian E, Van Paassen LA, Al Qabany A, Burns S, et al. Biogeochemical processes and geotechnical applications: progress, opportunities
and challenges. Geotechnique 2013;63:287–301.
[13] Chu J, Ivanov V, Naeimi M, Stabnikov V, Li B. Microbial method for construction of an aquaculture pond in sand. Geotechnique 2013;63:871–5.
[14] Feng K, Montoya BM. Influence of Confinement and Cementation Level on the Behavior of Microbial-Induced Calcite Precipitated Sands under Monotonic
Drained Loading. J Geotech Geoenviron 2016;142:04015057.
[15] Martinez BC, DeJong JT, Ginn TR, Montoya BM, Barkouki TH, Hunt C, et al. Experimental optimization of microbial-induced carbonate precipitation for soil
improvement. J Geotech Geoenviron 2013;139:587–98.
[16] Lin H, Suleiman MT, Brown DG, Kavazanjian Jr E. Mechanical Behavior of Sands Treated by Microbially Induced Carbonate Precipitation. J Geotech Geoenviron
2016;142:04015066.
[17] Sasaki T, Kuwano R. Undrained cyclic triaxial testing on sand with non-plastic fines content cemented with microbially induced CaCO3. Soils Found 2016;56:
485–95.
[18] Montoya BM, DeJong JT, Boulanger RW. Dynamic response of liquefiable sand improved by microbial-induced calcite precipitation. Geotechnique 2013;63:
302–12.
[19] Feng K, Montoya BM. Quantifying level of microbialiInduced cementation for cyclically loaded sand. J Geotech Geoenviron 2017;143:06017005.
[20] Montoya BM, Safavizadeh S, Gabr MA. Enhancement of coal ash compressibility parameters using microbial-induced carbonate precipitation. J Geotech
Geoenviron 2019;145:04019018.
[21] van Paassen LA, Ghose R, van der Linden TJM, van der Star WRL, van Loosdrecht MCM. Quantifying biomediated ground improvement by ureolysis: Large-scale
biogrout experiment. J Geotech Geoenviron 2010;136:1721–8.
[22] Gomez MG, Anderson CM, Graddy CMR, Dejong JT, Nelson DC, Ginn TR. Large-scale comparison of bioaugmentation and biostimulation approaches for
biocementation of sands. J Geotech Geoenviron 2017;143:04016124.
[23] Dejong JT, Martinez BC, Ginn TR, Hunt C, Major D, Tanyu B. Development of a scaled repeated five-spot treatment model for examining microbial induced
calcite precipitation feasibility in field applications. Geotech Test J 2014;37:424–35.
[24] Nassar MK, Gurung D, Bastani M, Ginn TR, Shafei B, Gomez MG, et al. Large-scale experiments in microbially induced calcite precipitation (MICP): reactive
transport model development and prediction. Water Resour Res 2018;54:480–500.
[25] Salifu E, Maclachlan E, Iyer KR, Knapp CW, Tarantino A. Application of microbially induced calcite precipitation in erosion mitigation and stabilisation of sandy
soil foreshore slopes: A preliminary investigation. Eng Geol 2016;201:96–105.
[26] Gowthaman S, Mitsuyama S, Nakashima K, Komatsu M, Kawasaki S. Biogeotechnical approach for slope soil stabilization using locally isolated bacteria and
inexpensive low-grade chemicals: A feasibility study on Hokkaido expressway soil. Japan. Soils Found 2019;59:484–99.
[27] Seifan M, Samani AK, Berenjian A. Bioconcrete: next generation of self-healing concrete. Appl Microbiol Biotechnol 2016;100:2591–602.

15
L. Wang et al. Engineering Fracture Mechanics 247 (2021) 107675

[28] Liu L, Liu HL, Xiao Y, Chu J, Xiao P, Wang Y. Biocementation of calcareous sand using soluble calcium derived from calcareous sand. Bull Eng Geol Environ
2018;77:1781–91.
[29] Xiao P, Liu H, Stuedlein AW, Evans TM, Xiao Y. Liquefaction resistance of bio-cemented calcareous sand. Soil Dyn Earthq Eng 2018;107:9–19.
[30] Schwiderski EW. On charting global ocean tides. Rev Geophys 1980;18:243–68.
[31] Singh SC, Carton H, Tapponnier P, Hananto ND, Chauhan APS, Hartoyo D, et al. Seismic evidence for broken oceanic crust in the 2004 Sumatra earthquake
epicentral region. Nat Geosci 2008;1:777–81.
[32] Wang XZ, Jiao YY, Wang R, Hu MJ, Meng QS, Tan FY. Engineering characteristics of the calcareous sand in Nansha Islands. South China Sea. Eng Geol 2011;120:
40–7.
[33] Xiao Y, Liu H, Chen Q, Long L, Xiang J. Evolution of particle breakage and volumetric deformation of binary granular soils under impact load. Granul Matter
2017;19:71.
[34] Coop MR, Sorensen KK, Bodas Freitas T, Georgoutsos G. Particle breakage during shearing of a carbonate sand. Geotechnique 2004;54:157–63.
[35] Salje EKH, Dahmen KA. Crackling noise in disordered materials. Ann Rev Cond Mat 2014;5:233–54.
[36] Perkovic O, Dahmen K, Sethna JP. Avalanches, Barkhausen noise, and plain old criticality. Phys Rev Lett 1995;75:4528–31.
[37] Salje EKH, Xue D, Ding X, Dahmen KA, Scott JF. Ferroelectric switching and scale invariant avalanches in BaTiO3. Phys Rev Mater 2019;3:014415.
[38] Vives E, Ortin J, Manosa L, Rafols I, Perezmagrane R, Planes A. Distributions of avalanches in martensitic transformations. Phys Rev Lett 1994;72:1694–7.
[39] Csikor FF, Motz C, Weygand D, Zaiser M, Zapperi S. Dislocation avalanches, strain bursts, and the problem of plastic forming at the micrometer scale. Science
2007;318:251–4.
[40] Vajente G. Crackling noise in advanced gravitational wave detectors: A model of the steel cantilevers used in the test mass suspensions. Phys Rev D 2017;96:
022003.
[41] Sheikh MA, Weaver RL, Dahmen KA. Avalanche statistics identify intrinsic stellar processes near criticality in KIC 8462852. Phys Rev Lett 2016;117:261101.
[42] Gutenberg B, Richter CF. Seismicity of the earth and associated phenomena. Princeton: Princeton Univ. Press; 1954.
[43] Christensen K, Moloney NR. Complexity and criticality. London: Imperial College Press; 2005.
[44] Salje EKH. Tweed, twins, and holes. Am Mineral 2015;100:343–51.
[45] Soto-Parra D, Zhang XX, Cao SS, Vives E, Salje EKH, Planes A. Avalanches in compressed Ti-Ni shape-memory porous alloys: An acoustic emission study. Phys
Rev E 2015;91:060401(R).
[46] Jiang X, Liu HL, Main IG, Salje EKH. Predicting mining collapse: Superjerks and the appearance of record-breaking events in coal as collapse precursors. Phys
Rev E 2017;96:023004.
[47] Kadar V, Kun F. System-size-dependent avalanche statistics in the limit of high disorder. Phys Rev E 2019;100:053001.
[48] Kun F, Varga I, Lennartz-Sassinek S, Main IG. Approach to failure in porous granular materials under compression. Phys Rev E 2013;88:062207.
[49] Kun F, Varga I, Lennartz-Sassinek S, Main IG. Rupture cascades in a discrete element model of a porous sedimentary rock. Phys Rev Lett 2014;112:065501.
[50] Utsu T. Representation and analysis of the earthquake size distribution: A historical review and some new approaches. Pure Appl Geophys 1999;155:509–35.
[51] Chanard K, Nicolas A, Hatano T, Petrelis F, Latour S, Vinciguerra S, et al. Sensitivity of acoustic emission triggering to small pore pressure cycling perturbations
during brittle creep. Geophys Res Lett 2019;46:7414–23.
[52] Aben FM, Brantut N, Mitchell TM, David EC. Rupture energetics in crustal rock from laboratory-scale seismic tomography. Geophys Res Lett 2019;46:7337–44.
[53] Marty S, Passelegue FX, Aubry J, Bhat HS, Schubnel A, Madariaga R. Origin of high-frequency radiation during laboratory earthquakes. Geophys Res Lett 2019;
46:3755–63.
[54] Hu W, Scaringi G, Xu Q, Huang R. Acoustic emissions and microseismicity in granular slopes prior to failure and flow-like motion: The potential for early
warning. Geophys Res Lett 2018;45. 10,406–10,15.
[55] Navas-Portella V, Corral A, Vives E. Avalanches and force drops in displacement-driven compression of porous glasses. Phys Rev E 2016;94:033005.
[56] Baro J, Corral A, Illa X, Planes A, Salje EKH, Schranz W, et al. Statistical similarity between the compression of a porous material and earthquakes. Phys Rev Lett
2013;110:088702.
[57] Nataf GF, Castillo-Villa PO, Baro J, Illa X, Vives E, Planes A, et al. Avalanches in compressed porous SiO2-based materials. Phys Rev E 2014;90:022405.
[58] Salje EKH, Soto-Parra DE, Planes A, Vives E, Reinecker M, Schranz W. Failure mechanism in porous materials under compression: crackling noise in mesoporous
SiO2. Philos Mag Lett 2011;91:554–60.
[59] Jiang X, Jiang D, Chen J, Salje EKH. Collapsing minerals: Crackling noise of sandstone and coal, and the predictability of mining accidents. Am Mineral 2016;
101:2751–8.
[60] Salje EKH, Lampronti GI, Soto-Parra DE, Baro J, Planes A, Vives E. Noise of collapsing minerals: Predictability of the compressional failure in goethite mines. Am
Mineral 2013;98:609–15.
[61] Nataf GF, Castillo-Villa PO, Sellappan P, Kriven WM, Vives E, Planes A, et al. Predicting failure: acoustic emission of berlinite under compression. J Phys:
Condens Matter 2014;26:275401.
[62] Castillo-Villa PO, Baro J, Planes A, Salje EKH, Sellappan P, Kriven WM, et al. Crackling noise during failure of alumina under compression: The effect of porosity.
J Phys: Condens Matter 2013;25:292202.
[63] Chen Y, Ding XD, Fang DQ, Sun J, Salje EKH. Acoustic emission from porous collapse and moving dislocations in granular Mg-Ho alloys under compression and
tension. Sci Rep 2019;9:1330.
[64] Navas-Portella V, Gonzalez A, Serra I, Vives E, Corral A. Universality of power-law exponents by means of maximum-likelihood estimation. Phys Rev E 2019;
100:062106.
[65] Xu YY, Borrego AG, Planes A, Ding XD, Vives E. Criticality in failure under compression: Acoustic emission study of coal and charcoal with different
microstructures. Phys Rev E 2019;99:033001.
[66] Salje EKH, Koppensteiner J, Schranz W, Fritsch E. Elastic instabilities in dry, mesoporous minerals and their relevance to geological applications. Mineral Mag
2010;74:341–50.
[67] Krajewska B, Ureases I. Functional, catalytic and kinetic properties: A review. J Mol Catal B: Enzym 2009;59:9–21.
[68] Burne RA, Chen YYM. Bacterial ureases in infectious diseases. Microbes Infect 2000;2:533–42.
[69] Xiao Y, Wang L, Jiang X, Evans TM, Stuedlein AW, Liu H. Acoustic emission and force drop in grain crushing of carbonate sands. J Geotech Geoenviron 2019;
145:04019057.
[70] Clauset A, Shalizi CR, Newman MEJ. Power-law distributions in empirical data. Siam Rev 2009;51:661–703.
[71] Salje EKH, Planes A, Vives E. Analysis of crackling noise using the maximum-likelihood method: Power-law mixing and exponential damping. Phys Rev E 2017;
96:042122.
[72] Soto-Parra D, Vives E, Botello-Zubiate ME, Matutes-Aquino JA, Planes A. Acoustic emission avalanches during compression of granular manganites. Appl Phys
Lett 2018;112:251906.
[73] Casals B, Dahmen KA, Gou B, Rooke S, Salje EKH. The duration-energy-size enigma for acoustic emission. Sci Rep 2021;11:5590.
[74] Bak P, Christensen K, Danon L, Scanlon T. Unified scaling law for earthquakes. Phys Rev Lett 2002;88:178501.
[75] Davidsen J, Stanchits S, Dresen G. Scaling and universality in rock fracture. Phys Rev Lett 2007;98:125502.
[76] Utsu T, Ogata Y, Matsuura RS. The centenary of the Omori formula for a decay law of aftershock activity. J Phys Earth 1995;43:1–33.
[77] Zhao YF, Liu HL, Xie KN, Salje EKH, Jiang X. Avalanches in compressed sandstone: Crackling noise under confinement. Crystals 2019;9:582.
[78] Salje EKH, Liu HL, Jin LS, Jiang DY, Xiao Y, Jiang X. Intermittent flow under constant forcing: Acoustic emission from creep avalanches. Appl Phys Lett 2018;
112:054101.
[79] Chen Y, Wang QB, Ding XD, Sun J, Salje EKH. Avalanches and mixing behavior of porous 316L stainless steel under tension. Appl Phys Lett 2020;116:111901.
[80] Zhu TT, Dittrich M. Carbonate precipitation through microbial activities in natural environment, and their potential in biotechnology: A review. Front Bioeng
Biotechnol 2016;4:4.

16
L. Wang et al. Engineering Fracture Mechanics 247 (2021) 107675

[81] Ramachandran SK, Ramakrishnan V, Bang SS. Remediation of concrete using micro-organisms. ACI Mater J 2001;98:3–9.
[82] Xie KN, Jiang X, Jiang D, Xiao Y, Chen SW, Dahmen KA, et al. Change of crackling noise in granite by thermal damage: Monitoring nuclear waste deposits. Am
Mineral 2019;104:1578–84.
[83] Zreihan N, Faran E, Vives E, Planes A, Shilo D. Relations between stress drops and acoustic emission measured during mechanical loading. Phys Rev Mater 2019;
3:043603.

17

You might also like