You are on page 1of 20

| |

Received: 4 April 2022    Revised: 23 June 2022    Accepted: 11 July 2022

DOI: 10.1111/gbi.12516

ORIGINAL ARTICLE

Interplay between abiotic and microbial biofilm-­mediated


processes for travertine formation: Insights from a thermal
spring (Piscine Carletti, Viterbo, Italy)

Stefania Venturi1,2 | Simona Crognale3  | Francesco Di Benedetto4 |


Giordano Montegrossi2 | Barbara Casentini3 | Stefano Amalfitano3  |
Tommaso Baroni1 | Simona Rossetti3 | Franco Tassi1,2 | Francesco Capecchiacci1,2,5 |
Orlando Vaselli1,2 | Stefano Fazi3

1
Department of Earth Sciences, University
of Florence, Florence, Italy Abstract
2
Institute of Geosciences and Earth Active hydrothermal travertine systems are ideal environments to investigate how
Resources (IGG), National Research
Council of Italy (CNR), Florence, Italy
abiotic and biotic processes affect mineralization mechanisms and mineral fabric for-
3
Water Research Institute (IRSA), National mation. In this study, a biogeochemical characterization of waters, dissolved gases,
Research Council of Italy (CNR), Rome, and microbial mats was performed together with a mineralogical investigation on
Italy
4 travertine encrustations occurring at the outflow channel of a thermal spring. The
Department of Physics and Earth
Sciences, University of Ferrara, Ferrara, comprehensive model, compiled by means of TOUGHREACT computational tool from
Italy
5
measured parameters, revealed that mineral phases were differently influenced by
Istituto Nazionale di Geofisica e
Vulcanologia, Sezione di Napoli, either abiotic conditions or microbially driven processes. Microbial mats are shaped
Osservatorio Vesuviano, Naples, Italy by light availability and temperature gradient of waters flowing along the channel.
Correspondence Mineralogical features were homogeneous throughout the system, with euhedral cal-
Francesco Di Benedetto, Department of cite crystals, related to inorganic precipitation induced by CO2 degassing, and calcite
Physics and Earth Sciences, University of
Ferrara, Ferrara, Italy. shrubs associated with organomineralization processes, thus indicating an indirect
Email: francesco.dibenedetto@unife.it microbial participation to the mineral deposition (microbially influenced calcite). The
Stefano Fazi, Water Research Institute microbial activity played a role in driving calcite redissolution processes, resulting in
(IRSA), National Research Council of Italy
circular pits on calcite crystal surfaces possibly related to the metabolic activity of
(CNR), Rome, Italy.
Email: stefano.fazi@irsa.cnr.it sulfur-­oxidizing bacteria found at a high relative abundance within the biofilm commu-
nity. Sulfur oxidation might also explain the occurrence of gypsum crystals embedded
in microbial mats, since gypsum precipitation could be induced by a local increase in
sulfate concentration mediated by S-­oxidizing bacteria, regardless of the overall un-
dersaturated environmental conditions. Moreover, the absence of gypsum dissolution
suggested the capability of microbial biofilm in modulating the mobility of chemical
species by providing a protective envelope on gypsum crystals.

KEYWORDS
biofilms, gypsum, hot spring, microbial mat, travertine

Geobiology. 2022;00:1–20. wileyonlinelibrary.com/journal/gbi© 2022 John Wiley & Sons Ltd.     1 |


|
2      VENTURI et al.

1  |  I NTRO D U C TI O N role of the microbial community assemblage and biofilm structure


at and below the water–­solid interface of travertine encrustations.
Terrestrial hot springs typically harbor rich and diverse microbial A cross-­disciplinary investigation was performed at a spring
communities able to thrive in extreme environmental conditions system from which thermal groundwaters spontaneously outflow
through peculiar adaptation strategies, including the formation of along an artificial channel suitable for travertine formation (Piscopo
heterogeneous syntrophic assemblages and microbial mats, adhe- et al.,  2006). All experimental findings were compared with those
sion to solid surfaces and mineral deposits, and establishment of calculated by a comprehensive model carried out by means of the
complex biotic and abiotic interactions with the surrounding highly TOUGHREACT computational tool.
dynamic geochemical environment (Des Marais & Walter,  2019; We hypothesized that travertine deposition and textural pat-
Schuler et al., 2017). terns will reflect the occurrence and structure of microbial biofilms
Microbial mats are vertically structured microbial communities and active microbial players providing specific biosignatures within
that significantly affect mineral precipitation and dissolution pro- the mineral deposits.
cesses occurring at the water–­s olid interface (e.g., Prieto-­B arajas
et al.,  2018; van Gemerden,  1993). Depending on the role of bi-
ological activity, microbe–­mineral interactions may result in di- 2  |  S T U DY A R E A
verse mineral precipitation processes, as follows: (i) biologically
controlled mineralization, in which biotic processes directly govern Piscine Carletti (hereafter named PC; also referred to as Bullicame
the crystal nucleation and growth shaping mineral morphologies; 3 or Bullicame West; Duchi et al., 1985; Pentecost, 1995; Minissale
(ii) biologically induced mineralization, resulting from the interac- et al., 2002; Di Benedetto et al., 2011; Rimondi et al., 2021) is lo-
tion between metabolic by-­p roducts and the chemical environ- cated in a thermal area sited between the Central Apennines and
ment; and (iii) biologically influenced mineralization (also referred Tyrrhenian coastline, few km west of the town of Viterbo (north-
to as organomineralization), induced by abiotic processes but with ern Latium, Central Italy; Figure  1). This region is characterized
the influence of active organisms on crystal morphology and com- by (i) a Paleozoic–­Triassic metamorphic basement, (ii) Mesozoic–­
position (Dupraz et al.,  2009; Castro-­A lonso et al.,  2019). In the Paleogene sedimentary rocks related to the Apennine orogenesis,
latter case, mineral precipitation is favored by nucleation on bac- and (iii) upper Mio-­Pleistocene sedimentary basins related to a
terially produced organic polymers (Görgen et al., 2021), such as post-­collisional extensional phase (Baiocchi et al., 2012; Piscopo
those forming cell surfaces or extracellular polymeric substances et al.,  2006) during which the Vicano–­C imino Volcanic District
(EPS). Similarly, microbial mats may induce or inhibit mineral dis- (VCVD) formed (Tassi et al., 2015, and references therein). Crustal
solution owing to either a direct result of metabolic activity or thinning, igneous processes, and high heat flow in this area are
metabolically induced chemical gradients at the microscale level responsible for the large number of CO2(H2 S)-­r ich thermomineral
(Wilmeth et al., 2018). springs and anomalously high CO2 diffuse degassing sites (Cinti
Several studies have focused on understanding the biogeo- et al., 2014), whose location is strictly controlled by (i) hydrogeo-
chemical functioning of early life-­forms on Earth, correctly iden- logical settings and (ii) brittle structural elements (Minissale, 2004).
tifying biosignatures in either rock records or extra-­terrestrial The resulting complex geological setting has originated the oc-
materials by gathering paleobiological and paleoecological infor- currence of two main aquifers, as follows: (i) a shallower volcanic
mation (Allen et al., 2000; Banfield et al., 2004; Tang et al., 2014; aquifer, characterized by cold and fresh waters with Ca-­H CO3
Della Porta et al., 2021). Moreover, increasing attention has been composition and low pCO2 values, located within the Pleistocene
paid in the recent years to microbially mediated carbon cap- volcanites, and (ii) a deeper aquifer, located in the deep Mesozoic–­
ture and storage in natural settings (Jansson & Northern,  2010; Cenozoic carbonate rocks, hosting thermal waters at 220°C with
McCutcheon et al., 2014). Disentangling the mutual influence be- a Ca-­SO 4 (HCO3) composition and high pCO2 values, mainly pro-
tween abiotic and biotic environmental components and unrav- duced by thermometamorphic decarbonation processes and
eling their impact on mineralization processes and mineral fabric mantle degassing (Baiocchi et al., 2012; Cinti et al., 2014; Piscopo
structuring have potentially relevant and multifaceted implica- et al.,  2006). The Viterbo geothermal system is located in coin-
tions to better estimate the significance and relevance of biomin- cidence with a structural high of the carbonate basement and a
eralization processes in the global cycles of major biogeochemical geothermal gradient greater than 100°C/km (Piscopo et al., 2006),
elements (Görgen et al., 2021). resulting in CO2-­ and H2 S-­r ich thermal waters, uprising through
The aim of this field study was to determine how and to what faults and fractures, with temperatures ranging around 50–­6 0°C
extent abiotic and biotic environmental factors can influence min- (Cinti et al., 2014; Duchi & Minissale, 1995; Minissale, 2004).
eral precipitation and textural patterns during travertine formation At PC, the hot water emerges at a relatively constant flow rate
at macro-­ and microscale levels. More specifically, we entailed to (i) (0.7 L/s; Di Benedetto et al., 2011) in a constructed pool (ca. 3 m in
link biogeochemical patterns of waters and dissolved gases to min- diameter), where vigorous gas bubbling occurs. Then, it flows along
eralization processes and travertine deposition, and (ii) explore the a narrow (ca. 14 cm) artificial channel, elevated with respect to the
VENTURI et al. |
      3

(a)

(b) (c)

F I G U R E 1  Location of the study area in (a) central Italy, (b) few km west of the town of Viterbo. In (c), the Piscine Carletti spring system is
shown, together with the sampling sites along the channel.

ground level (up to ca. 2 m). The channel ends into a pool where wa- Moreover, an unfiltered water aliquot was sampled in 125-­ml pol-
ters cool down to ambient temperature (Figure 1). The artificial el- yethylene bottles, where few milligrams of HgCl2 was added in
ements are covered by a thick travertine deposit with shrub fabrics the laboratory in order to prevent any microbial activity for the
13
(Di Benedetto et al.,  2011), partially coated by differently colored analysis of the C/12C ratios of the total dissolved inorganic car-
microbial mats. The channel is periodically maintained to guarantee bon (TDIC). Dissolved organic carbon (DOC) was sampled in HCl
the regular downstream water flow. preconditioned PTFE 20-­ml bottles, filtered on 25-­m m carbon
cleaned GF/F filters (combusted at 450°C for 4  h), and acidified
0.2% after sampling with Suprapur HCl. Samples were on site ana-
3  |   M ATE R I A L S A N D M E TH O DS lyzed for NO2− and NH4 + content by portable spectrophotometer
measurements (Hach DR 2800).
3.1  |  Sampling strategy Dissolved gases were collected at each site using pre-­evacuated
250-­ml Pyrex flasks equipped with Thorion® valve and filled with
Water, dissolved gas, biofilm, and travertine sampling at PC was water up to about 3/4 of their inner volume (Tassi et al., 2008, 2009).
carried out in March 2016 in eight selected sites along the artificial The chemical composition was calculated from the composition of
channel, at a distance of ca. 14 m from each other, starting from the the gas phase stored in the headspace of the sampling glass flasks on
farthest site with respect to the water spring and proceeding up- the basis of (i) gas pressure, (ii) headspace volume, and (iii) the solu-
stream (i.e., from C8 to C1; Figure 1) to avoid any sampling-­induced bility coefficients of each gas compound (Tassi et al., 2018).
contamination. The main anions and cations were analyzed by ion chromatog-
raphy (761 Compact IC-­Metrohm and 861 Advanced Compact IC-­
Metrohm, respectively). Trace elements (Mn, Fe, Co, Ni, Cu, Zn,
3.2  |  Water and dissolved gas Ba, and Sb) were determined by inductively coupled plasma optical
sampling and analysis emission spectrometry (ICP-­OES) with PerkinElmer Optima 8200.
The analytical errors for IC and ICP-­OES were ≤ 5 and ≤ 10%, re-
At each sampling site, pH, temperature (T, °C), and dissolved oxy- spectively. DOC samples were analyzed by a Shimadzu TOC ana-
gen (DO, mg/L) were measured in situ with a Hach HQ 40d probe. lyzer. Ammonium was spectrophotometrically detected at 420 nm (3
Alkalinity was determined directly in the field via acidimetric titra- drops of Seignette's salt and 0.3 ml of Nessler reagent were added
tion with 0.01N HCl and methyl orange as indicator. Three filtered to 5 ml sample and measured after 3 min). Nitrite was diazotated and
(0.45 μm) water aliquots were collected in polyethylene bottles spectrophotometrically detected at 543 nm (0.1  ml sulfanilamide
at each site, as follows: (i) 125 ml aliquot for the analysis of main and 0.1 ml NEDA reagent were added to 5 ml sample and measured
anions (SO 42−, − − − −
Cl , F , NO3 , and Br ), (ii) 50 ml aliquot acidified after 6 min).
with 0.5  ml of Suprapur HCl for the determination of the main The δ13C-­TDIC values (expressed as ‰ vs. V-­PDB) were analyzed
2+ 2+ + +
cations (Ca , Mg , Na , and K ), and (iii) 50 ml aliquot acidified with a Finnigan Delta Plus XL mass spectrometer in CO2 recovered
with 0.5 ml of Suprapur HNO3 for the analysis of trace elements. after the reaction of 3 ml of water with 2 ml of anhydrous H3PO 4 in
|
4      VENTURI et al.

pre-­evacuated sample holders, which were then left in a thermo- (C1) and last sampling (C8) points where only the superficial layer
static bath at 25 ± 0.1°C for 12 h. The CO2 was extracted and puri- was collected. Core biofilm samples were obtained by drilling the
fied using a two-­step cryogenic (liquid N2 and a mixture of liquid N2 travertine deposit using plastic cylinder with around 1  cm diame-
trichloroethylene; Evans et al., 1988; Vaselli et al., 2006) procedure ter to keep unaltered the biofilm structure and stratification. These
(Salata et al., 2000). The analytical error for δ13C-­TDIC values was core samples were immediately stored at −20°C and successively
±0.05 ‰. used for DNA extraction. Furthermore, around 1  g of superficial
The chemical composition of the main inorganic compounds and sub-­superficial layer was collected by using a plastic spoon and
(CO2, N2, Ar, O2, and He) stored in the headspace of the sampling immediately stored at +4°C until analysis. In the laboratory, these
flasks was analyzed using a Shimadzu 15A gas chromatograph (GC) aliquots were diluted (1:10 w/v) with sterilized buffer solution con-
equipped with a 9-­m-­long molecular sieve column and thermal con- taining formaldehyde in 15-­ml Falcon tubes and further processed
ductivity detector (TCD). The analysis of CH4 was carried out using a using Nycodenz density gradient centrifugation, as described else-
Shimadzu 14A gas chromatograph equipped with a flame ionization where (Amalfitano & Fazi, 2008). Then, the liquid suspensions were
detector (FID) and a 10-­m-­long stainless steel column packed with filtered on polycarbonate membrane filters and used for microscopy
Chromosorb PAW 80/100 mesh coated with 23% SP 1700 (Tassi observation.
et al., 2008; Vaselli et al., 2006). The analytical error for GC analysis Total prokaryotic abundance was estimated by DAPI stain-
was ≤5%. The total amount of dissolved gases was then calculated ing, while those of Bacteria and Archaea were determined by
according to Henry's law. catalyzed reporter deposition–­f luorescence in situ hybridiza-
13 12
The C/ C isotopic ratios of dissolved CO2 (expressed as tion (CARD-­FISH), following the protocol optimized by Fazi
δ13C-­CO2 in ‰ vs. V-­PDB) were determined on the basis of those et al. (2007, 2013) using specific rRNA-­t arget HRP-­labelled probes
measured on the gaseous CO2 stored in the sampling flask head- (Biomers, Ulm, Germany): EUB338 I-­III for Bacteria; ALF968 for
space (δ13C-­CO2_STRIP). The δ13C-­CO2_STRIP values were analyzed Alphaproteobacteria; BET42a for Betaproteobacteria; GAM42a
by a Finnigan Delta Plus XL mass spectrometer after a two-­step for Gammaproteobacteria; DELTA495 for Deltaproteobacteria;
extraction and purification procedure of the gas mixture, as de- CFX and GNSB for Chloroflexi; LGC354 for Firmicutes; CF319a
scribed for the analysis of the δ13C-­TDIC values. Internal (Carrara for Flavobacteria; PLA46 for Planctomycetes; TM7905 for TM7;
and San Vincenzo marbles) and international (NBS18 and NBS19) HGC69A for Actinobacteria; and ARCH915 for Archaea. Details
standards were used for the estimation of external precision. The of probes are available at probeBase (Greuter et al.,  2016). The
analytical error and the reproducibility were ± 0.05 ‰ and ± 0.1 stained filter sections were inspected on a Leica DM LB 30 epi-
‰. The δ13C-­CO2 values were then calculated from the measured fluorescence microscope (Leica Microsystems GmbH, Wetzlar,
δ13C-­CO2_STRIP (Venturi et al.,  2017) based on the enrichment fac- Germany) at 1000× magnification. At least 300 cells were counted
tor (ε1) for gas–­water isotope equilibrium proposed by Zhang in >10 microscopic fields randomly selected across the filter sec-
et al. (1995), as follows: tions. The relative abundance of hybridized cells was estimated as
the ratio of hybridized cells to total DAPI-­s tained cells.
𝛿 13 CO2 = 𝜀1 + 𝛿 13 CO2_STRIP = 0.0049 × T C − 1.31 + 𝛿 13 CO2_STRIP . The abundance of microbial free-­living cells and aggregates
( (◦ ))

was determined using the Flow Cytometer A50-­micro (Apogee


3.3  |  Sampling and analysis for microbial Flow System, Hertfordshire, England) equipped with a solid-­
community characterization of waters and biofilms state laser set at 20 mV and tuned to an excitation wavelength of
488 nm. The volumetric absolute cell counting was carried out on
At each sampling site, 1 L of water was filtered through polycarbon- samples stained with SYBR Green I (1:10,000 dilution; Molecular
ate membranes (pore size 0.2 μm, 47 mm diameter, Nuclepore) and Probes, Invitrogen). Apogee Histogram Software (v89.0) was used
immediately stored at −20°C for DNA extraction. Moreover, 100 ml to plot and analyze data; the light scattering signals (forward and
of water for the microbial community characterization was fixed side scatters) and the green fluorescence (530/30 nm) were con-
in formaldehyde solution (FA, 1% vol/vol final concentration) and sidered for the single-­cell characterization. Thresholding was set
kept at 4°C until analysis (performed within 24 h). Subaliquots (30–­ on the green channel, and voltages were adjusted to place the
80 ml) were filtered on polycarbonate membrane filters (Nuclepore background and instrumental noise below the first decade of
filters: 47 mm diameter with a pore size of 0.2 μm) by gentle vacuum green fluorescence. Samples were run at low flow rates to keep
(<0.2 bar), and the preparations were washed with 20 ml of Milli-­Q the number of events below 1000 events s−1. The intensity of
water. The obtained filters were stored at −20°C until further green fluorescence emitted by SYBR-­p ositive cells allowed for the
processing. discrimination among cell groups exhibiting two different nucleic
Biofilm grown on travertine deposit along the channel bed was acid content (cells with low nucleic acid (LNA) content and cells
collected at the same time as water sampling. In detail, 14 biofilm with high nucleic acid (HNA) content). Fixed gates were designed
samples were collected at the same sampling points of water sam- to discriminate between free-­living cells and aggregates accord-
ples from the superficial (2  cm depth) and sub-­superficial layer of ing to their signatures in a side scatter vs. green fluorescence plot
travertine deposit (8 cm depth), with the sole exception of the first (Amalfitano et al., 2018).
VENTURI et al. |
      5

Approximately 1  g of biofilm sample was used for DNA ex- overcome low complexity issue often observed with amplicon
traction with PowerSoil® DNA Isolation Kit (MoBio, Carlsbad, CA) samples.
by following the manufacturer's instructions. For water samples, Forward reads were trimmed for quality using Trimmomatic v.
DNA was extracted utilizing one entire polycarbonate membrane 0.32 (Bolger et al.,  2014) with the settings SLIDINGWINDOW:5:3
for each sample. The quality of extracted DNA (1.6<A260/280<1.8 and MINLEN:275. The dereplicated reads cut to 275 bp and clus-
and A260/230>2) was analyzed with a Nanodrop 3300 (Thermo tered using the usearch v. 7.0.1090 -­cluster_otus command with
Scientific, Italy). DNA was stored at −20°C in small aliquots. default settings. OTU abundance was estimated using the usearch
Bacterial and archaeal V3-­4 16S sequencing libraries were v. 7.0.1090 -­usearch_global command with -­id 0.97. Taxonomy was
prepared by a custom protocol based on an Illumina protocol assigned using the RDP classifier (Wang et al., 2007) as implemented
(Illumina, 2015). Up to 10 ng of extracted DNA was used as template in the parallel_assign_taxonomy_rdp.py script in QIIME (Caporaso
for PCR amplification of the 16S gene fragments. Each PCR (25 μl) et al., 2010), using the MiDAS database v.1.20 (McIlroy et al., 2015).
contained dNTPs (100 μM of each), MgSO 4 (1.5 mM), Platinum® The results were analyzed in R through the RStudio IDE using the
Taq DNA polymerase HF (1  U), 1X Platinum® High Fidelity buffer ampvis package v.1.27.0 (Albertsen et al., 2015).
(Thermo Fisher Scientific, USA), and tailed primermix (400 nM of
each forward and reverse). PCR was run according to the following
program: initial denaturation at 95°C for 2 min, 35 cycles of amplifi- 3.4  |  Tridimensional successional changes in biofilm
cation (95°C for 20 s, 50°C for 30 s, 72°C for 60 s), and a final elon-
gation at 72°C for 5 min. Duplicate PCRs were performed for each Clean microscopy slides were placed underwater in the central point
sample, and the duplicates were pooled after PCR. The forward and of channel (corresponding to sampling site C4; Figure 1c) with a pol-
reverse-­t ailed primers were designed according to Illumina  (2015) yvinyl chloride (PVC) support and collected overtime (2–­7–­12 days)
and contain primers targeting bacteria and archaea 16S gene V3-­4 to monitor biofilm development, biomass increments in the earlier
region (Sundberg et al.,  2013): 5′-­CCTAYGGGRBGCASCAG (341F) stages, and changes in the tridimensional microstructure during bio-
and 5′-­GGACTACNNGGGTATCTAAT (806R). The primer tails enable film maturation. The tridimensional structure, successional changes,
attachment of Illumina Nextera adaptors for sequencing in a sub- and microbial colonization were assessed using the CARD-­FISH
sequent PCR. The resulting amplicon libraries were purified using technique in combination with confocal laser scanning microscopy,
the standard protocol for Agencourt AMPure XP Bead (Beckman according to the protocol of Lupini et al. (2011). Image elaborations
Coulter, USA) with a modified bead-­to-­sample ratio of 4:5. The DNA were performed using the Imaris 6.2 software (Bitplane AG, Zurich,
was eluted in 33 μl of nuclease-­free water (Qiagen, Germany). DNA Switzerland).
concentration was measured using a Qubit™ HS DNA Assay kit
(Thermo Fisher Scientific, USA).
Sequencing libraries were prepared from the purified amplicon 3.5  |  Travertine analysis
libraries using a second PCR. Each PCR (25 μl) contained 1x PCRBIO
HiFi buffer (PCR Biosystems, UK), PCRBIO HiFi Polymerase (1  U) Travertine encrustations were collected immediately downstream
(PCR Biosystems, UK), adaptor mix (400 nm of each forward and re- with respect to each biofilm sampling point at the interface between
verse), and up to 10 ng of amplicon library template. PCR was run the channel bottom and water. Travertine samples were analyzed by
with the following program: initial denaturation at 95°C for 2  min, scanning electron microscopy (SEM, coupled to energy-­dispersive
8 cycles of amplification (95°C for 20 s, 55°C for 30 s, and 72°C for microanalysis, EDS) and X-­ray powder diffraction (XRPD). After
60 s), and a final elongation at 72°C for 5 min. The resulting sequenc- sampling, the eight collected encrustations (C1 to C8) were manu-
ing libraries were purified using the standard protocol for Agencourt ally separated at the binocular microscope according to their color
AMPure XP Bead (Beckman Coulter, USA) with a modified bead-­to-­ (white, red, and green, labelled W, R, and G, respectively). When
sample ratio of 4:5. The DNA was eluted in 33 μl of nuclease-­free possible, the original texture was preserved. The resulting travertine
water (Qiagen, Germany). The DNA concentration was measured samples were 15 (Table S1). In addition, two samples were also col-
using a Qubit™ HS DNA Assay kit (Thermo Fisher Scientific, USA). lected from the artificial deposits at different times (14 and 21 days)
Gel electrophoresis using TapeStation 2200 and D1000 ScreenTapes by inserting the slides underwater in the central point of the channel.
(Agilent, USA) was used to check the product size and purity of a The travertine samples were gently fixed over the stubs for SEM
subset of sequencing libraries. analysis, using a double-­sided conductive carbon tape, coated with
The purified sequencing libraries were pooled in equimolar a graphite layer to ensure their electrical conductivity, and ana-
concentrations and diluted to 2 nM. The samples were paired-­end-­ lyzed with a SEM ZEISS EVO MA15 (at MEMA—­Centro di Servizi
sequenced (2 × 301 bp) on a MiSeq (Illumina) using a MiSeq Reagent di Microscopia Elettronica e Microanalisi, University of Florence),
kit v3, 600 cycles (Illumina, USA) following the standard guidelines equipped with the Oxford INCA 250 Microanalysis. Backscattered
for preparing and loading samples on the MiSeq, as described in and secondary electron micrographs were registered, while the min-
Caporaso et al.  (2012). 20% PhiX control library was spiked in to eral identification was carried out by means of point and raster X-­ray
|
6      VENTURI et al.

EDS microanalysis. Measurements were carried out at an accelerat- injects a constant flow rate of 0.035 kg s−1 of H2O at 2.7867·105 J kg−1
ing voltage of 20 KV. enthalpy, and 1.52·10−5 kg s−1 CO2 at 1.528·106 J kg−1 enthalpy, cor-
The materials considered for the X-­ray powder diffraction responding to the flow rate measured at PC at 52.7°C. The chemical
(XRPD) were carried out after gently crushing each sample in an composition of the injected water was that analyzed in C1 (see pre-
agate mortar and analyzed with a XRD Bruker New D8 Da Vinci pow- vious paragraph) and was used to model the chemical evolution of
der diffractometer (at CRIST—­Centro di Cristallografia Strutturale, the channel. The atmospheric boundary was set as infinite volume
University of Florence), employing a Ni-­filtered Cu Kα (1.54187 Å) elements at 1.103 × 105 Pa pressure, 16.4°C temperature, and 40 Pa
radiation. XRPD patterns were registered at 1600 W (i.e., 40 kV, CO2 partial pressure (corresponding to 400 ppm of CO2). The mea-
40 mA) with a fast multichannel detector in the 2θ range 10–­90°, sured temperature and CO2 profiles along the channel were used
applying a step size of 0.0205° 2θ. The XRPD data were refined by to validate the model. The only calibrated variable was the reactive
means of full-­profile Rietveld algorithm, using the Fullprof software surface of the main mineral phase of the system, that is, calcite.
(Rodriguez-­C arvajal,  1993). The red and green travertine samples
obtained from C4 (i.e., C4R and C4G, respectively; Table  S1) were
investigated together with a synthetic commercial calcite (Rudi Pont, 4  |  R E S U LT S
Turin, Italy), the latter being used to calibrate the diffractometer ge-
ometry and line shape. 4.1  |  Physicochemical characteristics of waters and
dissolved gases

3.6  |  Computational model A strong decrease in water temperature was observed along the
thermal water outflow channel, passing from 52.4 (C1) to 16.4
Numerical modelling of the geochemical system (surface and bot- (C8)°C. Differently, an increase in DO values was observed along
tom fluids) was performed by means of the TOUGHREACT v (Xu the channel, with values from 1.84 to 9.08 mg/L (Table 1). Similarly,
et al.,  2006; Xu & Pruess,  2001) software, using the default data- pH increased from C1 to C4 (from 7.25 to 8.40), and then, it ranged
base included in the package (thermoxu.dat; Xu & Pruess, 2001; Xu between 8.63 in C5 and 8.61 in C8 (Table 1).
et al., 2006). The model uses the equation of state EOS2 for water The sampled waters were characterized by a Ca-­SO 4 (HCO3)
and CO2 (Pruess et al., 2012). The geochemical model strategy pro- composition, with TDS (total dissolved solid) values progressively
ceeds through a fully kinetic approach. With the kinetic database decreasing along the channel (from 3.2 to 2.8 g/L), mainly due to
from Palandri and Kharaka (2004), the mineral-­specific surface was a decrease in Ca2+ and HCO3− (Table  1). Similarly, an overall de-
calibrated against the measured data. In order to obtain an overall creasing trend was also observed in the concentrations of trace
estimate of the grain size of a monodisperse distribution of particles species, such as Fe, Mn and Ba, from C1 (40, 35, and 46 μg/L, re-
to be associated with the best-­fit specific surface, the relationship spectively) to C7 (6.9, 20, and 25 μg/L, respectively), whereas they
between radius and surface area was computed according to the slightly increased in C8 (Table  1). Similarly, NH4 + concentration
cubic array of truncated spheres model that makes up the framework also progressively decreased along the channel, whereas that of
of the rock used in TOUGHREACT (Sonnenthal & Ortoleva, 1994). NO2− showed a regular, though increasing, trend, mimicking that
This model at its limit (with 0 grain–­grain contact areas) reduces to a of TOC (Table 1).
simple spherical model of mineral grains. Among dissolved gases, a sharp decrease along the channel was
The relative permeability and capillary pressure equation was recorded for dissolved CO2, that is, from 9.86 to 4.46 mmol/L, as well
obtained according to Corey (1954), with an irreducible liquid satu- as CH4, from 0.089 to 0.048 mmol/L (Table 1). Differently, O2 con-
ration assumed at 0.2. Diffusivity coefficients of CO2 in water along centrations increased by one order of magnitude along the channel,
the vertical section of the channel and from the top of the channel that is, from 0.005 to 0.068 mmol/L (Table 1).
into atmosphere were computed, as a function of temperature, ac- The carbon isotopic composition of TDIC and dissolved CO2
cording to Cadogan et al. (2014). increased along the channel, ranging from +1.88 and −3.86 ‰ vs.
The channel was modelled as a material with a porosity of 0.9999 V-­PDB in C1, respectively, to +4.90 and −0.69 ‰ vs. V-­PDB in C3,
and permeability of 1·10−10 m2, with specific heat and heat conduc- respectively, and then, they were clustering around +5.26 and −0.05
tivity of pure water (1 kJ kg−1 K−1 and 0.6 W m−1 K−1, respectively). A ‰ vs. V-­PDB from C4 to C8, respectively (Table 1).
two-­dimensional x-­z model, whose dimensions were 110 m long and
10  cm high (plus one more cell for the atmosphere boundary), de-
scribed the channel. This system was modelled with an x,z grid made 4.2  |  Physicochemical patterns and mineral
of 100 × 21 computational elements. At the end side of the channel precipitation modelling along the outflow channel
(discharge pool), a column of infinite volume cells provided the right
boundary condition (i.e., at constant conditions: T = 16.4°C; P nearly A numerical model was defined by assuming a purely abiotic evo-
1  bar accounting for the hydrostatic pressure gradient within the lution of the water parameters provided by the spring supply.
channel). At the left side of the model (inlet pool), a column (20 cells) The validation of the thermal and flow rate model was carried
VENTURI et al. |
      7

TA B L E 1  Physicochemical and isotopic (δ13C-­TDIC and δ13C-­CO2) features of water and dissolved gas samples collected along the thermal
water outflow channel at the Piscine Carletti

Sample C1 C2 C3 C4 C5 C6 C7 C8

Distance (m) 0 14 28 42 56 70 84 100


T (°C) 52.4 42.6 39.1 30.8 25.7 23.5 20 16.4
DO (mg/L) 1.84 5.84 6.39 8.02 8.22 8.42 8.41 9.08
pH 7.25 7.76 8.16 8.4 8.63 8.62 8.69 8.61
TDS (g/L) 3.20 3.24 2.96 2.92 3.01 2.96 2.85 2.79
HCO3 (mg/L) 1180 1243 987 965 1007 982 885 854
F (mg/L) 2.60 2.40 2.02 2.73 1.96 1.98 1.73 1.88
Cl (mg/L) 17.7 21.6 21.5 23.3 21.7 21.7 22.9 22.2
Br (mg/L) 0.023 0.082 0.033 0.03 0.066 na na 0.104
NO3 (mg/L) 0.065 0.102 na na 0.074 0.057 0.490 0.340
NO2 (mg/L) 0.010 0.007 0.020 0.030 0.003 0.026 0.026 bdl
NH4 (mg/L) 1.417 1.159 0.863 0.708 0.799 0.631 0.657 0.631
SO 4 (mg/L) 1153 1175 1182 1180 1196 1187 1199 1186
Li (mg/L) 0.025 0.021 0.021 0.015 0.022 0.015 0.020 0.018
Na (mg/L) 38.8 35.9 37.6 37.5 38.9 40.1 37.5 35.1
K (mg/L) 41.5 38.9 39.5 39.9 41.0 38.5 41.1 36.2
Mg (mg/L) 136 131 133 130 135 135 132 128
Ca (mg/L) 627 593 561 541 569 551 533 525
Mn (μg/L) 35 33 27 23 23 28 20 22
Fe (μg/L) 40 21 19 7.1 6.8 11 6.9 33
Co (μg/L) 0.3 bdl 0.2 0.2 bdl bdl bdl bdl
Ni (μg/L) 3.4 3.1 3.2 2.8 2.7 3.4 1.7 2.6
Cu (μg/L) 5.4 5.4 4.9 5.0 5.0 6.6 5.1 5.1
Zn (μg/L) 6.1 6.2 3.3 4.7 5.4 14 3.1 5.7
Ba (μg/L) 46 44 35 26 28 26 25 44
Sb (μg/L) 7.5 6.4 6.7 2.8 2.7 3.4 bdl 5.7
As (μg/L) 176 215 209 219 202 193 162 187
TOC (mg/L) 0.23 0.34 0.35 0.70 0.44 0.22 0.47 0.53
δ13C-­TDIC (‰ vs. V-­PDB) 1.88 3.76 4.90 5.07 5.00 5.56 5.46 5.22
CO2 (mmol/L) 9.86 9.47 9.16 8.06 6.58 5.47 5.45 4.46
N2 (mmol/L) 0.51 0.56 0.55 0.52 0.51 0.55 0.58 0.61
CH4 (mmol/L) 0.089 0.079 0.081 0.075 0.069 0.056 0.051 0.048
Ar (mmol/L) 0.015 0.018 0.020 0.013 0.013 0.014 0.014 0.015
O2 (mmol/L) 0.005 0.005 0.009 0.009 0.011 0.031 0.051 0.068
He (mmol/L) 0.000028 0.000025 0.000022 0.000021 0.000025 0.000018 0.000016 0.000016
13
δ C-­CO2 (‰ vs. V-­PDB) −3.86 −2.80 −0.69 −0.03 na −0.23 0.23 −0.15

Abbreviations: bdl, below the detection limit; na, not analyzed.

out by comparing the theoretically predicted temperature and The dissolved CO2 amount was regulated by both exsolution
the dissolved CO2 parameters with those effectively measured in toward the atmosphere boundary and vertical diffusion of CO2 in
the field. The data reported in Figure  2a evidenced a very good water. Similar to the measured and computed temperatures, the
agreement between experimental and computed temperatures at model accurately reproduced all experimental evidences (Figure 2b).
increasing distances from the spring. Since flow rate affects the Once validated, the model was used to predict theoretical diffu-
thermal exchange with atmosphere, thus governing the cooling of sion and transport properties of chemical species involved in both
the channel, the calculated thermal profile allowed the correct pre- solution equilibria and precipitation of solid phases. The behavior of
diction of the temperature, evaluated at 1 cm from the bottom of HCO3− and Ca2+ was modelled by a kinetically controlled calcite pre-
the channel. cipitation (Figure 3a–­c). The kinetics of this process was accounted
|
8      VENTURI et al.

F I G U R E 2  Results of the numerical model of the channel: (a) temperature and (b) dissolved CO2. In the inset, the comparison of the
considered parameter, evaluated at 1 cm from the bottom of the channel (blue line), with the corresponding experimental measurement
(points), listed in Table 1.

for through a reaction surface of 1.105 × 106 m2/m3, which corre- gypsum that were calculated along the whole channel (Figure  3e).
sponded to a grain size of ~30 μm (Sonnenthal & Ortoleva,  1994). According to our model, gypsum precipitation did not occur due to
This size represented a maximum value, and it could be reduced the low saturation index (Figures 3e and S2).
to some micrometres taking into account additional factors such
as packing, biofilm occlusion, and active sites on the surface (e.g.,
Murphy et al., 1989; Steefel & Maher, 2009). The computed calcite 4.3  |  Prokaryotic abundance and community
grain size was then compared with that measured by SEM. Calcite in structure in water
each travertine sample was found in different aggregates. However,
individual grains were rarely exceeding 20 μm. In contrast, high sur- Total prokaryotic abundance in water samples increased from
face grain aggregates of definitely lower dimensions were relatively 6.8 × 105 to 8.2 × 105 cells/ml along the channel (Table S2). HNA cells
frequent (see section 4.6). Accordingly, a good match between the were dominant and slightly decreased along the channel, from C1
grain size evaluated by the numerical model and that observed in the (87.8% of total cells) to C8 (71.9% of total cells). Moreover, along with
collected samples was assessed. a high cytometric similarity among water samples, cytograms clearly
Concerning other ions, that is, F− and Cl− and cations, such as Li+, showed the occurrence of microbial aggregates, with density values
Mg2+, K+, and Na+, the experimental results and those provided by decreasing from C1 (3.4 × 10 4 aggregates/ml) to C8 (1.8 × 10 4 aggre-
the model point to a nearly constant behavior (Figure S1). This can gates/ml; Figure S3).
be explained with the fact that (i) their salts are highly soluble and (ii) Accordingly, CARD-­FISH analysis revealed a similar micro-
no reactions affect these ions. bial community composition in water column along the chan-
The model predicted that sulfate concentration is moderately nel (Figure  4a). The main microbial component in water samples
variable downstream and also from the water/air interface down to belonged to bacteria domain (on average 79.3% of total DAPI-­
the bottom of the channel (Figure  3d). However, the total amount stained cells) mainly affiliated with Proteobacteria (Figure  4a).
of dissolved sulfate was not consistently changing. The difference Gammaproteobacteria was the predominant group along the whole
between the measured and computed sulfate concentration is likely channel showing abundance between 1.2 × 105 ± 5.1 × 103 and
2+ 2+
due to the formation of soluble neutral complexes with Ca , Mg 5.3 × 105 ± 2.0 × 10 4 cells/ml (range: 16.7%–­77.1% of DAPI-­s tained
and Na+ cations (Dai et al., 2017; Krumgalz, 2018; Pillay et al., 2005), cells). On average, Alpha-­, Beta-­, and Deltaproteobacteria repre-
as evidenced by the model. It is however matter of fact that the mea- sented 7% of total prokaryotic abundance. The other bacterial
sured total amount of sulfate in the liquid phase maintained con- groups represented less than 2.2% of total cells. Meanwhile, 15.4%
stant. This finding is also in line with the saturation index values for of total prokaryotic cells (on average) belonged to Archaea, with
VENTURI et al. |
      9

F I G U R E 3  Results of the numerical model of the channel: (a) Ca2+ and (b) HCO3−. In the inset, the comparison of the considered
parameter, evaluated at 1 cm from the bottom of the channel (blue line), with the corresponding experimental measurement (points), listed in
Table 1. (c) Volume fraction of precipitated calcite. (d) Results of the numerical model of the channel for SO 42−. (e) Saturation index evaluated
for the most relevant mineral species accounted in the model.

a cell abundance ranging between 4.1 × 10 4 ± 1.2 × 103 cells/ml and represented by Gammaproteobacteria affiliated with genus
5 3
1.9 × 10  ± 6.4 × 10  cells/ml. Thiofaba (~85% of total OTUs) (Figure 4b). Epsilonproteobacteria
The outputs from high-­t hroughput sequencing were in represented around 9.2% of total reads. Archaea represented on
line with the results generated by CARD-­F ISH analysis. In average less than 1% of total reads. A low level of biodiversity
particular, a total of 57,943 reads were generated by C1, C4, was found in water samples with Shannon index values rang-
and C7 water samples. These reads resolved into 239 OTUs. ing between 0.6 and 0.8 and very similar Simpson index values
Overall, Proteobacteria was the most abundant phylum, mainly (range: 0.26–­0 .30).
|
10      VENTURI et al.

F I G U R E 4  Water: (a) abundance of phylogenetic taxa (Archaea and main phyla within Bacteria) and single classes within Proteobacteria (in
blue) in the different water sampling points. (b) OTU relative abundance in water samples estimated by NGS. Bacterial taxa accounting for
less than 1% of total composition were classified as “other Bacteria”.

4.4  |  Prokaryotic abundance and community Sphingomonadales. Furthermore, a high abundance of OTUs affili-
structure in biofilm ated with Rhodomicrobium, a genus belonging to order Rhizobiales
within Alphaproteobacteria, was mainly observed in C1 site (14.8%).
Along the channel, total prokaryotic abundance tended to de- OTUs affiliated with Gammaproteobacteria represented on average
crease in superficial biofilm layers, whereas it increased in sub-­ 3.4% and 1.7% of total OTUs in superficial and sub-­superficial layers,
superficial samples (Figure  5a). Values decreasing from 3.8 × 109 respectively.
cells/g to 4.0 × 107 cells/g were observed in superficial layers from The phylum Chloroflexi represented 8.8% (on average) of total
C2 to C8 samples. In contrast, increasing values from 5.2 × 107 to OTUs in superficial biofilms and up to 43.7% in sub-­superficial lay-
8.8 × 107 cells/g were observed in sub-­superficial layers. ers. These OTUs were mainly affiliated with family Anaerolineaceae
The microbial communities were dominated by Bacteria in in the superficial layer and genus Roseiflexus in the sub-­superficial
both the superficial and sub-­superficial layers representing the one. Members of this phylum were highly abundant, mainly in
90% (on average) of total prokaryotes estimated by CARD-­FISH. sub-­superficial biofilm in C7 site. OTUs affiliated with phylum
Among bacterial cells, Cyanobacteria were abundant in the su- Bacteroidetes, mostly belonging to family Saprospiraceae, increased
perficial layers showing values from 2.7 × 109 ± 3.0 × 107 cells/g along the channel up to 21.6% of total OTUs in site C8. The retrieved
in C2 to 1.9 × 107 ± 3.5 × 106 cells/g in C8 (Figure  5b). Lower cy- OTUs affiliated with phylum Chlorobi represented 2.6% (on average)
anobacterial abundance was observed in the sub-­superficial of total OTUs in both the superficial and sub-­superficial layers, with
7 6
layers with an average value of 1.2 × 10  ± 7.1 × 10  cells/g. The the sole exception of the C2 superficial layer in which the members
various classes within Proteobacteria counted together on average of Chlorobiaceae represented up to 12.1% of total OTUs.
1.3 × 10 8 ± 1.4 × 10 8 cells/g and 3.1 × 107 ± 2.1 × 107 cells/g in the su- Concerning the Archaea, Thaumarchaeota represented between
perficial and sub-­superficial layers, respectively. 14.9% and 41.0% of total reads in C1 and C4 sub-­superficial layer,
High-­throughput sequencing generated a total of 142,563 reads respectively. In the other biofilm samples, the OTUs belonging to
in the C1, C4, C7, and C8 superficial and sub-­superficial biofilm lay- archaea domain represented less than 0.5% of total reads.
ers that resolved into 949 OTUs. The results showed a high microbial
diversity among biofilm samples. The Shannon and Simpson indexes
ranged between 2.4 and 2.7 and between 0.8 and 0.9, respectively. 4.5  |  Tridimensional structure and successional
OTUs affiliated with Cyanobacteria were more abundant in the changes of biofilm
superficial layers (on average 37.3%) than in the sub-­superficial
ones (13.0%, on average) (Figure 5c). These OTUs mainly belonged Development, biomass increments, and three-­dimensional struc-
to genera Spirulina (around 35% in C4 and C7 superficial biofilm), ture of biofilm growing on microscopy slides placed underwater on
Leptolyngbya (on average 8.4% in superficial layer and 1.8% in sub-­ the central point of the channel (corresponding to sampling point
superficial one), and Fischerella (up to 7.7% in C4 sub-­superficial C4) were observed by combining CARD-­FISH and CLSM. After 2
biofilm). Overall, members of Proteobacteria represented on av- and 7 days, CLSM examination revealed biofilm assemblages with
erage around 20% of total reads in biofilm samples. In particular, similar microbial community equally represented by filamentous
OTUs affiliated with Alphaproteobacteria were the most abundant Cyanobacteria and other prokaryotes (Figure  6). After 12 days, a
in both superficial and sub-­superficial biofilms. Specifically, they highly complex and multistratified biofilm was observed, with a high
mainly belonged to orders Rhodobacterales, Rhodospirillales, and amount of filamentous spirulina-­like Cyanobacteria, dominating the
VENTURI et al. |
      11

F I G U R E 5  Biofilm: (a) Total prokaryotic abundance in superficial and sub-­superficial biofilm samples. (b) Total abundance of Cyanobacteria
and single classes within Proteobacteria in superficial and sub-­superficial biofilm samples. (c) OTU relative abundance in biofilm samples
estimated by NGS. Clusters making up less than 1% of total composition were classified as “other Bacteria” or “other Archaea”.

microbial community. Within the dense network of filamentous au- The second, and largely more frequent, calcite facies was a shrub
totrophs, non-­pigmented prokaryotic cells were also visible. precipitate in which the evolution of the crystal(s) occurred through
the surface uptake by rhombohedral lamellae (Figure  7b). In many
cases, calcite individual grains were no single crystals but aggregates
4.6  |  Mineralogical features of many individuals. Moreover, the overall surface of the grains ap-
peared highly structured and rough (Figure  7c,d). The individual
Three mineralogical phases were recognized throughout the chan- grains were significantly bigger than those of euhedral calcite. All
nel. Calcite was found in two apparently different and coexisting samples but C1 exhibited this kind of mineralization, irrespectively
morphologies. The first one exhibited a rhombohedral or, more to the sample type (green, red, or white).
rarely, scalenohedral habitus (Figure  7a). The crystals were almost Both calcite precipitates exhibited consumption from either
homogeneous with an average size of about 5  μm and displayed a the crystal edges or circular pits on the crystal surfaces (Figure 7).
euhedral aspect. Such calcite was significantly abundant in the C1 We attributed these features to the effect of crystal weathering,
precipitate although it was recorded throughout the whole channel. representing fingerprint of partial redissolution of the precipitated
|
12      VENTURI et al.

abundant biofilm occurrences. Crystal surfaces were well formed,


displaying neither consumption nor pitting evidences.
Finally, the third identified mineral phase was fluorite, with the
typical cubic habitus of 5–­10 μm in size (Figure 7d). The seldom fluo-
rite crystals identified were well formed, perfectly euhedral, without
evidence of active redissolution processes, and apparently unrelated
to the mineralogical context where they were recognized, and some-
times clustered in small regions of the samples.
All samples but C1 displayed extensive evidence of biofilm. This
provided a relevant network among all types of precipitated miner-
als, as it can be devised by comparing secondary and backscattered
electron micrographs (Figure 8a–­c). In the white-­colored encrusta-
tions (i.e., in C7W and C8W), biofilms were less extended, but still
present (Figure 8d). Regarding the relationships with the described
minerals (euhedral and shrub calcite, gypsum, fluorite), no specific
relationship occurred between biofilm and euhedral calcite, gypsum,
and fluorite. Their spatial co-­localization may result just from pre-
cipitation processes occurring almost simultaneously or in sequence
in the same volume of the biofilm/solution system. Conversely, the
mutual relationships between biofilm and shrub calcite were com-
pletely different and more complex (Figures  7b and 8e): in certain
cases, the grains were completely enveloped by biofilm, whereas,
more frequently, they were interconnected by filaments of organic
matter. Overall, the mutual relationships between biofilm and shrub
calcite appeared to be not occasional.
The main micromorphological difference in the glass slides po-
sitioned underwater on the channel (Figure  9a,b) was the net in-
crease in biofilm occurrence. In the 14-­day slide, biofilm was scarce,
whereas in the 21-­day slide, it appeared abundant. Concerning cal-
cite precipitate morphology, at intermediate time both the rhombo-
F I G U R E 6  CLSM combined images showing the spatial hedral and the shrub facies were concurring, with poor evidence of
distribution (X−Y, X−Z, and Y−Z planes) of Bacteria (green), crystal weathering (Figure  9a). Weathering and the occurrence of
Cyanobacteria (red) and other prokaryotes (blue) identified by topological relationships between calcite crystals and biofilm (al-
CARD-­FISH in biofilms. The hybridized bacterial cells were excited ready observed by confocal microscopy) were the main features of
with the 488 nm line of an Ar laser (excitation) and observed in
the calcite precipitates (almost exclusively shrub) in the 21-­day slide
the green channel from 500 to 530 nm (emission). Calcite crystals
were visualized by their reflection signal (405 nm line of a diode (Figure 9b). Interestingly, in the 14-­day slide also spherules of arago-
laser) and appear of gray color. nite needles, already described by Allen et al. (2000) for the nearby
Zitelle spring system, were identified (Figure 9a).
XRPD analysis on differently colored travertine samples (C4R
and C4G) showed a simple mineralogy, being largely dominated
crystals. These evidences were recorded in all the samples, although by calcite (Figure S4). Weak additional reflections in the 25–­35 2θ
their presence was reduced at >40°C (C1 and C2). range can be attributed to quartz and other carbonates (e.g., dolo-
Occasional, though not infrequent, findings of gypsum precip- mite). A refinement of the lattice constants, operated by means of
itates were also observed. All precipitates showed a unique com- the Rietveld method, provided the results shown in Table  S3. The
mon morphology, that is, aggregates of relatively small crystals, with calcite lattice parameters, the relative c/a ratio, and the cell volume
apparent traces of the {010} cleavage, assembled through multiple were rather similar in the two travertine samples. Conversely, when
branching scheme (Van Driessche et al., 2019), typical of fractal crys- compared to analogous parameters described for a pure inorganic
tallization (Van Driessche et al., 2019; Figure 7e). Moreover, these ag- calcite (Graf, 1961), a net lattice strain coupled to a slight increase
gregates grew up to relevant dimensions (up to 50 μm) and appeared in the cell volume was noticed. Accordingly, the red and green trav-
randomly but unevenly distributed in the precipitate (Figure 7f). In ertine encrustations resulted very similar in terms of mineralogical
particular, gypsum precipitates were found always associated with composition and encrustation micromorphology.
VENTURI et al. |
      13

(a) (b)

(c) (d)

(e) (f)

F I G U R E 7  Secondary (a,b) and backscattered (c-­f ) electron micrographs of representative regions of the investigated samples: (a) C1W
(magnification: 9580X): euhedral calcite crystals; (b) C3G (3410X): shrub aggregate of calcite lamellae connected with EPS filaments; (c) C2R
(1500X): calcite aggregates emerging from the biofilm; (d) C5G (1500X): rare euhedral crystals and shrub aggregates of calcite immersed
in the biofilm; redissolution pits are diffuse, as marked, for example, in the red circle; (e) C3R (2760X): aggregate of gypsum crystals;
(f) C2R (500X): aggregates of calcite lamellae and of gypsum crystals (marked in red circles) in biofilm.

5  |   D I S C U S S I O N coupled with an overall increase in NO3− contents (Table 1). On the


contrary, the rapid exsolution of CO2 toward the atmosphere drove
5.1  |  The geochemical context for travertine calcite precipitation along the outflow channel, resulting in a pro-
deposition along the thermal water outflow channel gressive decrease in Ca2+ and HCO3− (Table 1), as also supported by
the numerical model (Figure 3a–­c). Accordingly, CO2 degassing was
The Ca-­SO 4(HCO3) facies is typical of thermal waters circulating in expected to be the main driving factor of travertine precipitation, as
the Vicano–­Cimino Volcanic District, affected by water–­rock inter- observed in similar systems (e.g., Dupraz et al.,  2009). Coherently,
actions with the Mesozoic carbonate sequence and Triassic anhy- the carbon isotopic composition of both dissolved CO2 and TDIC
drites (e.g., Cinti et al.,  2014; Minissale,  2004), in agreement with showed a progressive increase along the outflow channel (Table 1),
previous results on this spring system (Di Benedetto et al.,  2011). being largely controlled by the fast CO2 exchange with the atmos-
Despite the limited spatial scale, the geochemical features evolved phere, as reported by Di Benedetto et al. (2011). The authors also
along the outflow channel with a sharp temperature gradient associ- evidenced the occurrence of an out-­of-­equilibrium isotopic frac-
ated with chemical changes. The water–­atmosphere exchange led tionation process between travertine and dissolved carbon species,
to a progressive increase in dissolved O2 concentrations as the dis- attributed to either kinetically controlled inorganic calcite precipi-
tance from the water spring increased, favoring the oxidation of re- tation or biomineralization processes. On the contrary, the SO 42−
+
duced species as indicated by the decrease in NH4 concentrations concentrations showed no significant variations along the channel
|
14      VENTURI et al.

(a) (b)

(c) (d)

(e)

F I G U R E 8  Secondary (a,b-­e) and backscattered (b) electron micrographs of representative regions of the investigated samples: (a) C3G
(magnification: 1580X) and (b) C3G (1580X): calcite aggregates in biofilm; (c) C8G (1610X) calcite aggregates dispersed in biofilm; (d) C8W
(851X): calcite aggregates with minor biofilm; (e) C7G (3000X): calcite aggregates immersed in biofilm, compared with a single euhedral
calcite crystal (marked by the red arrow).

(Table 1), in agreement with the thermodynamic model (Figure 3e), particles detected in a previous study (Casentini et al.,  2016). This
which predicted that gypsum precipitation did not occur along the may suggest the occurrence of specific particle-­associated microbial
channel. processes in the water flowing along the channel.
Overall, hot springs are characterized by a low biodiversity due
to their extreme conditions in terms of temperature and chemical
5.2  |  Microbial assemblages in waters and biofilms characteristics (Chiriac et al.,  2017; Kemp & Aller,  2004). In this
study, low levels of biodiversity were observed in the water col-
Microbiological characteristics in water samples were shaped by the umns probably due to the harsh environmental conditions more
physicochemical patterns found along the thermal water outflow suited to pioneer/resistant species (Giampaoli et al., 2013; Piscopo
channel. HNA cells, which are often considered as the most active et al., 2006; Valeriani et al., 2018). Accordingly, the microbial com-
fraction (Lebaron et al., 2001), showed a relatively higher abundance munity in the water column analyzed in this study was dominated
in the initial section of the outflow channel (Table S2), suggesting a by Thiofaba genus (Gammaproteobacteria class; Figure  4b), able to
higher microbial activity with respect to the final section. The de- grow under strict aerobic conditions with a chemolithoautotrophic
crease in HNA cells along the channel (Table S2) was likely related metabolism. Members of this genus are known to oxidize sulfur
to the transition from an anaerobic/anoxic high-­temperature aquatic compounds by utilizing H2S as electron donor for CO2 reduction
environment to aerobic and low temperature conditions. Moreover, (van Gemerden, 1993) and are widely reported in hot springs world-
the microbial aggregates showed a cytometric fingerprinting wide (e.g., Gulecal-­Pektas & Temel,  2017; Gumerov et al.,  2011;
and density level (Table  S2) comparable to those of the bioactive Huang et al.,  2011; Mori & Suzuki,  2008; Valeriani et al.,  2018).
VENTURI et al. |
      15

(a) Konhauser, 2007; Liu et al., 2011; Pagaling et al., 2012). Specifically,


micro-­organisms in the sub-­superficial layer were mainly repre-
sented by green and purple sulfur/non-­sulfur bacteria, such as
members of Chlorobi, Chloroflexi and Alphaproteobacteria (Figure 5c).
The family Chlorobiaceae is able to oxidize elemental sulfur and
sulfide and carry out photosynthesis only under anoxic conditions
(Asao & Madigan,  2010; Imhoff,  2014; Madigan et al.,  2017). The
occurrence of Chloroflexi, affiliated with family Anaerolineaceae or
genus Roseiflexus, currently known as “Filamentous Anoxygenic
Phototrophs,” was previously reported in thermophilic cyanobacte-
rial mats (Gaisin et al., 2016; Kambura et al., 2016; Tank et al., 2017;
Valeriani et al., 2018). Furthermore, members of Alphaproteobacteria,
mainly affiliated with Rhodomicrobium genus, preferably grow pho-
toheterotrophically under anoxic conditions in the light, but they
(b) can also use hydrogen, sulfide, or ferrous iron as electrons sources
(Imhoff, 2005).

5.3  |  Evidences from mineralogical


analyses and theoretical computations

Mineralogical investigations assessed the widespread presence of


calcite throughout the outflow channel, as also evidenced by the
numerical model for which calcite precipitation was driven by CO2
degassing from the supersaturated water. Nevertheless, calcite was
detected in two distinct morphologies (Figure 7), indicating the oc-
currence of different precipitation processes. Euhedral calcite can
be ascribed to a purely inorganic precipitation process, as expected
from the theoretical computations, whereas shrub precipitates are
F I G U R E 9  Backscattered electron micrographs of two
related to the interaction between microbial communities and bio-
representative regions of travertine precipitated after (a) 14 days
films (e.g., Allen et al., 2000; Görgen et al., 2021). Such a morphol-
and (b) 21 days over the glass slide. Magnification: 1000X.
Both micrographs highlight the abundant presencep of calcite ogy was found in association with euhedral calcite, irrespectively of
aggregates, and the clear development of biofilm on the slides after the color of the sampled encrustations and of the microbial commu-
21 days of field incubation. nity composition. This result is in line with the analysis of the lattice
strain, which showed structural features attributed to biomineral-
Furthermore, the occurrence of Epsilonproteobacteria in the water ized calcite (Di Benedetto et al.,  2011). In the present study, the
column (Figure 4b) confirmed the involvement of microbial commu- structural investigation was operated on samples discriminated by
nity in the sulfur cycle along the channel. Members of this class have color with no apparent differences.
indeed been proposed to being actively involved in sulfur cycling, Despite the occurrence of diverse cyanobacterial OTUs at in-
constituting one of the major sulfur oxidizer microbial group in high-­ creasing distance from the hot spring as a function of water tempera-
temperature thermal springs (Campbell et al., 2006). ture, calcite shrubs were constantly present along the whole channel.
The complex microbial community inhabiting the biofilm along Although cyanobacteria are well known to capture CO2 via photo-
the channel was mainly composed of phototrophic micro-­organisms, synthesis and mineralization (Kamennaya et al.,  2012), both empir-
in line with those previously observed in other hot springs (Bilyj ical and theoretical observations excluded a biologically controlled
et al., 2014; Coman et al., 2013; Portillo et al., 2009; Stal et al., 2017; or induced precipitation of calcite shrub deposits, which are ascribed
Wang et al., 2013). In particular, the predominant metabolic pathway to abiotic processes. However, while no evidence was found to indi-
was the oxygenic and anoxygenic phototrophy driven by members of cate an active role by microbial metabolism in calcite precipitation,
Cyanobacteria, Chloroflexi, Chlorobi and Alphaproteobacteria (Coman the shrub fabric might be influenced by the presence of microbial
et al., 2013). A stratified microbial community was retrieved in biofilm mats through organomineralization (e.g., Allen et al., 2000; Bastianini
along the channel (Figure 5), as previously observed in hot spring-­ et al., 2019; Dupraz et al., 2009; Perry et al., 2007). This mineral pre-
bearing microbial mats; in particular, the aerobic photoautotrophic cipitation typically occurs in supersaturated waters by nucleation on
Cyanobacteria inhabited the superficial layer, while the anoxic pho- bacterially produced polymers, such as those composing cell walls or
totrophs dominated underneath (Coman et al., 2013; Hanada, 2016; extracellular polymeric substances (EPS). EPS can indeed account for
|
16      VENTURI et al.

over 90% of biofilm dry mass (Flemming & Wingender, 2010), form- Nevertheless, the occurrence of this mineral phase was widely de-
ing the scaffold for the biofilm architecture and internal cohesion and tected along the channel, though in very localized assemblages.
allowing adhesion to surfaces. Accordingly, Cyanobacteria, which are The morphological observations on the retrieved crystals indicated
among the main contributors to the production of EPS in microbial that gypsum sporadically nucleated and grew very fast (Figure 7e,f).
mats (e.g., Rossi & De Philippis, 2015), dominated the microbial com- These evidences allowed to speculate that the observed gypsum
munity inhabiting the surficial layer of the biofilm along the chan- crystals were the result of a biologically induced mineralization,
nel (Figure  5c). Even though calcite precipitation did not seem to where microbial activity modified the local microenvironment cre-
be related to microbial activity, the cyanobacterial and prokaryotic ating favorable conditions for chemical precipitation, as previously
colonization along the channel occurred simultaneously with the for- observed in other hydrothermal aquatic systems (Tang et al., 2014).
mation of calcite nucleation. In fact, the mutual relationship between It can be hypothesized that S-­oxidizing bacteria produced a local
Spirulina-­like cells, Bacteria, and other prokaryotic cells was evident anomaly in sulfate concentration inducing local supersaturated con-
with the tridimensional structure examination of biofilm grown on ditions and, thus, gypsum precipitation in fractal clusters of crystals.
microscopy slides on the central point of the channel (Figure  6). This hypothesis could be supported by the presence of several S-­
This finding is paralleled with the evolution of the crystal weather- oxidizing bacteria (e.g., Thiofaba, Chloroflexi, and Chlorobi) along the
ing traced by SEM investigation of calcite precipitates as a function channel in both the water column and biofilm layers (Figures 4 and
of time (Figure 9). These evidences agree with those reported by a 5). However, a small source of sulfide (i.e., HS−), suitable for gypsum
recent study on active travertine deposits from Central Italy (Della precipitation, was present in the investigated spring system under
Porta et al., 2021), including the Viterbo thermal area, and with pre- the form of the HS− species, in a ratio to sulfate concentration of
vious observations on the channel (Di Benedetto et al., 2011), which ~3/1000 (Di Benedetto et al., 2011). A similar mechanism was previ-
ascribed the structural anomaly of calcite precipitates to the pres- ously proposed for both Bacteria (Canfora et al., 2016; Thompson &
ence of organic polymers embedded in the growing crystals. Ferris, 1990) and microfungi (Cecchi et al., 2018) in alkaline lake wa-
Concerning the structural anomaly, the presence in the PC cal- ters, saline soil crusts, and sulfide-­rich hardpans. This study showed
cite of the lattice strain is likely to be attributed to a bioinfluenced that the same process can occur even in highly dynamic geochemical
process (Görgen et al., 2021), driven by the presence of biomolecules environments such as travertine depositing hot springs. Contrarily
in the chemical environment during crystallization. Recent studies to what expected from the physicochemical conditions characteriz-
proved that the simple presence of monosaccharide or aminoac- ing the channel, no evidence of dissolution processes was observed
idic molecules, in a supersaturated Ca2+/HCO3− solution, enabled for gypsum. It appeared to be preserved from dissolution when em-
a different crystallization pathway and the presence of the lattice bedded in the biofilm, pointing to a critical role of microbial mats in
strain (Lang et al.,  2020; Mijowska et al.,  2020). A previous study governing the mobility of chemical species at the water–­travertine
showed similar results in a calcite precipitate found in a pure culture interface. These evidences were in line with previous studies (e.g.,
of Bacillus subtilis strain (Perito et al., 2018). Canfora et al., 2016), where cyanobacteria promoted, either directly
Euhedral and shrub calcite precipitates exhibited evidence of or indirectly, the formation of a protective carbonate-­ and sulfate-­
two distinct redissolution processes. The first one provided pro- based envelopes.
gressive consumption of crystals from the edges and/or some ex-
isting pits (Figure 7c), indicating a balance between crystal growth
and redissolution process. The second process consisted of a 6  |  CO N C LU S I O N S
specific pitting, starting from either crystal defects or centers of
well-­formed crystalline surfaces, to produce evident circular pits Travertine encrustations, mainly originated from inorganic precipi-
(Figure 7d). These features are commonly associated with microbial tation, were characterized by the presence of microbial mats with
mats containing sulfur-­oxidizing bacteria (Leprich et al.,  2021) and a well-­stratified microbial community. The latter, mainly shaped by
ascribed to acidity buffering generated during sulfur oxidation at light availability, influenced the morphology of the mineral phases,
the microscale through calcite weathering (e.g., Dupraz et al., 2009; with euhedral calcite crystals (inorganically produced by CO2 de-
Yang et al.,  2019). Accordingly, the observed circular pitting was gassing) coexisting with calcite shrub (likely ascribed to biologically
likely the result of the metabolic activity of the S-­oxidizing bacteria influenced mineralization). The mechanism for calcite precipita-
recognized in the water column (i.e., Thiofaba, Epsilonproteobacteria) tion was likely dependent on the occurrence of microbial EPS, thus
and microbial mats (i.e., Chlorobi, Chloroflexi). The temporal evolu- resulting in the alteration of the crystal structure (i.e., structural
tion investigated by the glass slides inserted in the water flow of the fingerprint), similar to that already described for other forms of bio-
channel allowed to give a rough estimation of the time necessary logically controlled mineralization.
for the precipitate–­biofilm system to attain a stable configuration. In Although the geochemical modelling based on measured ion
particular, we observed that this dynamical equilibrium was reached concentrations indicated undersaturation conditions with respect to
in about 3 weeks. gypsum, mineralogical analyses revealed the occurrence of this min-
Differently from calcite, the numerical model revealed that the eral embedded in the microbial mat. This apparent contradiction be-
PC waters were undersaturated with respect to gypsum (Figure 3e). tween physicochemical environmental conditions and mineralogical
VENTURI et al. |
      17

search for life on Mars. Icarus, 147, 49–­67. https://doi.org/10.1006/


evidences can be reconciled by hypothesizing the establishment of icar.2000.6435
chemical gradients at the microscale triggered by bacterial activity. Amalfitano, S., & Fazi, S. (2008). Recovery and quantification of bac-
In particular, the chemolithoheterotrophic oxidation of reduced sul- terial cells associated with streambed sediments. Journal of
Microbiological Methods, 75, 237–­243. https://doi.org/10.1016/j.
fur operated by the microbial community was expected to produce a
mimet.2008.06.004
local increase in sulfate concentration, determining supersaturated
Amalfitano, S., Fazi, S., Ejarque, E., Freixa, A., Romaní, A. M., & Butturini,
microenvironmental conditions that promoted precipitation of gyp- A. (2018). Deconvolution model to resolve cytometric microbial
sum crystals. This hypothesis was sustained by the identification of community patterns in flowing waters. Cytometry Part A, 93(2),
circular pitting on calcite crystals related to dissolution processes 194–­200. https://doi.org/10.1002/cyto.a.23304
Asao, M., & Madigan, M. T. (2010). Taxonomy, phylogeny, and ecology of
induced by the metabolic activity of sulfur-­oxidizing bacteria. On
the heliobacteria. Photosynthesis Research, 104, 103–­111. https://
the contrary, microbial biofilms may exert a protective functioning doi.org/10.1007/s1112​0 -­0 09-­9516-­1
with respect to gypsum crystals, preventing dissolution despite the Baiocchi, A., Lotti, F., & Piscopo, V. (2012). Conceptual hydrogeological
overall undersaturation conditions of the surrounding water, thus model and groundwater resource estimation in a complex hydro-
thermal area: The case of the viterbo geothermal area (Central
modulating the environmental mobility of chemical species in the
Italy). Journal of Water Resource and Protection, 4, 231–­247. https://
aquatic environment. doi.org/10.4236/jwarp.2012.44026
Banfield, J. F., Moreau, J. W., Chan, C. S., Welch, A., & Little, B. (2004).
AC K N OW L E D G M E N T S Mineralogical biosignatures and the search for life on Mars.
Astrobiology, 1, 447–­465. https://doi.org/10.1089/15311​07017​
CNR-­IRSA participation was supported by Fondazione CARIPLO
53593856
contract No. 2014-­1301 (BATA Project). Maria Rosa Siena e Giulia Bastianini, L., Rogerson, M., Mercedes-­M artín, R., Prior, T. J., Cesar,
Giagnoli (CNR-­IRSA) are gratefully acknowledged for the help in E. A., & Mayes, W. M. (2019). What causes carbonates to form
sampling and CARD-­FISH analysis. Laura Chiarantini and Mario “shrubby” morphologies? An Anthropocene limestone case
study. Frontiers in Earth Science, 7, 236. https://doi.org/10.3389/
Paolieri, and Laura Chelazzi are gratefully acknowledged for their
feart.2019.00236
unvaluable help in the SEM and XRD investigation, respectively. The Bilyj, M., Lepitzki, D., Hughes, E., Swiderski, J., Stackebrandt, E., Pacas,
MEMA (Microscopy and Microanalysis) and the CRIST (Structural C., & Yurkov, V. V. (2014). Abundance and diversity of the phototro-
Crystallography) Interdepartmental Centers of the University of phic microbial mat communities of Sulphur Mountain Banff Springs
and their significance to the endangered snail, Physella johnsoni.
Florence are acknowledged for provision of experimental access at
Open Journal of Ecology, 4, 488–­516.
their facilities. FDB acknowledges University of Florence, for grant-
Bolger, A. M., Lohse, M., & Usadel, B. (2014). Trimmomatic: A flexible
ing him under the Progetto di Ateneo funds, and the University of trimmer for Illumina sequence data. Bioinformatics, 30, 2114–­2120.
Ferrara, for granting him under the FAR2021 and FAR2022 funds. https://doi.org/10.1093/bioin​forma​tics/btu170
Financial support by the Laboratory of Fluid Geochemistry of Cadogan, S. P., Maitland, G. C., & Trusler, P. M. (2014). Diffusion coef-
ficients of CO2 and N2 in water at temperatures between 298.15
the Department of Earth Sciences (University of Florence) is also
K and 423.15 K at pressures up to 45 MPa. Journal of Chemical &
acknowledged. Engineering Data, 59, 519–­525.
Campbell, B. J., Engel, A. S., Porter, M. L., & Takai, K. (2006). The versatile
C O N FL I C T O F I N T E R E S T ε-­proteobacteria: Key players in sulphidic habitats. Nature Reviews
Microbiology, 4, 458–­468. https://doi.org/10.1038/nrmic​ro1414
The authors declare no conflict of interest.
Canfora, L., Vendramin, E., Vittori, A. L., Lo, P. G., Dazzi, C., Benedetti,
A., Iavazzo, P., Adamo, P., Jungblut, A., & Pinzari, F. (2016).
DATA AVA I L A B I L I T Y S TAT E M E N T Compartmentalization of gypsum and halite associate with cyano-
The data that support the findings of this study are available from bacteria in saline soil crusts. FEMS Microbiology Ecology, 92, fiw080.
https://doi.org/10.1093/femse​c/fiw080
the corresponding author upon reasonable request.
Caporaso, J. G., Kuczynski, J., Stombaugh, J., Bittinger, K., Bushman, F.
D., Costello, E. K., Fiere, N., Gonzalez, P. A., Goodrich, J. K., Gordon,
ORCID J. I., Huttley, G. A., Kelley, S. T., Knights, D., Koenig, J. E., Ley, R.
Simona Crognale  https://orcid.org/0000-0002-0322-9543 E., Lozupone, C. A., McDonald, D., Muegge, B. D., Pirrung, M., …
Knight, R. (2010). QIIME allows analysis of high-­throughput com-
Stefano Amalfitano  https://orcid.org/0000-0002-6148-1472
munity sequencing data. Nature Methods, 7, 335–­336. https://doi.
Stefano Fazi  https://orcid.org/0000-0001-7688-7179 org/10.1038/nmeth.f.303
Caporaso, J. G., Lauber, C. L., Walters, W. A., Berg-­Lyons, D., Huntley, J.,
REFERENCES Fierer, N., Owens, S. M., Betley, J., Fraser, L., Bauer, M., Gormley, N.,
Gilbert, J. A., Smith, G., & Knight, R. (2012). Ultra-­high-­throughput
Albertsen, M., Karst, S. M., Ziegler, A. S., Kirkegaard, R. H., & Nielsen, P.
microbial community analysis on the Illumina HiSeq and MiSeq plat-
H. (2015). Back to basics –­ The influence of DNA extraction and
forms. The ISME Journal, 6, 1621–­1624. https://doi.org/10.1038/
primer choice on phylogenetic analysis of activated sludge com-
ismej.2012.8
munities. PLoS One, 10, e0132783. https://doi.org/10.1371/journ​
Casentini, B., Falcione, F. T., Amalfitano, S., Fazi, S., & Rossetti, S. (2016).
al.pone.0132783
Arsenic removal by discontinuous ZVI two steps system for drink-
Allen, C. C., Albert, F. G., Chafetz, H. S., Combie, J., Graham, C. R., Kieft,
ing water production at household scale. Water Research, 106,
T. L., Kivett, S. J., McKay, D. S., Steele, A., Taunton, A. E., Taylor,
135–­145.
M. R., Thomas-­Keprta, K. L., & Westall, F. (2000). Microscopic
Castro-­Alonso, M. J., Montañez-­Hernandez, L. E., Sanchez-­Muñoz, M.
physical biomarkers in carbonate hot springs: Implications in the
A., Franco, M. R. M., Narayanasamy, R., & Balagurusamy, N. (2019).
|
18      VENTURI et al.

Microbially induced calcium carbonate precipitation (MICP) and community composition. PLoS One, 8, e64109. https://doi.
its potential in bioconcrete: Microbiological and molecular con- org/10.1371/journ​al.pone.0064109
cepts. Frontiers in Materials, 6, 126. https://doi.org/10.3389/ Flemming, H. C., & Wingender, J. (2010). The biofilm matrix. Nature
fmats.2019.00126 Reviews, 8, 623–­633.
Cecchi, G., Marescotti, P., Di Piazza, S., Lucchetti, G., Mariotti, M. G., & Gaisin, V. A., Grounzdev, D. S., Namsaraev, Z. B., Sukhacheva, M. V.,
Zotti, M. (2018). Gypsum biomineralization in sulphide-­rich hard- Gorlenko, V. M., & Kuznetsov, B. B. (2016). Biogeography of ther-
pans by a native Trichoderma harzianum Rifai strain. Geomicrobiology mophilic phototrophic bacteria belonging to Roseiflexus genus.
Journal, 35, 209–­214. https://doi.org/10.1080/01490​ FEMS Microbiology Ecology, 92, fiw012. https://doi.org/10.1093/
451.2017.1362077 femse​c/fiw012
Chiriac, C. M., Szekeres, E., Rudi, K., Baricz, A., Hegedus, A., Dragoş, N., Giampaoli, S., Valeriani, F., Gianfranceschi, G., Vitali, M., Delfini, M.,
& Coman, C. (2017). Differences in temperature and water chemis- Festa, M. R., Bottari, E., & Romano, S. V. (2013). Hydrogen sulfide
try shape distinct diversity patterns in thermophilic microbial com- in thermal spring waters and its action on bacteria of human origin.
munities. Applied and Environmental Microbiology, 83, e01363-­17. Microchemical Journal, 108, 210–­214. https://doi.org/10.1016/j.
https://doi.org/10.1128/AEM.01363​-­17 microc.2012.10.022
Cinti, D., Tassi, F., Procesi, M., Bonini, M., Capecchiacci, F., Voltattorni, N., Görgen, S., Benzerara, K., Skouri-­Panet, F., Gugger, M., Chauvat, F., &
Vaselli, O., & Quattrocchi, F. (2014). Fluid geochemistry and geo- Cassier-­Chauvat, C. (2021). The diversity of molecular mechanisms
thermometry in the unexploited geothermal field of the Vicano–­ of carbonate biomineralization by bacteria. Discover Materials, 1, 2.
Cimino Volcanic District (Central Italy). Chemical Geology, 371, https://doi.org/10.1007/s4393​9-­020-­0 0001​-­9
96–­114. Graf, D. L. (1961). Crystallographic tables for the rhombohedral carbon-
Coman, C., Druga, B., Hegedus, A., Sicora, C., & Dragoş, N. (2013). ates. American Mineralogist, 46, 1283–­1316.
Archaeal and bacterial diversity in two hot spring microbial mats Greuter, D., Loy, A., Horn, M., & Rattei, T. (2016). ProbeBase-­an online re-
from a geothermal region in Romania. Extremophiles, 17, 523–­534. source for rRNA-­t argeted oligonucleotide probes and primers: New
https://doi.org/10.1007/s0079​2-­013-­0537-­5 features 2016. Nucleic Acids Research, 44, D586–­D589. https://doi.
Corey, A. T. (1954). The interrelation between gas and oil relative perme- org/10.1093/nar/gkv1232
abilities. Producers Monthly, 19, 38–­41. Gulecal-­Pektas, Y., & Temel, M. (2017). A window to the subsur-
Dai, Z., Kan, A. T., Shi, W., Zhang, N., Zhang, F., Yan, F., Bhandari, N., face: Microbial diversity in Hot Springs of a sulfidic cave (Kaklik,
Zhang, Z., Liu, Y., Ruan, G., & Tomson, M. B. (2017). Solubility Turkey). Geomicrobiology Journal, 34, 374–­384. https://doi.
measurements and predictions of gypsum, anhydrite, and calcite org/10.1080/01490​451.2016.1204374
over wide ranges of temperature, pressure, and ionic strength Gumerov, V. M., Mardanov, A. V., Beletsky, A. V., Osmolovskaya, E. A.
with mixed electrolytes. Rock Mechanics and Rock Engineering, 50, B., & Ravin, N. V. (2011). Molecular analysis of microbial diversity
327–­339. in the Zavarzin spring, Uzon caldera, Kamchatka. Microbiology, 80,
Della Porta, G., Hoppert, M., Hallmann, C., Schneider, D., & Reitner, J. 244–­251. https://doi.org/10.1134/S0026​26171​102007X
(2021). The influence of microbial mats on travertine precipitation Hanada, S. (2016). Anoxygenic photosynthesis—­A photochemi-
in active hydrothermal systems (Central Italy). The Depositional cal reaction that does not contribute to oxygen reproduction.
Record, 8, 165–­209. https://doi.org/10.1002/dep2.147 Microbes and Environments, 31, 1–­3. https://doi.org/10.1264/
Des Marais, D. J., & Walter, M. R. (2019). Terrestrial hot spring systems: jsme2.ME3101rh
Introduction. Astrobiology, 19, 1419–­1432. https://doi.org/10.1089/ Huang, Q., Dong, C. Z., Dong, R. M., Jiang, H., Wang, S., Wang, G., Fang,
ast.2018.19761419 B., Ding, X., Niu, L., & Li, X. (2011). Archaeal and bacterial diver-
Di Benedetto, F., Montegrossi, G., Minissale, A., Pardi, L. A., Romanelli, sity in hot springs on the Tibetan plateau, China. Extremophiles, 15,
M., Tassi, F., Delgado, H. A., Pampin, E. M., Vaselli, O., & Borrini, 549–­563. https://doi.org/10.1007/s0079​2-­011-­0386-­z
D. (2011). Biotic and inorganic control on travertine deposition at Illumina. (2015). 16S metagenomic sequencing library preparation, Part
Bullicame 3 spring (Viterbo, Italy): A multidisciplinary approach. # 15044223 Rev B. Illumina https://suppo​r t.illum​ina.com/docum​
Geochimica et Cosmochimica Acta, 75, 4441–­4 455. https://doi. ents/docum​e ntat​i on/chemi​s try_docum​e ntat​i on/16s/16s-­m etag​
org/10.1016/j.gca.2011.05.011 enomi​c-­libra​r y-­prep-­guide​-­15044​223-­b.pdf
Duchi, V., & Minissale, A. (1995). Distribuzione delle manifestazioni gas- Imhoff, J. F. (2005). Genus XVI. Rhodomicrobium. In D. J. Brenner, N. R.
sose nel settore peritirrenico Tosco–­L aziale e loro interazione con Krieg, & J. T. Staley (Eds.), Bergey's manual of systematic and determi-
gli acquiferi superficiali. Bollettino della Societa Geologica Italiana, native bacteriology (pp. 543–­454). Springer.
114, 337–­351. Imhoff, J. F. (2014). Biology of green sulfur bacteria. In eLS. John Wiley
Duchi, V., Minissale, A., & Romani, L. (1985). Studio geochimico su acque & Sons Ltd. https://doi.org/10.1002/97804​70015​902.a0000​458.
e gas dell'area geotermica lago di Vico-­M.Cimini (Viterbo). Atti pub2
della Societa Toscana di Scienze Naturali Residente in Pisa, Memorie, Jansson, C., & Northern, T. (2010). Calcifying cyanobacteria –­ The po-
Processi Verbali, Serie A, 92, 237–­254. tential of biomineralization for carbon capture and storage. Current
Dupraz, C., Reid, R. P., Braissant, O., Decho, A. W., Norman, R. S., & Opinion in Biotechnology, 21, 365–­371. https://doi.org/10.1016/j.
Visscher, P. T. (2009). Processes of carbonate precipitation in mod- copbio.2010.03.017
ern microbial mats. Earth-­Science Reviews, 96, 141–­162. https://doi. Kambura, A. K., Mwirichia, R. K., Kasili, R. W., Karanja, E. N., Makonde,
org/10.1016/j.earsc​irev.2008.10.005 H. M., & Boga, H. I. (2016). Bacteria and archaea diversity within
Evans, W. C., White, L. D., & Rapp, J. B. (1988). Geochemistry of some the hot springs of Lake Magadi and Little Magadi in Kenya.
gases in hydrothermal fluids from the southern Juan de Fuca ridge. BMC Microbiology, 16, 1–­12. https://doi.org/10.1186/s1286​
Journal of Geophysical Research, 93(B12), 15305–­15313. 6-­016-­0748-­x
Fazi, S., Amalfitano, S., Pizzetti, I., & Pernthaler, J. (2007). Efficiency of Kamennaya, N. A., Ajo-­Franklin, C. M., Northern, T., & Jansson, C.
fluorescence in situ hybridization for bacterial cell identification (2012). Cyanobacteria as biocatalysts for carbonate mineralization.
in temporary river sediments with contrasting water content. Minerals, 2, 338–­364. https://doi.org/10.3390/min20​4 0338
Systematic and Applied Microbiology, 30, 463–­470. https://doi. Kemp, P., & Aller, J. (2004). Bacterial diversity in aquatic and other
org/10.1016/j.syapm.2007.03.003 environments: What 16S rDNA libraries can tell us. FEMS
Fazi, S., Vázquez, E., Casamayor, E. O., Amalfitano, S., & Butturini, A. Microbiology Ecology, 47, 161–­177. https://doi.org/10.1016/S0168​
(2013). Stream hydrological fragmentation drives bacterioplankton -­6 496(03)00257​-­5
VENTURI et al. |
      19

Konhauser, K. (2007). Introduction to geomicrobiology. Blackwell Pagaling, E., Grant, W. D., Cowan, D. A., Jones, B. E., Ma, Y., Ventosa,
Publishing. A., & Heaphy, S. (2012). Bacterial and archaeal diversity in two hot
Krumgalz, B. S. (2018). Temperature dependence of mineral solubil- spring microbial mats from the geothermal region of Tengchong,
ity in water. Part 3. Alkaline and alkaline earth sulfates. Journal of China. Extremophiles, 16, 607–­618. https://doi.org/10.1007/s0079​
Physical and Chemical Reference Data, 47, 023101–­1/30. https://doi. 2-­012-­0 460-­1
org/10.1063/1.5031951 Palandri, J. L., & Kharaka, Y. K. (2004). A compilation of rate parameters
Lang, A., Mijowska, S., Polishchuk, I., Fermani, S., Falini, G., Katsman, of water-­ mineral interaction kinetics for application to geochem-
A., Marin, F., & Pokroy, B. (2020). Acidic monosaccharides be- ical modeling. USGS open file report 2004-­1068, Menlo Park,
come incorporated into calcite single crystals. Chemistry, 26, California, National Energy Technology Laboratory –­United States
16860–­16868. Department of Energy, 64 pp.
Lebaron, P., Servais, P., Agogue, H., Courties, C., & Joux, F. (2001). Does Pentecost, A. (1995). Geochemistry of carbon dioxide in six travertine-­
the high nucleic acid content of individual bacterial cells allow us depositing waters of Italy. Journal of Hydrology, 167, 263–­278.
to discriminate between active cells and inactive cells in aquatic Perito, B., Romanelli, M., Buccianti, A., Passaponti, M., Montegrossi, G.,
systems? Applied and Environmental Microbiology, 67, 1775–­1782. & Di Benedetto, F. (2018). An XRPD and EPR spectroscopy study of
https://doi.org/10.1128/AEM.67.4.1775-­1782.2001 microcrystalline calcite bioprecipitated by Bacillus subtilis. Physics
Leprich, D. J., Flood, B. E., Schroedl, P. R., Ricci, E., Marlow, J. J., Girguis, and Chemistry of Minerals, 45, 935–­944.
P. R., & Bailey, J. V. (2021). Sulfur bacteria promote dissolution of Perry, R. S., Mcloughlin, N., Lynne, B. Y., Sephton, M. A., Oliver, J. D.,
authigenic carbonates at marine methane seeps. The ISME Journal, Perry, C. C., Campbell, K., Engel, M. H., Farmer, J. D., Brasier, M.
15, 2043–­2056. https://doi.org/10.1038/s4139​6-­021-­0 0903​-­3 D., & Staley, J. T. (2007). Defining biominerals and organominerals:
Liu, Z., Klatt, C. G., Wood, J. M., Rusch, D. B., Ludwig, M., Wittekindt, Direct and indirect indicators of life. Sedimentary Geology, 201, 157–­
N., Tomsho, L. P., Schuster, S. C., Ward, D. M., & Bryant, D. A. 179. https://doi.org/10.1016/j.sedgeo.2007.05.014
(2011). Metatranscriptomic analyses of chlorophototrophs of a hot-­ Pillay, V., Gartner, R. S., Himawan, C., Seckler, M. M., Lewis, A. E., &
spring microbial mat. The ISME Journal, 5, 1279–­1290. https://doi. Witkamp, G.-­J. (2005). MgSO4 + H2O system at eutectic conditions
org/10.1038/ismej.2011.37 and thermodynamic solubility products of MgSO4·12H2O(s) and
Lupini, G., Proia, L., Di Maio, M., Amalfitano, S., & Fazi, S. (2011). CARD-­ MgSO4·7H2O(s). Journal of Chemical & Engineering Data, 50, 551–­555.
FISH and confocal laser scanner microscopy to assess successional Piscopo, V., Barbieri, M., Monetti, V., Pagano, G., Pistoni, S., Ruggi, E., &
changes of the bacterial community in freshwater biofilms. Journal Stanzione, D. (2006). Hydrogeology of thermal waters in Viterbo
of Microbiological Methods, 86, 248–­251. https://doi.org/10.1016/j. area, Central Italy. Hydrogeology Journal, 14, 1508–­1521. https://
mimet.2011.05.011 doi.org/10.1007/s1004​0 -­0 06-­0 090-­8
Madigan, M. T., Schaaf, N. A. V., & Sattley, W. M. (2017). The chlorobia- Portillo, M. C., Sririn, V., Kanoksilapatham, W., & Gonzalez, J. M. (2009).
ceae, chloroflexaceae, and heliobacteriaceae. In P. Hallenbeck (Ed.), Differential microbial communities in hot spring mats from Western
Modern topics in the phototrophic prokaryotes. Springer. https://doi. Thailand. Extremophiles, 13, 321–­331. https://doi.org/10.1007/
org/10.1007/978-­3-­319-­46261​-­5_4 s0079​2-­0 08-­0219-­x
McCutcheon, J., Power, I. M., Harrison, A. L., Dipple, G. M., & Southam, Prieto-­B arajas, C. M., Valencia-­C antero, E., & Santoyo, G. (2018).
G. (2014). A greenhouse-­scale photosynthetic microbial bioreac- Microbial mat ecosystems: Structure types, functional di-
tor for carbon sequestration in magnesium carbonate minerals. versity, and biotechnological application. Electronic Journal
Environmental Science & Technology, 48, 9142–­9151. https://doi. of Biotechnology, 31, 48–­56. https://doi.org/10.1016/j.
org/10.1021/es500​3 44s ejbt.2017.11.001
McIlroy, S. J., Saunders, A. M., Albertsen, M., Nierychlo, M., McIlroy, B., Pruess, K., Oldenburg, C., & Moridis, G. (2012). TOUGH2 user's guide, ver-
Hansen, A. A., Karst, S. M., Nielsen, J. L., & Nielsen, P. H. (2015). sion 2. Lawrence Berkeley National Laboratory LBNL-­43134.
MiDAS: the field guide to the microbes of activated sludge. Rimondi, V., Costagliola, P., Lattanzi, P., Catelani, T., Fornasaro, S., Medas,
Database, 2015, bav062. https://doi.org/10.1093/datab​ase/ D., Morelli, G., & Paolieri, M. (2021). Bioaccessible arsenic in soil
bav062 of thermal areas of Viterbo, Central Italy: Implications for human
Mijowska, S., Polishchuk, I., Lang, A., Seknazi, E., Dejoie, C., Fermani, S., health risk. Environmental Geochemistry and Health, 44, 465–­485.
Falini, G., Demitri, N., Polentarutti, M., Katsman, A., & Pokroy, B. https://doi.org/10.1007/s1065​3-­021-­0 0914​-­1
(2020). High amino acid lattice loading at nonambient conditions Rodriguez-­C arvajal, J. (1993). Recent advances in magnetic structure
causes changes in structure and expansion coefficient of calcite. determination by neutron powder diffraction. Physica B: Condensed
Chemistry of Materials, 32, 4205–­4212. Matter, 192, 55–­69. https://doi.org/10.1016/0921-­4526(93)90108​-­I
Minissale, A. (2004). Origin, transport and discharge of CO2 in Central Rossi, F., & De Philippis, R. (2015). Role of cyanobacterial exopolysac-
Italy. Earth-­Science Reviews, 66, 89–­141. https://doi.org/10.1016/j. charides in phototrophic biofilms and in complex microbial mats.
earsc​irev.2003.09.001 Lifestyles, 5, 1218–­1238. https://doi.org/10.3390/life5​021218
Minissale, A., Kerrick, D. M., Magro, G., Murrell, M. T., Paladini, M., Rihs, Salata, G. G., Roelke, L. A., & Cifuentes, L. A. (2000). A rapid and precise
S., Sturchio, N. C., Tassi, F., & Vaselli, O. (2002). Geochemistry method for measuring stable carbon isotope ratios of dissolved in-
of quaternary travertines in the region north of Rome (Italy): organic carbon. Marine Chemistry, 69, 153–­161.
Structural, hydrologic and paleoclimatic implications. Earth and Schuler, C. G., Having, J. R., & Hamilton, T. L. (2017). Hot spring micro-
Planetary Science Letters, 203, 709–­728. https://doi.org/10.1016/ bial community composition, morphology, and carbon fixation:
S0012​-­821X(02)00875​- ­0 Implications for interpreting the ancient rock record. Frontiers in
Mori, K., & Suzuki, K. (2008). Thiofaba tepidiphila gen. Nov., sp. nov., a Earth Science, 5, 97. https://doi.org/10.3389/feart.2017.00097
novel obligately chemolithoautotrophic, sulfur-­oxidizing bacterium Sonnenthal, E., & Ortoleva, P. J. (1994). Numerical simulations of
of the Gammaproteobacteria isolated from a hot spring. International Overpressured compartments in sedimentary basins. In P. J.
Journal of Systematic and Evolutionary Microbiology, 58, 1885–­1891. Ortoleva & Z. Al-­Shaieb (Eds.), Basin compartments and seals,
https://doi.org/10.1099/ijs.0.65754​- ­0 Memoir 61 (pp. 403–­416). American Association of Petroleum
Murphy, W. M., Oelkers, E. H., & Lichtner, P. C. (1989). Surface reaction Geologists. https://doi.org/10.1306/M6158​8C26
versus diffusion control of mineral dissolution and growth rates in Stal, L. J., Bolhuis, H., & Cretoiu, M. S. (2017). Phototrophic microbial
geochemical processes. Chemical Geology, 78, 357–­380. Mats. In P. Hallenbeck (Ed.), Modern topics in the phototrophic
|
20      VENTURI et al.

prokaryotes. Springer. https://doi.org/10.1007/978-­3-­319-­46261​ Venturi, S., Tassi, F., Bicocchi, G., Cabassi, J., Capecchiacci, F., Capasso,
-­5_9 G., Vaselli, O., Ricci, A., & Grassa, F. (2017). Fractionation processes
Steefel, C. I., & Maher, K. (2009). Fluid-­rock interaction: A reactive trans- affecting the stable carbon isotope signature of thermal waters
port approach. Reviews in Mineralogy and Geochemistry, 70, 485–­ from hydrothermal/volcanic systems: The examples of Campi
532. https://doi.org/10.2138/rmg.2009.70.11 Flegrei and Volcano Island (southern Italy). Journal of Volcanology
Sundberg, C., Al-­Soud, W. A., Larsson, M., Alm, E., Shakeri, Y. S., Svensson, and Geothermal Research, 345, 46–­57. https://doi.org/10.1016/j.
B. H., Sørensen, S. J., & Karlsson, A. (2013). 454-­pyrosequencing jvolg​eores.2017.08.001
analyses of bacterial and archaeal richness in 21 full-­scale biogas Wang, Q., Garrity, G. M., Tiedje, J. M., & Cole, J. R. (2007). Naive Bayesian
digesters. FEMS Microbiology Ecology, 85, 612–­626. https://doi. classifier for rapid assignment of rRNA sequences into the new bac-
org/10.1111/1574-­6941.12148 terial taxonomy. Applied and Environmental Microbiology, 73, 5261–­
Tang, M., Ehreiser, A., & Li, Y. L. (2014). Gypsum in modern Kamchatka 5267. https://doi.org/10.1128/AEM.00062​- ­07
volcanic hot springs and the lower Cambrian black shale: Applied Wang, S., Hou, W., Dong, H., Jiang, H., Huang, L., Wu, G., Zhang, C.,
to the microbial-­mediated precipitation of sulfates on Mars. Song, Z., Zhang, Y., Ren, H., Zhang, J., & Zhang, L. (2013). Control
American Mineralogist, 99, 2126–­2137. https://doi.org/10.2138/ of temperature on microbial community structure in hot springs of
am-­2014-­4754 the Tibetan plateau. PLoS One, 8, e62901. https://doi.org/10.1371/
Tank, M., Thiel, V., Ward, D. M., & Bryant, D. A. (2017). A panoply of pho- journ​al.pone.0062901
totrophs: An overview of the thermophilic Chlorophototrophs of Wilmeth, D. T., Johnson, H. A., Stamps, B. W., Berelson, W. M., Stevenson,
the microbial Mats of alkaline siliceous Hot Springs in Yellowstone B. S., Nunn, H. S., Grim, S. L., Dillon, M. L., Paradis, O., Corsetti, F.
National Park, WY, USA. In P. Hallenbeck (Ed.), Modern topics in the A., & Spear, J. R. (2018). Environmental and biological influences
phototrophic prokaryotes. Springer. https://doi.org/10.1007/978-­3-­ on carbonate precipitation within hot spring microbial Mats in
319-­46261​-­5_3 Little Hot Creek, CA. Frontiers in Microbiology, 9, 1464. https://doi.
Tassi, F., Fazi, S., Rossetti, S., Pratesi, P., Ceccotti, M., Cabassi, J., org/10.3389/fmicb.2018.01464
Capecchiacci, F., Venturi, S., & Vaselli, O. (2018). The biogeochem- Xu, T., & Pruess, K. (2001). Modeling multiphase nonisothermal fluid flow
ical vertical structure renders a meromictic volcanic Lake a trap and reactive geochemical transport in variably saturated fractured
for geogenic CO2 (lake Averno, Italy). PLoS One, 13(3), e0193914. rocks: 1 Methodology. American Journal of Science, 301, 16–­33.
https://doi.org/10.1371/journ​al.pone.0193914 Xu, T., Sonnenthal, E. L., Spycher, N., & Pruess, K. (2006). TOUGHREACT:
Tassi F., Vaselli O., Luchetti G., Montegrossi G., & Minissale A. (2008). A simulation program for non-­isothermal multiphase reactive geo-
Metodo per la determinazione dei gas disciolti in acque naturali. Int. chemical transport in variably saturated geologic media. Computers
Rep., CNR-­IGG, Florence, Italy, pp. 11. & Geosciences, 32, 145–­165.
Tassi, F., Vaselli, O., Tedesco, D., Montegrossi, G., Darrah, T., Cuoco, Yang, T., Teske, A., Ambrose, W., Salman-­C arvalho, V., Bagnell, R., &
E., Mapendano, M. Y., Poreda, R., & Delgado, H. A. (2009). Water Nielsen, L. P. (2019). Intracellular calcite and sulfur dynamics of
and gas chemistry at Lake Kivu (DRC): Geochemical evidence of Achromatium cells observed in a lab-­based enrichment and aerobic
vertical and horizontal heterogeneities in a multi-­basin structure. incubation experiment. Antonie Van Leeuwenhoek, 112, 263–­274.
Geochemistry, Geophysics, Geosystems, 10(2), 1–­22. https://doi. https://doi.org/10.1007/s1048​2-­018-­1153-­2
org/10.1029/2008G​C002191 Zhang, J., Quay, P. D., & Wilbur, D. O. (1995). Carbon isotope frac-
Tassi, F., Venturi, S., Cabassi, J., Vaselli, O., Gelli, I., Cinti, D., & tionation during gas-­water exchange and dissolution of CO2.
Capecchiacci, F. (2015). Biodegradation of CO2, CH4 and volatile Geochimica et Cosmochimica Acta, 59(1), 107–­114. https://doi.
organic compounds (VOCs) in soil gas from the Vicano-­Cimino hy- org/10.1016/0016-­7037(95)91550​-­D
drothermal system (Central Italy). Organic Geochemistry, 86, 81–­93.
https://doi.org/10.1016/j.orgge​ochem.2015.06.004
Thompson, J. B., & Ferris, F. G. (1990). Cyanobacterial precipitation of S U P P O R T I N G I N FO R M AT I O N
gypsum, calcite, and magnesite from natural alkaline lake water.
Additional supporting information can be found online in the
Geology, 18, 995–­998.
Valeriani, F., Crognale, S., Protano, C., Gianfranceschi, G., Orsini, M., Supporting Information section at the end of this article.
Vitali, M., & Spica, V. R. (2018). Metagenomic analysis of bacterial
community in a travertine depositing hot spring. New Microbiologica,
41, 126–­135. How to cite this article: Venturi, S., Crognale, S.,
Van Driessche, A. E. S., Stawski, T. M., & Kellermeier, M. (2019). Calcium
Di Benedetto, F., Montegrossi, G., Casentini, B., Amalfitano,
sulfate precipitation pathways in natural and engineered environ-
S., Baroni, T., Rossetti, S., Tassi, F., Capecchiacci, F., Vaselli,
ments. Chemical Geology, 530, 119274. https://doi.org/10.1016/j.
chemg​eo.2019.119274 O., & Fazi, S. (2022). Interplay between abiotic and microbial
van Gemerden, H. (1993). Microbial mats: A joint venture. Marine biofilm-­mediated processes for travertine formation: Insights
Geology, 113, 3–­25. from a thermal spring (Piscine Carletti, Viterbo, Italy).
Vaselli, O., Tassi, F., Montegrossi, G., Capaccioni, B., & Giannini, L. (2006).
Geobiology, 00, 1–20. https://doi.org/10.1111/gbi.12516
Sampling and analysis of volcanic gases. Acta Vulcanologica, 18,
65–­76.

You might also like