Next Article in Journal
Nitric Oxide Increases the Physiological and Biochemical Stability of Soybean Plants under High Temperature
Next Article in Special Issue
Bioregulators Can Improve Biomass Production, Photosynthetic Efficiency, and Ornamental Quality of Gazania rigens L.
Previous Article in Journal
Multi-Factor Diagnostic and Recommendation System for Boron in Neutral and Acidic Soils
Previous Article in Special Issue
Application of Proximal Optical Sensors to Fine-Tune Nitrogen Fertilization: Opportunities for Woody Ornamentals
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Cultivation Substrate Composition Influences Morphology, Volatilome and Essential Oil of Lavandula Angustifolia Mill.

1
Department of Pharmacy, University of Pisa, Via Bonanno 6, 56126 Pisa, Italy
2
Department of Agricultural, Forest and Food Sciences, University of Torino, Largo Paolo Braccini 2, 10095 Grugliasco (TO), Italy
*
Author to whom correspondence should be addressed.
Agronomy 2019, 9(8), 411; https://doi.org/10.3390/agronomy9080411
Submission received: 28 June 2019 / Revised: 22 July 2019 / Accepted: 25 July 2019 / Published: 26 July 2019

Abstract

:
Aromatic plants are commonly produced for ornamental, cosmetic and medicinal purposes. Their morphological traits and the amounts and compositions of the volatile substances and essential oils (EOs) produced can be influenced by several factors, including the cultivation technique. In the present study, the influence of substrate composition on Lavandula angustifolia Mill. production was evaluated. In particular, substrates containing mixes of peat (P), green compost (C) and/or demolition aggregates (A) were tested in the following ratios: 70%:30% v/v, P:C; 70%:30% v/v, P:A; and 40%:30%:30% v/v, P:C:A. The P:C mixture allowed to obtain the best results in terms of survival rate, compactness of the plant and flower production. The P:C:A led to higher yields and better quality of EOs, with higher amounts of linalool, an important compound for medicinal uses. The volatiles and the blooming trend were not affected by the different cultivation substrates. Therefore, substrate composition in pot lavender cultivation can be regulated depending on the final use of the plant, successfully using locally sourced material in addition to peat.

1. Introduction

Lavender, Lavandula angustifolia Mill. (Lamiaceae family), is a small evergreen shrub, with aromatic foliage and flowers, mostly cultivated in open fields on well-drained and lime-rich soils in full sun [1,2]. The great economic importance of lavender is due to the high quality of its essential oil (EO), of which 200 tons are produced worldwide every year [3]. The lavender EO is considered one of the most medically useful EOs. It has antibacterial and antifungal activities [4] and is being used to treat infections [5] and neurological disorders [1]. The delightful perfume of this oil has also been widely used in different industries such as perfumery and cosmetics [6] as well as flavoring in food manufacturing [7]. The first phase of lavender cultivation generally occurs in pots, with peat as the main constituent of growing media used throughout Europe (European Peat and Growing Media Association, EPAGMA). However, numerous issues are related to the use of peat as a cultivation substrate [8,9] and its demand and costs are rising [10]. Currently, composts are widely used as component for potting mixtures in many peat-based cultivations [9], and composted green waste is the most widely utilized compost in Europe [8,10,11]. However, composted materials often lack the large particles necessary for adequate aeration, which is improved by the addition of coarse components that can be made of several materials [12]. The use of locally sourced materials as potting media is an object of intensive investigation in floriculture and horticulture. Indeed, several studies have investigated the potential of waste derived from agriculture, specifically nutshells, rice husks, coconut fiber, cattle manure and peanut shells [13,14,15,16,17,18]. Municipal wastes are also of interest; these include green compost, vermicompost and sewage sludge [16,19,20]. Finally, industrial by-products (e.g., biochar, pine bark and bamboo residues) have been studied [16,21,22]. The suitability of a growing medium for pot cultivation depends on several physical, chemical and biological characteristics [23], and establishing the best proportion of materials to obtain good plant growth and productivity results is a very complex process [24].
The scientific research to improve the quality of cultivated medicinal and aromatic plants commonly focuses on fertilization [25,26,27,28,29,30] or irrigation regimes [31,32,33,34]. Nonetheless, previous experimental studies revealed that many other factors interfere with the volatile and essential oil composition: genotype [35], latitude [36], developmental conditions [37,38], harvesting time [39], development stage and environmental conditions [38,40], such as temperature [41]. The effects of different mixtures used as substrates for potting plants have been reported in few studies on lavender [42] and other Lamiaceae plants, such as Thymus vulgaris L. [43], Ocimum basilicum L. [39] and Rosmarinus officinalis L. [44]. Whilst the impact of a growing medium on plant performance is important, it does need to be evaluated in the context of a commercial plant production system while using an adaptive approach [8]. This work deals with the effects of different peat-based potting substrates composed of a mixture of peat and organic and/or mineral material from local markets (green compost and demolition aggregates) on the growth, appearance and composition of aroma and EOs of L. angustifolia in pots in a nursery production system.

2. Materials and Methods

2.1. Soilless Cultivation

Three populations of Lavandula angustifolia of North West Italian Alps and referred to as Susa Valley (Sus), Stura Valley (Stu) and Tanaro Valley (Tan) according to their geographical origin, were selected based on their different phytochemical profile [36]. Cutting propagation occurred in September 2014 under plastic tunnels in the nursery Fratelli Gramaglia (Collegno, Italy; 45°05′22.4′′ N, 7°34′26.4′′ E, 302 m.a.s.l.). The soilless cultivation trial started in March 2015, when rooted plants were transplanted in 1.2 L pots (one rooted plant each). Peat (P), green compost (C) and demolition aggregates (A) were used to prepare three different mixtures to be tested as substrates (Table 1): P:C, 70%:30% v/v; P:A, 70%:30% v/v; and P:C:A, 40%:30%:30% v/v, with 180 plants grown in each substrate (60 Sus, 60 Stu, 60 Tan). Peat and green compost were provided by the local fertilizer producer ItalConcimi S.r.l., while the demolition aggregates were supplied by Perino Piero & C. S.n.c.; both companies were located within 20 km from the nursery. Demolition aggregates were composed of 70% bricks and 30% concrete, with particle size of 0.01–2 cm (40% < 0.02 cm). The chemical and physical characteristics are shown in (Table 1). The P:C substrate generally had a higher cation exchange capacity, as well as carbon, nitrogen and phosphorous content, which was conversely very low in P:A. The presence of the demolition aggregates in the substrates P:A and P:C:A markedly increased the content of the heavy metals Cr and Ni.
Cultivation was performed in open air. Water was provided when needed (pH 7.4, conductivity 505 µS cm−1 at 20 °C), while fertilizer (Peters© Professional Allrounder 20-20-20, Scotts Professional, Geldermalsen, The Netherlands) was applied three times during spring and once during autumn. Cultivation lasted until the summer of 2016, when survival percentage, flowering and morphology were evaluated and volatile organic compound (VOC) and EO profiles were analyzed. The weather parameters were monitored using the closest weather station (Latitude N 450447, Longitude E 073639; WGS84). During the blooming period (June–August 2016), the average maximum temperature was 29.6 °C (34.9 °C was the highest), the average minimum temperature was 18 °C (12.5 °C was the lowest) and the average solar radiation was 23.29 MJ m−2 per day.

2.2. Analysis of Biometric Parameters and Performance

During the first flowering season (Summer 2015), only 8.9% of plants bloomed; thus, the biometric parameters were recorded in the second flowering season (Summer 2016) among the surviving plants (110 plants in P:C, 92 in P:A, and 71 in P:C:A), from the beginning of June until mid-August. The percentages of flowering plants in each substrate per week were recorded. Concurrently, morphological characteristics of each flowering plant were evaluated according to selected guidelines of the International Union for the Protection of New Varieties of Plants (UPOV) proposed for lavenders. The parameters were number of spikes per plant (n), spike length (cm), plant height (cm) and plant diameter (cm). Afterwards, the flowered spikes of each plant were cut, air-dried and weighed and used for the analyses of emissions profiles and essential oils.

2.3. Analysis of Secondary Metabolites

Spontaneous emission profiles and essential oil compositions of the lavender plants were analyzed using methods described in a previous study [36]. Emitted volatiles were sampled from the headspace of each plant with a Supelco (Bellefonte, PA, USA) solid phase microextraction (SPME) device (Supelcor, Bellafonte, PA, USA) coated with polydimethylsiloxane (PDMS, 100 μm coating thickness, St. Louis, MO, USA). They were then injected into a Varian CP-3800 apparatus (Varian Inc., Palo Alto, Santa Clara, CA, USA) coupled to a Varian Saturn 2000 (Varian Inc., Palo Alto, Santa Clara, CA, USA) for the gas chromatography–flame ionization detection (GC-FID) and gas chromatography-mass spectrometry (GC-MS) analyses. Essential oils were obtained by the hydro-distillation of dried lavender inflorescences using a Clevenger-type apparatus (Tecnovetro, Milan, Italy), and those oils were then injected into the GC-FID and GC-MS devices for the identification of oil constituents.

2.4. Statistical Analyses

All of the data were first tested for the homogeneity of variances (Levene test). A one-way ANOVA (Analysis of Variance) was performed on biometric parameters of the three local lavender selections to test the effect of cultivation substrate and means were separated according to Tukey post-hoc tests; a two-way ANOVA was performed to analyze the interaction between the substrate and lavender selection effects. Statistically significant differences induced by substrate and lavender selection on VOC, EO compounds and EO yields were assessed with one- or two-way PERMANOVAs (Permutational Multivariate Analysis of Variance) with Euclidean distances, which were based on a distribution-free analysis of variance. The percentage contribution of each compound to the observed dissimilarity was assessed through similarity percentage analysis (SIMPER, Euclidean distance). For each compound, the differences between substrates and lavender selections were tested with the Mann-Whitney pairwise test. The value for statistical significance was p < 0.05. All statistical analyses were performed by Past 3 software, version 3.15.

3. Results and Discussion

3.1. Influence of Cultivation Substrate on Morphology and Performance

Lavandula angustifolia plants performed differently after a year and a half under cultivation in the various substrates, which influenced all the measured parameters except for the spike length (Table 2). The best results were obtained in the P:C substrate, where more plants survived (61.1%), had a higher spike number (9.0) and higher flower yield (3.7 g of dried flowers per plant). In the P:A substrate, plants flowered less (6.3 spikes and 1.4 g of dried flowers per plant) and P:C:A showed lower plant survival (39.4%) and compactness (height = 45.3 cm and diameter = 15 cm). The lavender selection factor interacted with the substrate factor only for the survival rate, which ranged from 45.4% for Susa, 50.2% for Stura, and 57.6% for Tanaro lavender selections
The addition of demolition aggregates to peat and compost raised the substrate pH in P:A and P:C:A when compared to P:C, similarly to what was observed by [45] after mixing crushed bricks to compost-based pot media for the cultivation of parsley (Petroselinum crispum (Mill.) Fuss) and coriander (Coriandrum sativum L.). Lavender usually thrives best in neutral to alkaline soils [2]; however, plants grown in P:A (pH 7) and P:C:A (pH 8.2) showed low performances, especially concerning flower production and survival rate in P:C:A, thus adapting better to acidic media. Adaptation of plants to pH ranges different from species optima has been previously observed in other ornamental plants [46,47]. P:C had generally higher nutrient levels compared to P:A and P:C:A; nonetheless, conflicting results on how mineral nutrients affect Lamiaceae morphology and performance are reported in the literature [29,43,45,48,49,50]. Lavenders were generally scarce after cultivation in P:C:A, with the lowest survival rate and spread plants, even if peat, compost and mineral material mixtures have been reported to support an adequate establishment of lavender [42] and other Lamiaceae species, namely L. dentata L., Satureja montana L., Thymus pseudolanuginosus Ronniger and T. caespititius Brot. [51,52]. Both substrates with demolition aggregates had very high heavy metal concentrations (Cr and Ni), which are known to reduce plant growth and biomass, as reviewed by [53]. References [54,55] have not recorded any negative influence of heavy metals on peppermint (Mentha × piperita L.), basil and lavender, but two of the studied heavy metals, Cr and Ni, exceeded in P:C:A the accepted limits in soils according to Italian legislation (Cr: 150 mg kg−1 and Ni: 120 mg kg−1; D. Lgs. 152/2006–Norme in materia ambientale–G.U.88), possibly contributing to the limited lavender survival.
The flowering period (Figure 1) lasted 10 weeks (from 6 June 2016 to 19 August 2016). Every potted plant bloomed, and almost 90% of lavenders bloomed within the first four weeks. Plants cultivated on P:C and P:C:A had peak blooming during the second week, with 37% and 30% of plants flowering, respectively, whereas in P:A, the blooming was more gradual and delayed by one or two weeks. However, no significant differences were recorded between the different substrates.

3.2. Influence of Cultivation Substrate on Vocs

Volatile compounds analyzed by the SPME technique are reported in Table 3. The most prevalent class of compounds was oxygenated monoterpenes (OMs), with percentages ranging from 68.8% (in P:A-Sus) to 84.9% (in P:A-Tan), followed by monoterpene hydrocarbons (MHs, 5.7% in P:A-Stu, 15.5% in P:C-Sus). This study is in agreement with previous papers on the same species [36,56]. It is important to emphasize the highest percentage of non-terpene derivatives (NTs) was found in Sus samples grown in P:A substrate (13.4%). The percentage of sesquiterpenoids (SHs) in the aroma profile was very low (from 1.7 in P:C-Stu to 3.5% in P:C:A-Stu). Interestingly, all samples were oxygenated sesquiterpene (OS) free, except for P:C-Stu and P:A-Sus where caryophyllene oxide was the unique compound. Da Prorto and Decorti (2008) [57] noted the absence of OS when analyzing the volatile compounds of lavender flowers cultivated in the north east regions of Italy, while a considerable amount of SH was measured (7% of the whole composition).
Seventy-five VOCs were identified in all the analyzed samples, with a percentage of identification ranging from 96.5–99.9%. The number of compounds in each sample varied from 37 (P:A-Sus) to 49 (P:C-Tan). Among them, only 25 were common in all of the lavender profiles. Linalyl acetate (accounting at least 20.6% of the total composition) and linalool (ranging from 18.7% in P:A-Sus to 42.7% in P:C-Tan) were identified as the main compounds of the aroma emission, with values higher than the amounts recorded the previous year in the same lavender selections [36]. The amount of linalyl acetate is in accordance with the content in the aroma of several Bulgarian lavender varieties [58], while linalool reached higher values in this study. Both isomers trans- and cis-linalool oxide (furanoids) were present in good amounts and the highest percentages were measured in P:C-Stu (8.3 and 7.1%, respectively). All these latter compounds accounted for at least 50% of the total identified and detected, up to 75.3% in P:A-Tan.
Volatile organic compounds can vary drastically depending on several other factors [59]. Among these, abiotic and biotic stresses have affected the emission rate of the hydrophilic oxygenated monoterpene 1,8-cineole in Eucalyptus globulus Labill., through the regulation of terpene synthase and stomatal conductance [60]. The emission rates of the same compound were found to be also light- and temperature-dependent in Pinus sylvestris L. [61], Hesperis matronalis L. [62] and Pinus pinea L. [63]. In medicinal and aromatic plants under moderately water-stressed conditions, volatile oxygenated and monoterpene hydrocarbons increased in Salvia sinaloensis Fern. and Helichrysum petiolare Hilliard & B.L. Burtt plants [31,32]. While, under similar conditions, an increase in sesquiterpene hydrocarbons was revealed in Salvia dolomitica Codd. Plants [33]. In lavenders, factors such as latitude [36], stage of flower development [64] and genotype [6,57,58,65,66] may affect the volatile composition. However, in this study, the two-way PERMANOVA (using 9999 permutations) performed on VOCs revealed that the cultivation substrate did not significantly affect the volatilome of lavender (F = 0.69, p = 0.2156, Table 4).

3.3. Influence of Cultivation Substrate on EO Yield and Composition

Essential oil yields varied from very low in P:C:A-Sus to a maximum of 1% in P:A-Sus (Table 5). Reference [67] showed that the average yield of lavender EO was 0.14% (w/w), although sometimes it can reach 5% [68]. The cultivation substrate influenced the EO yields according to the one-way PERMANOVA analysis (F = 3.225, p = 0.047), with yields obtained from plants in the P:C substrates significantly lower than yields obtained from plants in P:C:A substrates. The production of secondary metabolites, including EOs, is regulated by genetics and edaphic factors [69], water stress [31,32,33] and light, as well [70]. Differences in EO yield caused by substrates have been observed in other Lamiaceae species, such as Thymus caespititius Brot. [69] and Ocimum basilicum [39]. Particularly in the latter species, the yields of EOs increased up to 40% with the addition of 20–60% mineral material to peat. Likewise, a higher EO yield was obtained in this study from lavenders grown in both substrates containing mineral material (P:A). Contrasting results have been achieved concerning the effect of mineral nutrients on EO production in Lamiaceae species. Elevated levels of potassium can decrease the EO content in many plants [40], as highlighted in Origanum dictamnus [49] and lavender [30] with >300 mg L−1 in hydroponics; however, no evident relations between potassium and yields were found in Rosmarinus officinalis [48]. Lavender EO production was not influenced by 30 to 70 mg L−1 of available phosphorus, according to [29], in contrast with another study on lavender, where higher amounts of oil were produced with increased P applications in field trial [50]. Essential oil synthesis can be also fostered by stress conditions [70]. Plants in P:C:A substrate indeed had a poor performance; however, they produced higher amounts of EOs.
The main class of compounds in EOs was OM (Table 5), which varied from 69.6% in P:A-Stu to 81.8% in P:C:A-Tan. OS followed, with 5.9% in P:C:A-Tan and 13.7% in P:C-Stu. The EOs included apocarotenoids (AC), represented only by hexahydro farnesyl acetone in all samples, except for P:C:A-Tan. MHs, a very important class in the aroma profile, were present in all EOs with varying percentages (2.8% in P:C-Tan and 5.6% in P:C:A-Tan). It is important to highlight that the number of NT compounds decreased by 38.1% (from 21 to 13 constituents in VOCs and EOs, respectively) in comparison with those reported in VOCs, and NT percentages were also drastically decreased (from 2.3% in P:C-Tan to 6.5% in P:A-Sus). On the other hand, the OSs, which were almost absent in VOCs, were present as a considerable amount (5.9% in P:C:A-Tan-13.7% in P:C-Stu) in EOs.
A total of 101 compounds were identified in lavender EOs (Table 5), representing more than 99.2% of the total EO compositions. [71] found only 21 different compounds in L. angustifolia EO, while in this study higher numbers of components were found, ranging from 62 in P:C:A-Sus to 86 in P:A-Stu. Only 39 compounds were the same in all of the studied samples. Linalool was the major compound, and its percentage represented at least 21.0% in P:C-Sus, followed by linalyl acetate (from 7.5% in P:A-Tan to 17.7% in P:C-Sus). The percentages of the major compounds were inverted in comparison to VOC composition. The obtained value in this work showed that the amount of linalool in Tan valley and in all substrates agreed with those reported in the European Pharmacopeia (E.Ph.), which mentioned that the percentage in linalool had to be ranged between 25 and 45% of the total composition. On the contrary, linalyl acetate was very low in comparison with the value reported by E.Ph. Taking into account the substrate, both P:C:A (36.3%, average of three values) and P:A (27.0%) pointed out a value of linalool which was accepted by both E.Ph. and AFNOR. Linalool and linalyl acetate are important compounds in the cosmetic and pharmaceutical industry for their numerous biological and therapeutic activities [72]. Usually, they are the most abundant compounds in lavender oil [4], but their proportion can vary widely within the species [2,36,37,73]. Reference [74], for instance, found that the sum of these compounds reached more than 70% of the total EO composition.
The two-way PERMANOVA performed on the EO composition showed that both factors (substrates and lavender selection) had a significant influence on the EO composition (Table 4). Sus and Tan selections differently produced EOs, as seen in a previous study [36]. In the case of cultivation substrates, the pair-wise test showed significant differences between samples grown in P:A and those grown in P:C:A (Table 4). The profile of an essential oil is complex to evaluate due to its numerous constituents, particularly in L. angustifolia, since it is the most variable in the genus [75]. Nonetheless, a few studies on Lamiaceae species have shown that oil composition varied in Thymus caespititius Brot. [69], Thymus vulgaris [43] and Origanum vulgare L. [76] according to substrate, ratio and type of compost. Moreoever, there is the evidence that heavy metal content (Cd, Cu, Pb) can alter the EO composition of basil and peppermint in a peat-based medium experiment, though without affecting the marketability of the product [55]. Similarity percentage (SIMPER) analysis was performed to determine the compounds that contribute to the differences among the substrates. The results (Table 6) indicate that linalool and linalyl acetate together were responsible for 76.86%. This dissimilarity reached 95.06% when adding the effect of 1,8-cineol, caryophyllene oxide, lavandulyl acetate, 4-terpineol, borneol, 1-octan-3-yl acetate, τ-cadinol, α-terpineol, isopulegol acetate and (E)-caryophyllene to the previous compounds. Linalool, caryophyllene oxide, (E)-γ-bisabolene and (Z)-α-santalol showed a significant difference for substrate factors at 0.05 criterions. Caryophyllene oxide, (E)-γ-bisabolene and (Z)-α-santalol were higher in plants grown on P:C substrates, while plants grown on P:C:A substrates had higher recorded amounts of linalool. This latter compound has positive effects on the central nervous system, thus important for medicinal purposes.

4. Conclusions

An overview on the influence of different cultivation substrates on L. angustifolia morphology as well as VOC and EO profiles was provided in this study. All morphological parameters evaluated were affected by substrate composition, except for spike length. Generally, plants performed better in terms of survival rate, compactness, number of spikes and flower yield when cultivated in the substrate with peat and compost, being the best out of the three substrates tested from an ornamental horticulture perspective. The VOC profile after cultivation in different substrates did not change, and it had never been studied before in lavender. Interestingly, the highest EO yield and amounts of linalool were obtained by cultivating plants in the mixture of peat, compost and demolition aggregates, even though plants had a lower survival rate. Thus, locally sourced materials, such as green compost or demolition aggregates, can be effectively used in the preparation of pot mixtures. In the cultivation of lavender, substrate composition can be regulated depending on the final use of the plant, whether as an ornamental or for cosmetic, industrial and medicinal purposes. This will help the environmental protection by reducing waste material and supporting recycling, together with reducing the use of peat in horticulture.

Author Contributions

B.N. and S.D. contributed equally to this work. Conceptualization, V.S.; formal analysis B.N. and S.D.; investigation B.N., S.D., M.C., W.G. and P.L.C.; resources, V.S. and L.P.; data curation, B.N. and S.D.; writing—original draft preparation, B.N. and S.D.; writing—review and editing, V.S. and L.P.; visualization, B.N. and S.D.; supervision, V.S. and L.P.; project administration, V.S.; funding acquisition, V.S.

Funding

This research was partially funded by Fondazione Cassa di Risparmio di Torino, grant number 2014.0976 and by the program Interreg V-A Francia Italia Alcotra, n. 1139.

Acknowledgments

The authors thank Paolo Lo Turco for his contribution to the propagation of lavender plants, Fratelli Gramaglia nursery for hosting cultivation, ItalConcimi S.r.l for providing peat and green compost and Perino Piero and C. S.n.c. for providing demolition aggregates.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Prusinowska, R.; Śmigielski, K.B. Composition, biological properties and therapeutic effects of lavender (Lavandula angustifolia L): A review. Herba Pol. 2014, 60, 56–66. [Google Scholar] [CrossRef]
  2. Lim, T.K. Lavandula angustifolia. In Edible Medicinal and Non Medicinal Plants: Volume 8, Flowers; Springer: Dordrecht, The Netherlands, 2014; pp. 156–185. ISBN 978-94-017-8748-2. [Google Scholar]
  3. Lesage-Meessen, L.; Bou, M.; Sigoillot, J.-C.; Faulds, C.B.; Lomascolo, A. Essential oils and distilled straws of lavender and lavandin: A review of current use and potential application in white biotechnology. Appl. Microbiol. Biotechnol. 2015, 99, 3375–3385. [Google Scholar] [CrossRef] [PubMed]
  4. Cavanagh, H.M.A.; Wilkinson, J.M. Lavender essential oil: A review. Aust. Infect. Control 2005, 10, 35–37. [Google Scholar] [CrossRef]
  5. Soltani, R.; Soheilipour, S.; Hajhashemi, V.; Asghari, G.; Bagheri, M.; Molavi, M. Evaluation of the effect of aromatherapy with lavender essential oil on post-tonsillectomy pain in pediatric patients: A randomized controlled trial. Int. J. Pediatr. Otorhinolaryngol. 2013, 77, 1579–1581. [Google Scholar] [CrossRef] [PubMed]
  6. Palá-Paúl, J.; Brophy, J.J.; Goldsack, R.J.; Fontaniella, B. Analysis of the volatile components of Lavandula canariensis (L.) Mill., a Canary Islands endemic species, growing in Australia. Biochem. Syst. Ecol. 2004, 32, 55–62. [Google Scholar] [CrossRef]
  7. Hassiotis, C.N.; Lazari, D.M.; Vlachonasios, K.E. The effects of habitat type and diurnal harvest on essential oil yield and composition of Lavandula angustifolia Mill. Fresenius Environ. Bull. 2010, 19, 1491–1498. [Google Scholar]
  8. Barrett, G.E.; Alexander, P.D.; Robinson, J.S.; Bragg, N.C. Achieving environmentally sustainable growing media for soilless plant cultivation systems—A review. Sci. Hortic. (Amsterdam) 2016, 212, 220–234. [Google Scholar] [CrossRef]
  9. Kern, J.; Tammeorg, P.; Shanskiy, M.; Sakrabani, R.; Knicker, H.; Kammann, C.; Tuhkanen, E.-M.; Smidt, G.; Prasad, M.; Tiilikkala, K.; et al. Synergistic use of peat and charred material in growing media–an option to reduce the pressure on peatlands? J. Environ. Eng. Landsc. Manag. 2017, 25, 160–174. [Google Scholar] [CrossRef]
  10. Fascella, G. Growing substrates alternative to peat for ornamental plants. In Soilless Culture–Use of Substrates for the Production of Quality Horticultural Crops; Asaduzzaman, M., Ed.; InTech: London, UK, 2015; pp. 47–67. [Google Scholar]
  11. Balliu, A.; Sallaku, G.; Nasto, T. Nursery management practices influence the quality of vegetable seedlings. Italus Hortus 2017, 24, 39–52. [Google Scholar]
  12. Bilderback, T.E.; Warren, S.L.; Owen, J.S.; Albano, J.P. Healthy substrates need physicals too! Horttechnology 2005, 15, 747–751. [Google Scholar] [CrossRef]
  13. Berruti, A.; Scariot, V. Efficacy of flurprimidol and peat alternatives on growth control of potted camellias. New Zeal. J. Crop Hortic. Sci. 2013, 41, 230–239. [Google Scholar] [CrossRef]
  14. Ganesh, S.; Kannan, M.; Lal, M.J.; Arulmozhiyan, R.; Jeyakumar, P. Standardization of growing medium for cut chrysanthemum (Dendranthema grandiflora Tzvelev) cv. Amalfi under protected conditions. J. Ornam. Hortic. 2015, 18, 48–55. [Google Scholar]
  15. Pagani, A.; Molinari, J.; Lavado, R.; Di Benedetto, A. Behavior of Impatiens wallerana Hook. F in alternative pot substrates: Mechanisms involved and research perspectives. J. Plant Nutr. 2015, 38, 2185–2203. [Google Scholar] [CrossRef]
  16. Hernández-Apaolaza, L.; Gascó, A.M.; Gascó, J.M.; Guerrero, F. Reuse of waste materials as growing media for ornamental plants. Bioresour. Technol. 2005, 96, 125–131. [Google Scholar] [CrossRef] [PubMed]
  17. Jayasinghe, G.Y.; Arachchi, I.D.D.L.L.; Tokashiki, Y. Evaluation of containerized substrates developed from cattle manure compost and synthetic aggregates for ornamental plant production as a peat alternative. Resour. Conserv. Recycl. 2010, 54, 1412–1418. [Google Scholar] [CrossRef]
  18. Khomami, A.M.; Guilan, I. The possibility using the composted peanut shells in the growth of Marigold and Viola tricolor plants. J. Ornam. Plants 2015, 5, 61–66. [Google Scholar]
  19. Gong, X.; Li, S.; Sun, X.; Wang, L.; Cai, L.; Zhang, J.; Wei, L. Green waste compost and vermicompost as peat substitutes in growing media for geranium (Pelargonium zonale L.) and calendula (Calendula officinalis L.). Sci. Hortic. (Amsterdam) 2018, 236, 186–191. [Google Scholar] [CrossRef]
  20. Massa, D.; Malorgio, F.; Lazzereschi, S.; Carmassi, G.; Prisa, D.; Burchi, G. Evaluation of two green composts for peat substitution in geranium (Pelargonium zonale L.) cultivation: Effect on plant growth, quality, nutrition, and photosynthesis. Sci. Hortic. (Amsterdam) 2018, 228, 213–221. [Google Scholar] [CrossRef]
  21. Guo, Y.; Niu, G.; Starman, T.; Gu, M. Growth and development of Easter lily in response to container substrate with biochar. J. Hortic. Sci. Biotechnol. 2019, 94, 80–86. [Google Scholar] [CrossRef]
  22. Zhong, Z.; Bian, F.; Zhang, X. Testing composted bamboo residues with and without added effective microorganisms as a renewable alternative to peat in horticultural production. Ind. Crops Prod. 2018, 112, 602–607. [Google Scholar] [CrossRef]
  23. Di Lorenzo, R.; Pisciotta, A.; Santamaria, P.; Scariot, V. From soil to soil-less in horticulture: Quality and typicity. Ital. J. Agron. 2013, 8, 255–260. [Google Scholar] [CrossRef]
  24. Ceglie, F.G.; Bustamante, M.A.; Ben Amara, M.; Tittarelli, F. The challenge of peat substitution in organic seedling production: Optimization of growing media formulation through mixture design and response surface analysis. PLoS ONE 2015, 10, e0128600. [Google Scholar] [CrossRef] [PubMed]
  25. Zubek, S.; Stefanowicz, A.M.; Błaszkowski, J.; Niklińska, M.; Seidler-Łozykowska, K. Arbuscular mycorrhizal fungi and soil microbial communities under contrasting fertilization of three medicinal plants. Appl. Soil Ecol. 2012, 59, 106–115. [Google Scholar] [CrossRef]
  26. Bufalo, J.; Cantrell, C.L.; Astatkie, T.; Zheljazkov, V.D.; Gawde, A.; Boaro, C.S.F. Organic versus conventional fertilization effects on sweet basil (Ocimum basilicum L.) growth in a greenhouse system. Ind. Crops Prod. 2015, 74, 249–254. [Google Scholar] [CrossRef]
  27. Najafian, S.; Zahedifar, M. Antioxidant activity and essential oil composition of Satureja hortensis L. as influenced by sulfur fertilizer. J. Sci. Food Agric. 2015, 95, 2404–2408. [Google Scholar] [CrossRef] [PubMed]
  28. Egamberdieva, D.; Shrivastava, S.; Varma, A. Plant-Growth-Promoting Rhizobacteria (PGPR) and Medicinal Plants; Egamberdieva, D., Shrivastava, S., Varma, A., Eds.; Soil Biology; Springer International Publishing: Cham, Switzerland, 2015; Volume 42, ISBN 978-3-319-13400-0. [Google Scholar]
  29. Chrysargyris, A.; Panayiotou, C.; Tzortzakis, N. Nitrogen and phosphorus levels affected plant growth, essential oil composition and antioxidant status of lavender plant (Lavandula angustifolia Mill.). Ind. Crops Prod. 2016, 83, 577–586. [Google Scholar] [CrossRef]
  30. Chrysargyris, A.; Drouza, C.; Tzortzakis, N. Optimization of potassium fertilization/nutrition for growth, physiological development, essential oil composition and antioxidant activity of Lavandula angustifolia Mill. J. Soil Sci. Plant Nutr. 2017, 17, 291–306. [Google Scholar] [CrossRef]
  31. Caser, M.; D’Angiolillo, F.; Chitarra, W.; Lovisolo, C.; Ruffoni, B.; Pistelli, L.L.; Pistelli, L.L.; Scariot, V. Water deficit regimes trigger changes in valuable physiological and phytochemical parameters in Helichrysum petiolare Hilliard & BL Burtt. Ind. Crops Prod. 2016, 83, 680–692. [Google Scholar]
  32. Caser, M.; D’Angiolillo, F.; Chitarra, W.; Lovisolo, C.; Ruffoni, B.; Pistelli, L.L.; Pistelli, L.L.; Scariot, V. Ecophysiological and phytochemical responses of Salvia sinaloensis Fern. to drought stress. Plant Growth Regul. 2018, 84, 383–394. [Google Scholar] [CrossRef]
  33. Caser, M.; Chitarra, W.; D’Angiolillo, F.; Perrone, I.; Demasi, S.; Lovisolo, C.; Pistelli, L.; Pistelli, L.; Scariot, V. Drought stress adaptation modulates plant secondary metabolite production in Salvia dolomitica Codd. Ind. Crops Prod. 2019, 129, 85–96. [Google Scholar] [CrossRef]
  34. Kleinwachter, M.; Selmar, D. New insights explain that drought stress enhances the quality of spice and medicinal plants: Potential applications. Agron. Sustain. Dev. 2014, 35, 121–131. [Google Scholar] [CrossRef]
  35. Chu, C.J.; Kemper, K.J. Lavender (Lavandula spp.) 2001.
  36. Demasi, S.; Caser, M.; Lonati, M.; Cioni, P.L.; Pistelli, L.; Najar, B.; Scariot, V. Latitude and altitude influence secondary metabolite production in peripheral alpine populations of the mediterranean species Lavandula angustifolia Mill. Front. Plant Sci. 2018, 9, 983. [Google Scholar] [CrossRef] [PubMed]
  37. Hassiotis, C.N.; Ntana, F.; Lazari, D.M.; Poulios, S.; Vlachonasios, K.E. Environmental and developmental factors affect essential oil production and quality of Lavandula angustifolia during flowering period. Ind. Crops Prod. 2014, 62, 359–366. [Google Scholar] [CrossRef]
  38. Usano-Alemany, J.; Palá-Paúl, J.; Rodríguez, M.S.-C.; Herraiz-Peñalver, D. Chemical description and essential oil yield variability of different accessions of Salvia lavandulifolia. Nat. Prod. Commun. 2014, 9. [Google Scholar] [CrossRef]
  39. Burdina, I.; Priss, O. Effect of the substrate composition on yield and quality of basil (Ocimum basilicum L.). J. Hortic. Res. 2016, 24, 109–118. [Google Scholar] [CrossRef]
  40. Sangwan, N.S.; Farooqi, A.H.A.; Shabih, F.; Sangwan, R.S. Regulation of essential oil production in plants. Plant Growth Regul. 2001, 34, 3–21. [Google Scholar] [CrossRef]
  41. Usano-Alemany, J.; Palá-Paúl, J.; Herráiz-Peñalver, D. Temperature stress causes different profiles of volatile compounds in two chemotypes of Salvia lavandulifolia Vahl. Biochem. Syst. Ecol. 2014, 54, 166–171. [Google Scholar] [CrossRef]
  42. Kotsiris, G.; Nektarios, P.A.; Paraskevopoulou, A.T. Lavandula angustifolia growth and physiology is affected by substrate type and depth when grown under mediterranean semi-intensive green roof conditions. HortScience 2012, 47, 311–317. [Google Scholar] [CrossRef]
  43. Bolechowski, A.; Moral, R.; Bustamante, M.A.; Bartual, J.; Paredes, C.; Pérez-Murcia, M.D.; Carbonell-Barrachina, A.A. Winery-distillery composts as partial substitutes of traditional growing media: Effect on the volatile composition of thyme essential oils. Sci. Hortic. (Amsterdam) 2015, 193, 69–76. [Google Scholar] [CrossRef]
  44. Boyle, T.H.; Craker, L.E.; Simon, J.E. Growing medium and fertilization regime influence growth and essential oil content of rosemary. Hortscience 1991, 26, 33–34. [Google Scholar] [CrossRef]
  45. Sousa, G.; Monteiro, F.G.; Vasconcelos, E.; Ribeiro, H.M. Valorization of sieved crushed bricks as a component of compost-based substrates. Acta Hortic. 2017, 1168, 303–310. [Google Scholar] [CrossRef]
  46. Demasi, S.; Caser, M.; Handa, T.; Kobayashi, N.; De Pascale, S.; Scariot, V. Adaptation to iron deficiency and high pH in evergreen azaleas (Rhododendron spp.): Potential resources for breeding. Euphytica 2017, 213, 148. [Google Scholar] [CrossRef]
  47. Demasi, S.; Caser, M.; Kobayashi, N.; Kurashige, Y.; Scariot, V. Hydroponic screening for iron deficiency tolerance in evergreen azaleas. Not. Bot. Horti Agrobot. Cluj-Napoca 2015, 43, 210–213. [Google Scholar] [CrossRef]
  48. Puttanna, K.; Rao, E.V.S.P.; Singh, R.; Ramesh, S. Influence of Nitrogen and Potassium Fertilization on Yield and Quality of Rosemary in Relation to Harvest Number. Commun. Soil Sci. Plant Anal. 2010, 41, 190–198. [Google Scholar] [CrossRef]
  49. Economakis, C.D. Effect of potassium on growth and yield of Origanum dictamnus L. in solution culture. Acta Hortic. 1993, 339–344. [Google Scholar] [CrossRef]
  50. Șekeroğlu, N.; Özgüven, M. Determination of optimum phosphorus doses for high flower yield and essential oil content in common Lavender (Lavandula angustifolia Mill.). In Proceedings of the Fifth Conference on Medicinal and Aromatic Plants of Southeast European Countries, (5th CMAPSEEC), Brno, Czech Republic, 2–5 September 2008. [Google Scholar]
  51. Monteiro, C.M.; Calheiros, C.S.C.; Martins, J.P.; Costa, F.M.; Palha, P.; de Freitas, S.; Ramos, N.M.M.; Castro, P.M.L. Substrate influence on aromatic plant growth in extensive green roofs in a Mediterranean climate. Urban Ecosyst. 2017, 20, 1347–1357. [Google Scholar] [CrossRef]
  52. Monteiro, C.M.; Calheiros, C.S.C.; Palha, P.; Castro, P.M.L. Growing substrates for aromatic plant species in green roofs and water runoff quality: Pilot experiments in a Mediterranean climate. Water Sci. Technol. 2017, 76, 1081–1089. [Google Scholar] [CrossRef]
  53. Nagajyoti, P.C.; Lee, K.D.; Sreekanth, T.V.M. Heavy metals, occurrence and toxicity for plants: A review. Environ. Chem. Lett. 2010, 8, 199–216. [Google Scholar] [CrossRef]
  54. Zheljazkov, V.D.; Nielsen, N.E. Studies on the effect of heavy metals (Cd, Pb, Cu, Mn, Zn and Fe) upon the growth, productivity and quality of lavender (Lavandula angustifolia Mill.) production. J. Essent. Oil Res. 1996, 8, 259–274. [Google Scholar] [CrossRef]
  55. Zheljazkov, V.D.; Craker, L.E.; Xing, B. Effects of Cd, Pb, and Cu on growth and essential oil contents in dill, peppermint, and basil. Environ. Exp. Bot. 2006, 58, 9–16. [Google Scholar] [CrossRef]
  56. Pistelli, L.; Najar, B.; Giovanelli, S.; Lorenzini, L.; Tavarini, S.; Angelini, L.G. Agronomic and phytochemical evaluation of lavandin and lavender cultivars cultivated in the Tyrrhenian area of Tuscany (Italy). Ind. Crops Prod. 2017, 109, 37–44. [Google Scholar] [CrossRef]
  57. Da Porto, C.; Decorti, D. Analysis of the volatile compounds of flowers and essential oils from Lavandula angustifolia cultivated in northeastern Italy by headspace solid-phase microextraction coupled to gas chromatography-mass spectrometry. Planta Med. 2008, 74, 182–187. [Google Scholar] [CrossRef] [PubMed]
  58. Milina, R.; Mustafa, Z.; Stanev, S.; Zvezdova, D.; Stoeva, S. Headspace gas chromatographic analysis of Bulgarian Lavandula Angustifolia mill Herbs. I.optimization of the analysis conditions. Научни Трудoве На Русенския Университет (Sci. Works Univ. (Bulgarian) 2012, 51, 50–56. [Google Scholar]
  59. Holopainen, J.K.; Gershenzon, J. Multiple stress factors and the emission of plant VOCs. Trends Plant Sci. 2010, 15, 176–184. [Google Scholar] [CrossRef] [PubMed]
  60. Kanagendran, A.; Pazouki, L.; Bichele, R.; Külheim, C.; Niinemets, Ü. Temporal regulation of terpene synthase gene expression in Eucalyptus globulus leaves upon ozone and wounding stresses: Relationships with stomatal ozone uptake and emission responses. Environ. Exp. Bot. 2018, 155, 552–565. [Google Scholar] [CrossRef] [PubMed]
  61. Tarvainen, V.; Hakola, H.; Hellén, H.; Bäck, J.; Hari, P.; Kulmala, M. Temperature and light dependence of the VOC emissions of Scots pine. Atmos. Chem. Phys. Discuss. 2004, 4, 6691–6718. [Google Scholar] [CrossRef]
  62. Nielsen, J.K.; Jakobsen, H.B.; Friis, P.; Hansen, K.; Møller, J.; Olsen, C.E. Asynchronous rhythms in the emission of volatiles from Hesperis matronalis flowers. Phytochemistry 1995, 38, 847–851. [Google Scholar] [CrossRef]
  63. Staudt, M.; Bertin, N.; Hansen, U.; Seufert, G.; Ciccioli, P.; Foster, P.; Frenzel, B.; Fugit, J.L. Seasonal and diurnal patterns of monoterpene emissions from Pinus pinea (L.) under field conditions. Atmos. Environ. 1997, 31, 145–156. [Google Scholar] [CrossRef]
  64. Woronuk, G.; Demissie, Z.; Rheault, M.; Mahmoud, S. Biosynthesis and therapeutic properties of lavandula essential oil constituents. Planta Med. 2011, 77, 7–15. [Google Scholar] [CrossRef]
  65. Sanz, J.; Soria, A.C.; Garcıa-Vallejo, M.C. Analysis of volatile components of Lavandula luisieri L. by direct thermal desorption–gas chromatography–mass spectrometry. J. Chromatogr. A 2004, 1024, 139–146. [Google Scholar] [CrossRef]
  66. An, M.; Haig, T.; Hatfield, P. On-site field sampling and analysis of fragrance from living Lavender (Lavandula angustifolia L.) flowers by solid-phase microextraction coupled to gas chromatography and ion-trap mass spectrometry. J. Chromatogr. A 2001, 917, 245–250. [Google Scholar] [CrossRef]
  67. Cardia, G.F.E.; Silva-Filho, S.E.; Silva, E.L.; Uchida, N.S.; Cavalcante, H.A.O.; Cassarotti, L.L.; Salvadego, V.E.C.; Spironello, R.A.; Bersani-Amado, C.A.; Cuman, R.K.N. Effect of lavender (Lavandula angustifolia) essential oil on acute inflammatory response. Evid.-Based Complement. Altern. Med. 2018, 2018, 1–10. [Google Scholar] [CrossRef] [PubMed]
  68. Al-Younis, F.; Al-Naser, Z.; Al-Hakim, W. Chemical composition of lavandula angustifolia miller and rosmarinus officinalis L. Essential oils and fumigant toxicity against larvae of ephestia kuehniella zeller. Int. J. ChemTech Res. 2015, 8, 1382–1390. [Google Scholar]
  69. Pereira, S.I.; Santos, P.A.G.; Barroso, J.G.; Figueiredo, A.C.; Pedro, L.G.; Salgueiro, L.R.; Deans, S.G.; Scheffer, J.J.C. Chemical polymorphism of the essential oils from populations of Thymus caespititius grown on the island S. Jorge (Azores). Phytochemistry 2000, 55, 241–246. [Google Scholar] [CrossRef] [Green Version]
  70. Mandoulakani, B.A.; Eyvazpour, E.; Ghadimzadeh, M. The effect of drought stress on the expression of key genes involved in the biosynthesis of phenylpropanoids and essential oil components in basil (Ocimum basilicum L.). Phytochemistry 2017, 139, 1–7. [Google Scholar] [CrossRef] [PubMed]
  71. Tomescu, A.; Rus, C.; Pop, G.; Alexa, E.; Șumălan, R.; Copolovici, D.; Negrea, M. Chemical composition of Lavandula angustifolia L. and Rosmarinus officinalis L. essential oils cultivated in west Romania. Res. J. Agric. Sci. 2015, 47, 246–253. [Google Scholar]
  72. Cavanagh, H.M.A.; Wilkinson, J.M. Biological activities of Lavender essential oil. Phyther. Res. 2002, 16, 301–308. [Google Scholar] [CrossRef] [PubMed]
  73. Venskutonis, P.R.; Dapkevicius, A.; Baranauskiene, M. Composition of the essential oil of Lavender (Lavandula angustifolia Mill.) from Lithuania. J. Essent. Oil Res. 1997, 9, 107–110. [Google Scholar] [CrossRef]
  74. Moon, T.; Cavanagh, H.M.A.; Wilkinson, J.M. Antifungal activity of australian grown Lavandula spp. essential oils against Aspergillus nidulans, Trichophyton mentagrophytes, Leptosphaeria maculans and Sclerotinia sclerotiorum. J. Essent. Oil Res. 2007, 19, 171–175. [Google Scholar] [CrossRef]
  75. Lis-Balchin, M. Lavender: The Genus Lavandula, 1st ed.; CRC Press: London, UK, 2005; Volume 12, ISBN 9780415284868. [Google Scholar]
  76. Bolechowski, A.; Moral, R.; Bustamante, M.A.; Paredes, C.; Agulló, E.; Bartual, J.; Carbonell-Barrachina, Á.A. Composition of oregano essential oil (origanum vulgare) as affected by the use of winery-distillery composts. J. Essent. Oil Res. 2011, 23, 32–38. [Google Scholar] [CrossRef]
Figure 1. Blooming trend per week of Lavandula angustifolia plants after two cultivation cycles in different substrates (P:C = peat and green compost, 70%:30% v/v; P:A = peat and demolition aggregates, 70%:30% v/v; P:C:A = peat, green compost and demolition aggregates, 40%:30%:30% v/v).
Figure 1. Blooming trend per week of Lavandula angustifolia plants after two cultivation cycles in different substrates (P:C = peat and green compost, 70%:30% v/v; P:A = peat and demolition aggregates, 70%:30% v/v; P:C:A = peat, green compost and demolition aggregates, 40%:30%:30% v/v).
Agronomy 09 00411 g001
Table 1. Physical properties, chemical properties and heavy metal content of raw materials (P = peat; C = compost; A = demolition aggregates) and mixtures tested as cultivation substrates (P:C = peat and green compost, 70%:30% v/v; P:A = peat and demolition aggregates, 70%:30% v/v; P:C:A = peat, green compost and demolition aggregates, 40%:30%:30% v/v) with standard deviation.
Table 1. Physical properties, chemical properties and heavy metal content of raw materials (P = peat; C = compost; A = demolition aggregates) and mixtures tested as cultivation substrates (P:C = peat and green compost, 70%:30% v/v; P:A = peat and demolition aggregates, 70%:30% v/v; P:C:A = peat, green compost and demolition aggregates, 40%:30%:30% v/v) with standard deviation.
ParameterUnitsPCAP:CP:AP:C:A
pH 1 4.1 ± 0.708.0 ± 0.109.8 ± 0.205.1 ± 0.517.0 ± 0.118.2 ± 0.49
C tot 2%25.4 ± 1.4027.0 ± 0.602.0 ± 2.2041.5 ± 0.7028.1 ± 0.7016.1 ± 1.30
N tot 2%1.8 ± 0.041.9 ± 0.080.02 ± 0.001.6 ± 0.080.8 ± 0.050.6 ± 0.00
C/N 14.4 ± 0.0314.6 ± 0.01109.7 ± 0.0726.1 ± 0.0933.8 ± 0.5125.5 ± 1.00
Available P 3mg/kg18.1 ± 0.57291.0 ± 0.4011.6 ± 0.51190.7 ± 0.5214.7 ± 0.45130.9 ± 0.49
Carbonates 4%0.4 ± 0.102.2 ± 0.0513.1 ± 0.300.8 ± 0.716.6 ± 0.629.8 ± 0.10
CEC 5meq/100g108.4 ± 7.1082.4 ± 12.007.4 ± 0.0355.5 ± 0.0525.9 ± 0.0224.8 ± 0.40
Exchangeable Cameq/100g31.6 ± 2.5039.1 ± 2.8016.5 ± 0.0632.7 ± 0.0527.3 ± 0.0327.5 ± 0.03
Exchangeable Kmeq/100g0.2 ± 0.2018.7 ± 0.030.9 ± 0.064.8 ± 0.060.5 ± 0.053.9 ± 0.03
Exchangeable Mgmeq/100g5.7 ± 0.0119.2 ± 0.200.2 ± 0.017.6 ± 0.002.4 ± 0.004.2 ± 0.00
Cr 6mg/kg137.7 ± 8.70135.8 ± 1.40320.0 ± 0.04137.2 ± 0.02128.6 ± 0.11232.6 ± 0.02
Cu 6mg/kg68.0 ± 4.7069.2 ± 1.2026.5 ± 0.0123.3 ± 0.0114.5 ± 0.0029.6 ± 0.02
Ni 6mg/kg100.8 ± 5.5099.6 ± 9.50155.8 ± 0.0629.5 ± 0.0679.4 ± 0.05119.7 ± 0.02
Gravel%--40-5040
1 ISO 10390; 2 ISO 10694; 3 Olsen; 4 ISO 10693; 5 CEC: cation exchange capacity, ISO 11260; 6 EPA 3051A.
Table 2. Differences in Lavandula angustifolia survival rate, height, diameter, spike number, spike length and flower yield (grams of dry flowers per plant) after two cultivation cycles in different substrates (P:C = peat and green compost, 70%:30% v/v; P:A = peat and demolition aggregates, 70%:30% v/v; P:C:A = peat, green compost and demolition aggregates, 40%:30%:30% v/v) with standard deviation.
Table 2. Differences in Lavandula angustifolia survival rate, height, diameter, spike number, spike length and flower yield (grams of dry flowers per plant) after two cultivation cycles in different substrates (P:C = peat and green compost, 70%:30% v/v; P:A = peat and demolition aggregates, 70%:30% v/v; P:C:A = peat, green compost and demolition aggregates, 40%:30%:30% v/v) with standard deviation.
SubstrateSurvival (%)Height (cm)Diameter (cm)Spike (number)Spike Length (cm)Flower Yield (g Dry Flowers per Plant)
P:C61.1 ± 5.54 a 141.1 ± 0.77 b13.7 ± 0.42 b9.0 ± 0.52 a6.2 ± 0.203.7 ± 0.48 a
P:A51.1 ± 3.49 a43.9 ± 0.77 a12.3 ± 0.32 c6.3 ± 0.33 c6.6 ± 0.431.4 ± 2.10 b
P:C:A39.4 ± 0.44 b45.3 ± 0.75 a15.0 ± 0.41 a7.6 ± 0.44 b6.1 ± 0.191.6 ± 10.18 ab
P***********ns*
Substrate × Lavender selection p**nsnsnsnsns
1 Means followed by the same letter in the same column denote no significant differences according to Tukey test (p < 0.05). Statistical significance: * p < 0.05, ** p < 0.01, *** p < 0.001, ns, not significant.
Table 3. Solid phase microextraction (SPME) volatile profiles (%) of lavender selections from Susa (Sus), Stura (Stu) and Tanaro (Tan) valleys, grown in different substrates (P:C = peat and green compost, 70%:30% v/v; P:A = peat and demolition aggregates, 70%:30% v/v; P:C:A = peat, green compost and demolition aggregates, 40%:30%:30% v/v) with standard deviation.
Table 3. Solid phase microextraction (SPME) volatile profiles (%) of lavender selections from Susa (Sus), Stura (Stu) and Tanaro (Tan) valleys, grown in different substrates (P:C = peat and green compost, 70%:30% v/v; P:A = peat and demolition aggregates, 70%:30% v/v; P:C:A = peat, green compost and demolition aggregates, 40%:30%:30% v/v) with standard deviation.
CompoundsClass *LRI 1P:CP:AP:C:A
SusStuTanSusStuTanSusStuTan
11-octeneNT810--0.1 2 ± 0.05------
21-butyl acetate NT811-0.1 ± 0.08-0.1 ± 0.15-----
3Hexyl methyl etherNT832---0.2 ± 0.23-0.1 ± 0.10---
4α-thujeneMH932--0.2 ± 0.06--0.1 ± 0.050.1 ± 0.050.1 ± 0.050.4 ± 0.27
5tricycleneMH9380.1 ± 0.020.2 ± 0.030.7 ± 0.510.1 ± 0.100.2 ± 0.180.4 ± 0.340.2 ± 0.180.4 ± 0.170.8 ± 0.20
6β-citronellene MH946-0.1 ± 0.09----0.1 ± 0.09--
74-methyl pent-2-enolideNT9510.1 ± 0.070.2 ± 0.200.1 ± 0.06-0.4 ± 0.040.1 ± 0.080.1 ± 0.040.1 ± 0.16-
8campheneMH9550.4 ± 0.050.2 ± 0.030.3 ± 0.090.2 ± 0.090.2 ± 0.040.2 ± 0.061.0 ± 0.320.4 ± 0.030.3 ± 0.05
9sabineneMH978--0.2 ± 0.05-0.2 ± 0.07--0.1 ± 0.040.1 ± 0.08
101-octen-3-oneNT980--0.1 ± 0.13------
11β-pineneMH9810.1 ± 0.04-0.1 ± 0.080.1 ± 0.13--0.2 ± 0.090.1 ± 0.080.1 ± 0.08
121-octen-3-olNT9820.1 ± 0.090.3 ± 0.260.2 ± 0.05-0.2 ± 0.050.3 ± 0.270.4 ± 0.390.4 ± 0.090.4 ± 0.04
133-octanoneNT9872.3 ± 0.531.9 ± 0.182.7 ± 1.022.7 ± 1.251.1 ± 0.392.1 ± 1.073.3 ± 1.521.9 ± 0.361.9 ± 0.30
14butanoic acid, butyl esterNT9930.1 ± 0.060.3 ± 0.090.4 ± 0.020.4 ± 0.07 0.5 ± 0.460.2 ± 0.090.2 ± 0.030.2 ± 0.09
15myrceneMH9936.5 ± 1.253.9 ± 0.270.8 ± 0.416.4 ± 0.550.9 ± 0.140.8 ± 0.033.7 ± 0.222.4 ± 0.501.3 ± 0.48
163-octanolNT9980.1 ± 0.080.1 ± 0.070.1 ± 0.03---0.1 ± 0.70.1 ± 0.080.1 ± 0.04
17α-phellandreneMH1006--0.1 ± 0.01--0.1 ± 0.01-0.1 ± 0.100.1 ± 0.03
18δ-3-careneMH1012--0.1 ± 0.06--0.1 ± 0.09-0.1 ± 0.060.2 ± 0.23
19n-hexyl acetateNT10130.9 ± 0.100.8 ± 0.030.2 ± 0.092.1 ± 0.630.3 ± 0.010.2 ± 0.010.9 ± 0.860.6 ± 0.010.1 ± 0.08
20α-terpineneMH1019----0.1 ± 0.01----
21o-cymene MH1026----0.2 ± 0.03-0.1 ± 0.10--
22p-cymeneMH10280.2 ± 0.140.5 ± 0.040.8 ± 0.540.3 ± 0.210.7 ± 0.050.7 ± 0.600.3 ± 0.270.4 ± 0.310.9 ± 0.57
23limoneneMH10321.2 ± 0.970.5 ± 0.300.3 ± 0.08-1.0 ± 0.060.3 ± 0.12-1.7 ± 0.810.6 ± 0.41
241,8-cineoleOM10364.3 ± 2.761.3 ± 1.000.9 ± 0.417.3 ± 3.253.0 ± 0.560.7 ± 0.706.1 ± 3.580.4 ± 0.060.9 ± 1.12
25(Z)-β-ocimeneMH10422.9 ± 0.831.8 ± 1.011.6 ± 1.023.3 ± 1.341.0 ± 0.162.6 ± 0.122.9 ± 0.894.4 ± 1.343.3 ± 0.51
26lavender lactoneNT10460.5 ± 0.381.2 ± 0.760.9 ± 0.720.3 ± 0.410.6 ± 0.580.5 ± 0.460.4 ± 0.190.3 ± 0.090.5 ± 0.79
27(E)-β-ocimeneMH10534.1 ± 0.652.5 ± 1.691.8 ± 1.64.5 ± 1.231.2 ± 1.081.6 ± 0.922.5 ± 1.862.9 ± 0.053.6 ± 2.38
28γ-terpineneMH1062--0.2 ± 0.25--0.1 ± 0.18-0.2 ± 0.110.4 ± 0.37
29trans-linalool oxide (furanoid)OM10692.3 ± 1.148.3 ± 2.266.8 ± 4.51.7 ± 1.005.5 ± 3.825.8 ± 2.202.7 ± 1.222.7 ± 0.125.4 ± 0.11
30cis-sabinene hydrateOM10720.1 ± 0.110.4 ± 0.350.9 ± 0.650.1 ± 0.110.4 ± 0.310.8 ± 0.78-0.3 ± 0.061.1 ± 0.44
31cis-linalool oxide (furanoid)OM10752.1 ± 0.937.1 ± 3.545.5 ± 3.701.3 ± 1.024.5 ± 2.024.5 ± 3.082.2 ± 1.021.8 ± 0.724.5 ± 0.03
326,7-eoxymyrceneOM1095-------0.4 ± 0.15-
33linaloolOM110219.1 ± 3.0423.5 ± 1.9033.4 ± 11.6518.7 ± 1.0526.1 ± 5.9442.7 ± 2.4428.5 ± 4.1140.3 ± 7.8635.4 ± 6.56
34n-nonanalNT11040.1 ± 0.27--------
35(E)-2-heptyl acetate NT1114-0.1 ± 0.18-------
361-octen-3-yl acetateNT11174.9 ± 3.263.6 ± 2.821.0 ± 0.305.3 ± 4.276.5 ± 0.640.5 ± 0.513.9 ± 3.941.0 ± 0.260.3 ± 0.26
373-octanol acetateNT11290.5 ± 0.080.3 ± 0.050.1 ± 0.020.7 ± 0.570.1 ± 0.03-0.8 ± 0.780.1 ± 0.06-
38allo-ocimeneOM11330.1 ± 0.090.1 ± 0.03-0.1 ± 0.02--0.1 ± 0.080.1 ± 0.020.1 ± 0.07
39nopinoneOM1142----0.1 ± 0.07----
40trans-pinocarveol OM1144----0.1 ± 0.09----
41eucarvone OM1146----0.1 ± 0.07----
42camphorOM11480.6 ± 0.070.6 ± 0.010.2 ± 0.140.4 ± 0.100.9 ± 0.780.2 ± 0.200.7 ± 0.270.3 ± 0.080.4 ± 0.17
43hexyl isobutyrateNT11530.1 ± 0.060.2 ± 0.12-0.2 ± 0.180.2 ± 0.020.1 ± 0.080.1 ± 0.030.1 ± 0.01-
44pinocarvoneOM1166----0.1 ± 0.10----
45borneolOM11690.5 ± 0.050.5 ± 0.070.6 ± 0.430.3 ± 0.031.0 ± 0.800.6 ± 0.330.9 ± 0.740.2 ± 0.081.0 ± 0.43
46pinocampheolOM1170-------0.1 ± 0.02-
47lavandulolOM11720.1 ± 0.03-0.2 ± 0.07---0.1 ± 0.090.2 ± 0.030.1 ± 0.04
48trans-linalool oxide (pyranoid)OM1177-0.4 ± 0.04-------
494-terpineolOM11801.8 ± 0.962.5 ± 1.925.1 ± 1.361.1 ± 0.222.6 ± 1.095.3 ± 3.690.9 ± 0.503.3 ± 1.967.4 ± 1.98
50cryptoneNT11870.1 ± 0.080.3 ± 0.31--0.4 ± 0.58-0.3 ± 0.29--
51octanoic acidNT11910.1 ± 0.08-----0.1 ± 0.04--
52α-terpineolOM1192--0.1 ± 0.08------
53dihydro carveolOM1194-0.1 ± 0.10-------
54hexyl butyrateNT11950.7 ± 0.511.2 ± 0.680.5 ± 0.071.2 ± 0.530.8 ± 0.730.7 ± 0.630.2 ± 0.030.3 ± 0.080.3 ± 0.04
55myrtinalOM1196----0.1 ± 0.18----
56n-decanal NT12060.1 ± 0.02--0.1 ± 0.070.2 ± 0.19----
57verbenoneOM1214----0.1 ± 0.18----
58isobornyl formateOM1230-0.1 ± 0.101.4 ± 0.36-0.2 ± 0.13-0.1 ± 0.7--
59hexyl 3-methyl butanoate NT12420.1 ± 0.030.1 ± 0.04-0.1 ± 0.100.1 ± 0.09----
60cuminaldehydeOM1244--0.1 ± 0.04------
61linalyl acetateOM126027.0 ± 6.3524.6 ± 6.7223.0 ± 5.2126.6 ± 5.9532.0 ± 7.2722.3 ± 3.7524.4 ± 3.3222.6 ± 4.6620.6 ± 1.20
62isobornyl acetateOM12870.9 ± 0.380.1 ± 0.10-0.2 ± 0.19--0.4 ± 0.280.1 ± 0.03-
63lavandulyl acetateOM12894.1 ± 1.983.4 ± 2.413.7 ± 1.905.3 ± 3.273.8 ± 0.741.4 ± 0.584.5 ± 3.032.7 ± 0.753.5 ± 1.84
64(Z)-8-hydroxylinalolOM13600.6 ± 0.461.1 ± 0.850.4 ± 0.010.4 ± 0.450.4 ± 0.410.1 ± 0.200.6 ± 0.560.2 ± 0.150.2 ± 0.06
65neryl acetateOM13681.8 ± 0.641.1 ± 0.910.2 ± 0.061.8 ± 0.040.2 ± 0.130.2 ± 0.031.1 ± 1.020.6 ± 0.030.2 ± 0.02
66α-copaeneSH1376--0.1 ± 0.08----0.1 ± 0.040.1 ± 0.09
67β-bourboreneSH1383--0.1 ± 0.04------
68geranyl acetateOM13863.9 ± 0.942.3 ± 1.740.4 ± 0.123.3 ± 0.110.5 ± 0.080.4 ± 0.042.2 ± 1.901.3 ± 0.670.5 ± 0.09
69(E)-caryophylleneSH14182.7 ± 1.741.3 ± 1.001.7 ± 1.271.6 ± 1.031.3 ± 0.802.0 ± 0.582.0 ± 1.163.0 ± 1.221.8 ± 1.15
70α-santhalene SH1419---0.2 ± 0.27-----
71trans-α-bergamoteneSH1437-0.1 ± 0.100.1 ± 0.02--0.1 ± 0.01--0.1 ± 0.02
72(E)-β-farneseneSH14600.5 ± 0.040.3 ± 0.070.3 ± 0.271.0 ± 0.070.1 ± 0.090.4 ± 0.360.2 ± 0.190.3 ± 0.150.2 ± 0.21
73germacrene DSH14810.1 ± 0.08-0.2 ± 0.19--0.2 ± 0.01-0.2 ± 0.160.2 ± 0.17
74trans-γ-cadineneSH15130.1 ± 0.09-----0.2 ± 0.07--
75caryophyllene oxideOS1582-0.1 ± 0.02-0.1 ± 0.02-----
Class of Compounds P:CP:AP:C:A
SusStuTanSusStuTanSusStuTan
Monoterpene hydrocarbons (MH) 15.5 ± 2.39.7 ± 2.367.2 ± 4.8114.9 ± 3.345.7 ± 1.427.0 ± 2.8511.1 ± 3.313.3 ± 5.812.1 ± 4.46
Oxygenated monoterpenes (OM) 69.4 ± 3.8877.3 ± 4.1883.0 ± 6.0668.8 ± 4.9081.8 ± 3.8984.9 ± 5.7375.5 ± 6.7377.6 ± 4.0081.4 ± 1.33
Sesquiterpene hydrocarbons (SH) 3.4 ± 0.641.7 ± 0.802.4 ± 1.762.8 ± 1.962.5 ± 0.212.7 ± 0.832.4 ± 1.133.5 ± 1.412.3 ± 0.90
Oxygenated sesquiterpenes (OS) -0.1 ± 0.12-0.1 ± 0.12-----
Non-terpene derivatives (NT) 10.8 ± 2.8010.7 ± 2.336.4 ± 3.5313.4 ± 5.4710.9 ± 1.034.4 ± 1.2410.8 ± 4.114.7 ± 2.513.7 ± 1.50
Total identified 99.1 ± 0.9197.5 ± 2.5798.5 ± 1.2396.5 ± 3.3199.9 ± 0.1099.8 ± 0.2299.9 ± 0.1299.7 ± 0.1199.7 ± 0.18
1 LRI: Linear retention indices on DB-5 column; 2 The percentages are averages of at least three independent samples for each lavender selection and for each substrate. Compounds with abundance < 0.1% are not present in the table.
Table 4. Effects of lavender selection (Susa (Sus), Stura (Stu) and Tanaro (Tan)) and cultivation substrate (P:C = peat and green compost, 70%:30% v/v; P:A = peat and demolition aggregates, 70%:30% v/v; P:C:A = peat, green compost and demolition aggregates, 40%:30%:30% v/v) on volatile organic compounds (VOCs) and essential oils (Eos), according to the two-way PERMANOVA analysis.
Table 4. Effects of lavender selection (Susa (Sus), Stura (Stu) and Tanaro (Tan)) and cultivation substrate (P:C = peat and green compost, 70%:30% v/v; P:A = peat and demolition aggregates, 70%:30% v/v; P:C:A = peat, green compost and demolition aggregates, 40%:30%:30% v/v) on volatile organic compounds (VOCs) and essential oils (Eos), according to the two-way PERMANOVA analysis.
VOCsFpSignificant Pair-Wise Comparisons at p < 0.05
Lavender selection3.270.0001Sus versus Tan
Substrate0.690.2156-
EOs
Lavender selection2.940.0013Sus versus Tan
Substrate2.480.0065P:A versus P:C:A
Table 5. Percentage of major constituents in lavender essential oils obtained from lavender selections collected in Susa (Sus), Stura (Stu), and Tanaro (Tan) valleys and grown in pot with different substrates (P:C = peat and green compost, 70%:30% v/v; P:A = peat and demolition aggregates, 70%:30% v/v; P:C:A = peat, green compost and demolition aggregates, 40%:30%:30% v/v) with standard deviation.
Table 5. Percentage of major constituents in lavender essential oils obtained from lavender selections collected in Susa (Sus), Stura (Stu), and Tanaro (Tan) valleys and grown in pot with different substrates (P:C = peat and green compost, 70%:30% v/v; P:A = peat and demolition aggregates, 70%:30% v/v; P:C:A = peat, green compost and demolition aggregates, 40%:30%:30% v/v) with standard deviation.
CompoundsClassLRI 1P:CP:AP:C:A
SusStuTanSusStuTanSusStuTan
1α-thujeneMH932-----0.1 2 ± 0.07--0.1 ± 0.04
2tricycleneMH9380.1 ± 0.100.1 ± 0.070.1 ± 0.070.1 ± 0.01-0.3 ± 0.060.1 ± 0.080.1 ± 0.090.3 ± 0.01
3campheneMH9550.6 ± 0.290.2 ± 0.030.2 ± 0.020.6 ± 0.20.1 ± 0.030.5 ± 0.110.4 ± 0.090.4 ± 0.190.6 ± 0.16
41-octen-3-oneNT9800.2 ± 0.030.2 ± 0.10-0.1 ± 0.02--0.8 ± 0.470.1 ± 0.06-
5β-pineneMH9810.3 ± 0.06--0.4 ± 0.040.1 ± 0.010.2 ± 0.06-0.1 ± 0.030.2 ± 0.05
61-octen-3-ol NT9820.2 ± 0.160.4 ± 0.260.3 ± 0.110.3 ± 0.120.3 ± 0.160.3 ± 0.12-0.7 ± 0.180.3 ± 0.09
73-octanoneNT9870.7 ± 0.340.5 ± 0.170.4 ± 0.260.9 ± 0.380.5 ± 0.350.7 ± 0.261.3 ± 0.160.5 ± 0.070.6 ± 0.14
8myrceneMH9931.0 ± 0.240.7 ± 0.070.5 ± 0.101.0 ± 0.110.7 ± 0.160.9 ± 0.201.2 ± 0.590.6 ± 0.221.0 ± 0.39
9butanoic acid butyl esterNT994-0.1 ± 0.08-0.2 ± 0.10.1 ± 0.08----
103-octanolNT9980.1 ± 0.080.1 ± 0.100.1 ± 0.050.1 ± 0.060.1 ± 0.090.2 ± 0.080.2 ± 0.040.2 ± 0.050.2 ± 0.01
11cis-dehydroxylinalool oxideOM10090.1 ± 0.060.1 ± 0.09-0.1 ± 0.080.1 ± 0.06-0.1 ± 0.080.1 ± 0.09-
12N-hexyl acetateNT10130.3 ± 0.040.2 ± 0.10-0.2 ± 0.520.2 ± 0.070.1 ± 0.090.7 ± 0.17--
13α-terpineneMH10190.1 ± 0.020.1 ± 0.080.1 ± 0.090.1 ± 0.090.1 ± 0.090.2 ± 0.03-0.1 ± 0.100.2 ± 0.05
14o-cymene MH10260.1 ± 0.010.1 ± 0.100.1 ± 0.090.1 ± 0.060.1 ± 0.090.1 ± 0.01-0.1 ± 0.090.1 ± 0.04
15p-cymene MH10280.3 ± 0.040.5 ± 0.240.4 ± 0.210.3 ± 0.060.5 ± 0.140.5 ± 0.070.2 ± 0.030.4 ± 0.240.5 ± 0.24
16limoneneMH1032-0.2 ± 0.060.3 ± 0.07-0.1 ± 0.030.2 ± 0.03-0.3 ± 0300.3 ± 0.03
171,8-cineole OM10366.3 ± 0.471.5 ± 0.451.0 ± 0.155.2 ± 1.841.6 ± 0.762.3 ± 0.293.6 ± 0.914.3 ± 0.882.0 ± 0.46
18(Z)-β-ocimene MH10420.8 ± 0.330.4 ± 0.270.5 ± 0.200.9 ± 0.250.5 ± 0.330.9 ± 0.461.0 ± 0.160.8 ± 0.250.9 ± 0.47
19(E)-β-ocimene MH10530.9 ± 0.240.6 ± 0.120.7 ± 0.181.0 ± 0.020.7 ± 0.381.1 ± 0.211.2 ± 0.420.5 ± 0.421.1 ± 0.42
20γ-terpineneMH1062--0.1 ± 0.06--0.2 ± 0.02--0.2 ± 0.01
21trans-linalool oxide (furanoid)OM10691.2 ± 0.282.0 ± 0.671.3 ± 0.431.7 ± 0.701.7 ± 0.182.2 ± 0.161.0 ± 0.131.7 ± 0.702.5 ± 0.57
22cis-sabinene hydrate OM10720.2 ± 0.100.4 ± 0.160.4 ± 0.270.3 ± 0.040.4 ± 0.050.5 ± 0.350.2 ± 0.010.4 ± 0.180.4 ± 0.14
23cis-linalool oxide (furanoid)OM10751.2 ± 0.201.8 ± 0.631.2 ± 0.331.6 ± 0.571.6 ± 0.922.2 ± 0.181.1 ± 0.251.5 ± 0.862.5 ± 0.17
24campheniloneOM1086-0.1 ± 0.09-------
256,7-epoxymyrceneOM10950.5 ± 0.130.5 ± 0.100.3 ± 0.150.5 ± 0.210.5 ± 0.020.2 ± 0.160.4 ± 0.060.3 ± 0.020.2 ± 0.05
26linaloolOM110221.0 ± 3.2423.9 ± 3.3629.5 ± 3.5123.7 ± 3.7720.4 ± 3.4036.9 ± 5.3833.8 ± 3.7236.7 ± 4.0338.5 ± 0.69
271-octen-3-yl acetate NT11173.0 ± 0.732.4 ± 1.300.6 ± 0.083.1 ± 1.323.0 ± 0.560.5 ± 0.761.6 ± 0.170.8 ± 0.730.5 ± 0.06
28cis-p-menth-2-en-1-olOM1125-0.2 ± 0.110.1 ± 0.09-0.2 ± 0.140.1 ± 0.080.1 ± 0.080.2 ± 0.070.1 ± 0.02
293-octanol acetateNT11290.5 ± 0.340.1 ± 0.070.1 ± 0.010.5 ± 0.230.1 ± 0.050.1 ± 0.060.5 ± 0.06-0.1 ± 0.02
30α-campholenalOM1130-0.1 ± 0.07-----0.1 ± 0.08-
31(Z)-myroxide OM11370.2 ± 0.090.2 ± 0.050.1 ± 0.090.2 ± 0.070.2 ± 0.050.1 ± 0.100.1 ± 0.090.1 ± 0.04-
32trans-pinocarveolOM11420.1 ± 0.10--0.1 ± 0.10.1 ± 0.08-0.1 ± 0.080.1 ± 0.08-
33eucarvone OM1146---0.1 ± 0.08---0.1 ± 0.05-
34camphorOM11481.6 ± 0.221.8 ± 0.161.2 ± 0.521.3 ± 0.271.4 ± 0.341.4 ± 0.61.3 ± 0.351.6 ± 0.381.3 ± 0.71
35trans-verbenolOM11500.1 ± 0.01-0.1 ± 0.040.1 ± 0.090.1 ± 0.080.1 ± 0.010.1 ± 0.02-0.1 ± 0.07
36hexyl isobutyrateNT11530.2 ± 0.080.2 ± 0.12-0.2 ± 0.020.2 ± 0.060.1 ± 0.080.1 ± 0.070.1 ± 0.080.1 ± 0.03
37nerol oxideOM1158-0.1 ± 0.05-0.1 ± 0.080.1 ± 0.09--0.1 ± 0.030.1 ± 0.03
38pinocarvoneOM11660.3 ± 0.100.1 ± 0.100.1 ± 0.080.2 ± 0.110.2 ± 0.160.1 ± 0.080.1 ± 0.070.1 ± 0.08-
39borneolOM11694.4 ± 0.585.3 ± 1.094.6 ± 1.663.6 ± 0.454.6 ± 1.485.3 ± 2.084.0 ± 0.675.1 ± 0.595.3 ± 0.39
404-terpinenolOM11800.9 ± 0.561.4 ± 0.553.0 ± 0.621.2 ± 0.672.3 ± 0.764.9 ± 0.801.7 ± 0.111.6 ± 0.114.1 ± 0.30
41cryptoneNT11870.7 ± 0.471.1 ± 0.560.8 ± 0.540.3 ± 0.061.1 ± 0.570.8 ± 0.300.3 ± 0.061.3 ± 0.280.8 ± 0.24
42p-cymen-8-olOM11890.2 ± 0.130.2 ± 0.100.1 ± 0.100.2 ± 0.050.2 ± 0.090.1 ± 0.010.1 ± 0.08-0.1 ± 0.04
43α-terpineolOM11924.1 ± 0.983.1 ± 0.362.9 ± 0.514.5 ± 0.563.6 ± 1.104.1 ± 1.134.3 ± 0.202.9 ± 0.224.4 ± 0.13
44verbenoneOM12140.3 ± 0.120.4 ± 0.170.3 ± 0.110.2 ± 0.080.4 ± 0.190.3 ± 0.140.1 ± 0.030.3 ± 0.140.2 ± 0.10
45trans-carveolOM12210.5 ± 0.380.6 ± 0.150.3 ± 0.200.5 ± 0.230.6 ± 0.080.2 ± 0.070.4 ± 0.260.3 ± 0.080.1 ± 0.06
46cis-p-mentha-1(7),8-dien-2-olOM12290.2 ± 0.170.3 ± 0.090.2 ± 0.110.1 ± 0.070.3 ± 0.160.1 ± 0.12-0.2 ± 0.100.1 ± 0.05
47isobornyl formateOM12300.3 ± 0.050.3 ± 0.210.3 ± 0.140.2 ± 0.020.3 ± 0.130.2 ± 0.100.2 ± 0.010.3 ± 0.020.2 ± 0.13
48nerolOM12321.0 ± 0.360.6 ± 0.080.7 ± 0.160.9 ± 0.070.7 ± 0.260.8 ± 0.231.1 ± 0.340.5 ± 0.170.9 ± 0.50
493-methyl-3hexen-1-yl butanoateNT12360.1 ± 0.070.1 ± 0.09-0.1 ± 0.030.1 ± 0.07-0.1 ± 0.03--
50cumin aldehydeOM12440.4 ± 0.260.7 ± 0.410.4 ± 0.070.2 ± 0.170.7 ± 0.390.3 ± 0.180.2 ± 0.020.8 ± 0.260.2 ± 0.05
51carvoneOM12480.2 ± 0.130.3 ± 0.200.2 ± 0.170.1 ± 0.020.3 ± 0.200.1 ± 0.030.1 ± 0.010.4 ± 0.110.1 ± 0.06
52linalyl acetateOM126017.7 ± 1.5515.3 ± 2.3314.9 ± 1.3216.7 ± 3.9516.2 ± 2.027.5 ± 2.8916.5 ± 1.2410.3 ± 0.798.8 ± 1.81
53isopulegol acetateOS12730.1 ± 0.08--0.1 ± 0.08-1.8 ± 0.170.1 ± 0.07--
54isobornyl acetateOM12871.6 ± 0.310.6 ± 0.050.3 ± 0.061.1 ± 0.280.5 ± 0.40.2 ± 0.10.9 ± 0.290.3 ± 0.110.3 ± 0.09
55lavandulyl acetateOM12893.3 ± 1.653.9 ± 1.083.8 ± 1.214.8 ± 1.035.0 ± 0.484.4 ± 1.192.7 ± 1.322.1 ± 0.885.1 ± 0.79
56carvacrolOM1301-0.3 ± 0.140.4 ± 0.06-0.4 ± 0.230.1 ± 0.02-0.1 ± 0.080.1 ± 0.03
57hexyl tiglateNT1333----0.1 ± 0.01-0.1 ± 0.02--
58δ-elemeneSH13400.1 ± 0.010.5 ± 0.250.5 ± 0.230.1 ± 0.080.6 ± 0.40.2 ± 0.08-0.3 ± 0.140.3 ± 0.13
59(Z)-8-hydroxylinalol OM13600.6 ± 0.280.5 ± 0.130.2 ± 0.050.6 ± 0.230.7 ± 0.260.2 ± 0.090.3 ± 0.080.4 ± 0.14-
60neryl acetateOM13681.5 ± 0.361.3 ± 0.331.3 ± 0.271.5 ± 0.091.4 ± 0.411.3 ± 0.331.9 ± 0.400.9 ± 0.201.2 ± 0.60
61α-copaeneSH1376--0.1 ± 0.10------
62geranyl acetateOM13863.1 ± 0.792.8 ± 0.572.8 ± 0.643.1 ± 0.232.9 ± 0.852.6 ± 0.733.6 ± 0.571.9 ± 0.202.5 ± 0,23
63(E)-caryophylleneSH14181.3 ± 0.481.8 ± 0.683.2 ± 1.001.4 ± 0.492.4 ± 0.891.6 ± 0.282.0 ± 0.911.9 ± 0.381.4 ± 0.12
64trans-γ-bergamoteneSH1437-0.2 ± 0.100.3 ± 0.11-0.2 ± 0.120.1 ± 0.01-0.1 ± 0.070.1 ± 0.07
65aromadendreneSH1445-0.1 ± 0.080.2 ± 0.07-0.1 ± 0.120.1 ± 0.08-0.1 ± 0.09-
66epi-β-santaleneSH1447-0.1 ± 0.090.1 ± 0.08-0.1 ± 0.01----
67α-humuleneSH1456----0.1 ± 0.07----
68(E)-β-farnesene SH14600.3 ± 0.060.4 ± 0.090.7 ± 0.060.4 ± 0.200.4 ± 0.230.5 ± 0.140.4 ± 0.080.3 ± 0.080.3 ± 0.16
69germacrene DSH14810.1 ± 0.070.1 ± 0.070.5 ± 0.040.1 ± 0.070.1 ± 0.030.3 ± 0.110.2 ± 0.020.1 ± 0.160.3 ± 0.08
70γ-curcumeneSH1484--0.2 ± 0.09--0.1 ± 0.06-0.1 ± 0.06-
71β-bisaboleneSH1509--0.6 ± 0.09------
72trans-γ-cadineneSH15130.5 ± 0.160.1 ± 0.06-0.3 ± 0.230.2 ± 0.11-0.2 ± 0.060.2 ± 0.25-
73(Z)-γ-bisaboleneSH15150.5 ± 0.101.9 ± 0.801.3 ± 0.190.5 ± 0.182.0 ± 0.471.1 ± 0.380.3 ± 0.041.0 ± 0.540.9 ± 0.31
74(E)-γ-bisaboleneSH15350.2 ± 0.050.8 ± 0.410.9 ± 0.370.2 ± 0.071.1 ± 0.710.5 ± 0.16-0.3 ± 0.070.3 ± 0.13
75cis-sesquisabinene hydrateOS1545-0.3 ± 0.160.3 ± 0.120.1 ± 0.040.3 ± 0.250.1 ± 0.03-0.1 ± 0.020.1 ± 0.07
76elemolOS15530.5 ± 0.220.7 ± 0.060.6 ± 0.260.4 ± 0.060.6 ± 0.200.3 ± 0.110.3 ± 0.010.5 ± 0.210.3 ± 0.10
77germacrene BSH1556-0.2 ± 0.080.3 ± 0.09-0.2 ± 0.080.1 ± 0.02-0.1 ± 0.080.1 ± 0.02
78spathulenolOS15810.1 ± 0.010.3 ± 0.130.2 ± 0.08-0.3 ± 0.220.2 ± 0.04-0.2 ± 0.050.1 ± 0.02
79caryophyllene oxideOS15826.1 ± 1.217.4 ± 1.406.6 ± 1.416.0 ± 2.526.2 ± 1.083.6 ± 1.054.8 ± 0.205.8 ± 0.673.4 ± 0.49
80thujapsan-2-α-olOS1589-1.4 ± 0.811.5 ± 0.770.1 ± 0.091.4 ± 0,420.9 ± 0.32-0.6 ± 0.260.7 ± 0.19
81β-oplopenoneOS1606----0.1 ± 0.09----
82humulene epoxide IIOS16070.2 ± 0.060.2 ± 0.040.2 ± 0.070.1 ± 0.050.2 ± 0.050.1 ± 0.09-0.2 ± 0.070.1 ± 0.10
831,10-di-epi-cubenolOS16140.3 ± 0.030.2 ± 0.070.2 ± 0.050.1 ± 0.070.2 ± 0.06-0.1 ± 0.020.2 ± 0.07-
84α-acorenolOS1633-0.1 ± 0.01--0.1 ± 0.02-0.1 ± 0.08--
85β-acorenolOS1636-0.1 ± 0.090.1 ± 0.12-0.1 ± 0.02----
86β-caryophylla-4(14),8(15)-dien-5-olOS1639-0.2 ± 0.040.2 ± 0.09-0.1 ± 0.040.1 ± 0.03-0.1 ± 0.080.1 ± 0.04
87τ-cadinol OS16423.0 ± 0.310.6 ± 0.370.2 ± 0.101.9 ± 0.151.1 ± 0.590.1 ± 0.030.8 ± 0.152.0 ± 0.970.1 ± 0.08
88α-cadinolOS1655-0.1 ± 0.100.4 ± 0.11-0.2 ± 0.020.2 ± 0.060.2 ± 0.120.1 ± 0.060.1 ± 0.02
89neo-intermediolOS1660-0.3 ± 0.020.2 ± 0.12-0.3 ± 0.08--0.2 ± 0.08-
90(Z)-α-santalolOS16650.7 ± 0.250.9 ± 0.341.0 ± 0.300.7 ± 0.340.4 ± 0.060.5 ± 0.080.3 ± 0.040.5 ± 0.210.4 ± 0.11
91(Z)-nerolidol acetateOS1668--0.2 ± 0.150.1 ± 0.08-0.2 ± 0.08--0.1 ± 0.02
9214-hydroxy-9-epi-(E)-caryophylleneOS16720.1 ± 0.050.2 ± 0.100.1 ± 0.020.4 ± 0.050.3 ± 0.02--0.1 ± 0.09-
93elemol acetateOS1675-0.4 ± 0.330.4 ± 0.26-0.6 ± 0.060.3 ± 0.04-0.3 ± 0.090.1 ± 0.03
94cis-14-muurol-5-en-4-oneOS16840.7 ± 0.130.2 ± 0.160.3 ± 0.120.2 ± 0.020.4 ± 0.170.1 ± 0.010.2 ± 0.050.5 ± 0.070.1 ± 0.10
9514-hydroxy-α-humuleneOS1714--0.1 ± 0.02-0.1 ± 0.10----
96curcuphenolOS1720-------0.1 ± 0.01-
97cedr-8(15)-en-9-α-ol acetateOS17430.1 ± 0.02--0.1 ± 0.04-----
98cyclocolorenoneOS17580.5 ± 0.060.1 ± 0.10-0.4 ± 0.230.1 ± 0.02-0.2 ± 0.020.2 ± 0.06-
99benzyl benzoateNT17600.1 ± 0.090.2 ± 0.10-0.1 ± 0.08--0.1 ± 0.080.1 ± 0.08-
100(Z)-α-santalol acetateOS17630.1 ± 0.070.1 ± 0.09--0.2 ± 0.14--0.1 ± 0.08-
101hexahydrofarnesylacetone AC18450.2 ± 0.060.4 ± 0.120.2 ± 0.200.1 ± 0.030.2 ± 0.020.1 ± 0.070.1 ± 0.040.3 ± 0.14-
Yield of EO (% w/w) 0.4 ± 0.080.1 ± 0.050.2 ± 0.021.0 ± 0.530.4 ± 0.100.6 ± 0.25vl 30.2 ± 0.100.8 ± 0.37
P:CP:AP:C:A
Class of Compounds SusStuTanSusStuTanSusStuTan
Monoterpene Hydrocarbons (MH) 4.4 ± 0.972.9 ± 0.572.8 ± 0.674.4 ± 1.082.9 ± 0.765.1 ± 0.514.1 ± 1.063.5 ± 1.355.6 ± 1.37
Oxygenated Monoterpenes (OM) 73.2 ± 3.9970.6 ± 2.3672.2 ± 3.6574.6 ± 3.3769.6 ± 4.2878.8 ± 1.4579.6 ± 1.4675.8 ± 2.9781.8 ± 1.12
Sesquiterpene Hydrocarbons (SH) 3.1 ± 0.476.2 ± 0.328.8 ± 1.973.1 ± 0.237.5 ± 1.974.6 ± 0.233.1 ± 1.044.5 ± 0.703.8 ± 0.33
Oxygenated Sesquiterpenes (OS) 12.7 ± 2.2013.7 ± 1.3513.2 ± 2.8810.9 ± 1.6813.1 ± 2.548.7 ± 1.827.0 ± 1,8111.6 ± 1.685.9 ± 0.73
Non-terpene derivatives (NT) 6.0 ± 0.235.5 ± 0.222.4 ± 0.776.1 ± 1.706.0 ± 0.602.8 ± 0.855.7 ± 0.313.8 ± 1.122.8 ± 0.84
Apocarotenoids (AC) 0.2 ± 0.060.4 ± 0.120.2 ± 0.200.1 ± 0.030.2 ± 0.020.1 ± 0.070.1 ± 0.040.3 ± 0.14-
Total Identified 99.5 ± 0.4099.3 ± 0.3699.5 ± 0.2699.2 ± 0.1499.3 ± 0.43100.0 ± 0.0199.7 ± 0.1899.4 ± 0.5599.9 ± 0.10
1 LRI: Linear retention indices on DB-5 column; 2 The percentages are averages of at least three independent samples for each lavender selection and for each substrate. Compounds with abundance < 0.1% are not present in the table; 3 vl: very low amount.
Table 6. List of compounds responsible for dissimilarity in lavender EOs induced by substrate (P:C = peat and green compost, 70%:30% v/v; P:A = peat and demolition aggregates, 70%:30% v/v; P:C:A = peat, green compost and demolition aggregates, 40%:30%:30% v/v), according to the similarity percentage (SIMPER) analysis.
Table 6. List of compounds responsible for dissimilarity in lavender EOs induced by substrate (P:C = peat and green compost, 70%:30% v/v; P:A = peat and demolition aggregates, 70%:30% v/v; P:C:A = peat, green compost and demolition aggregates, 40%:30%:30% v/v), according to the similarity percentage (SIMPER) analysis.
Contribution %Cumulative %Substrate P:CSubstrate P:ASubstrate P:C:AStat. Sign.Significant Pair-Wise Comparisons at p < 0.05
linalool58.1158.1126.6027.0036.60*P:C versus P:C:A
linalyl acetate18.7576.8615.5013.5011.40ns-
1,8-cineole4.1681.022.133.073.13ns-
caryophyllene oxide3.1084.126.695.304.50*P:C versus P:C:A
lavandulyl acetate2.7886.903.754.713.56ns-
4-terpineol1.8088.702.232.822.84ns-
borneol1.7690.464.744.514.90ns-
1-octen-3-yl acetate1.2691.721.492.220.91ns-
τ-cadinol1.2392.950.821.030.86ns-
α-terpineol0.7493.693.174.063.94ns-
isopulegol acetate0.7194.400.030.660.01ns-
(E)-caryophyllene0.6695.062.501.811.70ns-
(Z)-γ-bisabolene 0.5795.631.291.210.79ns-
thujapsan-2-α-ol 0.5196.141.190.800.50ns-
geranyl acetate0.3296.462.872.862.66ns-
trans-linalool oxide0.3296.781.441.871.83ns-
cis-linalool oxide0.2797.051.341.781.80ns-
β-bisabolene0.1997.240.360.000.00ns-
camphor0.1797.411.421.401.34ns-
cryptone0.1797.580.830.740.79ns-
isobornyl acetate0.1697.740.630.610.43ns-
(E)-γ-bisabolene0.1697.900.740.560.24*P:C versus P:C:A
(Z)-α-santalol0.1198.010.900.520.40*P:C versus P:A; P:C versus P:C:A
Stat. Sign.: Statistical significance: * p < 0.05; ns, not significant.

Share and Cite

MDPI and ACS Style

Najar, B.; Demasi, S.; Caser, M.; Gaino, W.; Cioni, P.L.; Pistelli, L.; Scariot, V. Cultivation Substrate Composition Influences Morphology, Volatilome and Essential Oil of Lavandula Angustifolia Mill. Agronomy 2019, 9, 411. https://doi.org/10.3390/agronomy9080411

AMA Style

Najar B, Demasi S, Caser M, Gaino W, Cioni PL, Pistelli L, Scariot V. Cultivation Substrate Composition Influences Morphology, Volatilome and Essential Oil of Lavandula Angustifolia Mill. Agronomy. 2019; 9(8):411. https://doi.org/10.3390/agronomy9080411

Chicago/Turabian Style

Najar, Basma, Sonia Demasi, Matteo Caser, Walter Gaino, Pier Luigi Cioni, Luisa Pistelli, and Valentina Scariot. 2019. "Cultivation Substrate Composition Influences Morphology, Volatilome and Essential Oil of Lavandula Angustifolia Mill." Agronomy 9, no. 8: 411. https://doi.org/10.3390/agronomy9080411

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop