Skip to main content
Horticulture Research logoLink to Horticulture Research
. 2022 Jan 19;9:uhab057. doi: 10.1093/hr/uhab057

Cucurbitaceae genome evolution, gene function, and molecular breeding

Lili Ma 1,2,4, Qing Wang 1,4, Yanyan Zheng 1,4, Jing Guo 3, Shuzhi Yuan 1, Anzhen Fu 1, Chunmei Bai 1, Xiaoyan Zhao 1, Shufang Zheng 1, Changlong Wen 1, Shaogui Guo 1, Lipu Gao 1, Donald Grierson 4,, Jinhua Zuo 1,, Yong Xu 1,
PMCID: PMC8969062  PMID: 35043161

Abstract

Cucurbitaceae is one of the most genetically diverse plant families in the world. Many of them are important vegetables or medicinal plants and are widely distributed worldwide. The rapid development of sequencing technologies and bioinformatic algorithms has enabled the generation of genome sequences of numerous important Cucurbitaceae species. This has greatly facilitated research on gene identification, genome evolution, genetic variation, and molecular breeding of cucurbit crops. So far, genome sequences of 18 different cucurbit species belonging to tribes Benincaseae, Cucurbiteae, Sicyoeae, Momordiceae, and Siraitieae have been deciphered. This review summarizes the genome sequence information, evolutionary relationships, and functional genes associated with important agronomic traits (e.g. fruit quality). The progress of molecular breeding in cucurbit crops and prospects for future applications of Cucurbitaceae genome information are also discussed.

Introduction

Cucurbitaceae is the second largest fruit and vegetable family and its members are among the most important edible plants in the world, next only to Solanaceae [1, 2]. The family contains ~115 genera and 960 species, which are mostly herbaceous annual vines or perennial lianas, often with tendrils [3]. They can be monoecious or dioecious (occasionally hermaphrodite) and are mainly distributed in tropical and subtropical zones, rarely in temperate zones [3]. A characteristic feature of the Cucurbitaceae is the existence of bicollateral vascular bundles where the phloem is present on both the outer and the inner side of the xylem [4]. Cucurbits frequently contain cucurbitacin, which is the main substance causing the bitter taste [5]. The family Cucurbitaceae contains a variety of vegetables or fruit crops, which are of great significance to the global or local economy. The vegetables include cucumber (Cucumis sativus), zucchini (Cucurbita pepo), pumpkin (Cucurbita maxima, Cucurbita moschata, and Cucurbita argyrosperma), wax gourd (Benincasa hispida), bottle gourd (Lagenaria siceraria), bitter gourd (Momordica charantia), ridge gourd (Luffa acutangula), sponge gourd (Luffa cylindrica), chayote (Sechium edule), and snake gourd (Trichosanthes anguina), and the fruits include melon (Cucumis melo), horned cucumber (Cucumis metuliferus), watermelon (Citrullus lanatus), and luo-han-guo (Siraitia grosvenorii) [2, 3]. Among them, bitter gourd and luo-han-guo both have rich edible and medicinal value [6] and snake gourd and bottle gourd can be used as food and ornaments [7, 8].

Recently, thanks to the rapid advances in sequencing technologies and bioinformatic algorithms, the application of whole-genome sequencing technology in biology has become more and more common [9]. Due to the high cost and low throughput of Sanger sequencing, the initial genome sequencing work was limited to few plant species, mainly model species such as Arabidopsis thaliana [10] and Oryza sativa [11]. The first Cucurbitaceae crop genome, that of cucumber, was sequenced using Sanger and next-generation Illumina sequencing technologies and released in 2009 [12]. With the emergence of next-generation sequencing, the cost of sequencing was greatly reduced and efficiency was substantially improved, making possible whole-genome sequencing of many commercially important plants in addition to model organisms. Most importantly, third-generation sequencing technologies (e.g. Oxford Nanopore and Pacific Biosciences) that produce longer reads instead of short reads, chromosome conformation capture techniques, and novel computational methods have together improved the completeness and contiguity of genome assemblies [13, 14]. To date, a number of Cucurbitaceae genomes have been assembled, including Cucumis sativus [12, 1519], Cucumis melo [2025], Cucumis hystrix [26], Cucumis × hystrix [27], Cucumis metuliferus [28], Cucurbita pepo [29], Luffa siceraria [30], C.itrullus lanatus [3133], Cucurbita moschata, Cucurbita maxima [34], Cucurbita argyrosperma [35, 36], B. hispida [37], Luffa cylindrica [38–40], Luffa acutangula [40], M. charantia [6, 41, 42], Siraitia grosvenorii [43, 44], T. anguina [45], and Sechium edule [46].

The completion of genome sequencing of several Cucurbitaceae crops has injected new impetus into the study of genome structure and functional evolution of Cucurbitaceae. This is of great practical significance for further study of the Cucurbitaceae at the genomic level, understanding biological mechanisms, and improving the quality of Cucurbitaceae crops at the molecular level. This review summarizes the findings of whole-genome sequencing and resequencing of Cucurbitaceae plants, which have provided basic data for genome-wide studies of important Cucurbitaceae plants, and discusses the molecular regulation of important traits and prospects for their application to improving fruit quality and promoting plant breeding.

Whole-genome sequencing of cucurbit crops

As sequencing technologies have developed rapidly, the experimental data and genome sequences of some species have been reinterpreted or revised and improved using new technologies, and this has enabled more complete genome assemblies to be constructed [16, 24]. The cucumber genome sequence [12] was quickly followed by melon [23] and watermelon [31] sequences. Many improved or new genome assemblies of Cucurbitaceae species have been produced during the past 5 years (Table 1). The assembled genome sizes of Cucurbitaceae crops range from 204.8 to 919.76 Mb with a scaffold N50 ranging from 620.88 kb to 82.12 Mb.

Table 1.

Overview of genome sequences of Cucurbitaceae plant s.

Date Sequencing technologies Cucurbitaceae species Accession name Chromosome number (n) Genome size (Mb) Contig N50 (kb) Scaffold N50 (Mb) Anchored (Mb) Oriented (Mb) Complete BUSCOs (%) Repetitive sequences (%) Protein-coding genes
2009 Sanger and Illumina Cucumber [12] (Cucumis sativus var. sativus) ‘Chinese long’ inbred line 9930 7 243.50 19.80 1.14 177.30 (72.80%) 24.00 26 682
2011 454, Sanger-Celera/Arachne Cucumber [17] (Cucumis sativus var. sativus) B10 7 323.00 23.28 0.32 16.09 26 587
2012 454 pyrosequencing and Sanger Melon [23] (Cucumis melo) Double-haploid line DHL92 12 375.00 18.20 4.68 316.30 (87.52%) 291.90 (80.77%) 19.70 27 427
2012 Cucumber [18] (Cucumis sativus var. sativus) Gy14 7 193.00 173.10 (89.90%)
2013 Illumina Watermelon 31 (Citrullus lanatus ssp. vulgaris) Inbred line 97 103 11 353.50 26.38 2.38 330.00 (93.48%) 45.20 23 440
2013 Illumina Wild cucumber [15] (Cucumis sativus var. hardwickii) PI183967 (CG0002) 7 204.80 119.00 4.20 31.10 23 836
2016 Illumina Luo-han-guo [43] (Siraitia grosvenorii) 14 420.15 34.15 0.10
2017 Illumina Bottle gourd [30] (Lagenaria siceraria) Inbred line USVL1VR-Ls 11 313.40 28.30 8.70 308.00 (98.30%) 284.00 (90.60%) 95.40 46.90 22 472
2017 Illumina Bitter gourd [6] (Momordica charantia) Inbred line OHB3-1 11 285.50 21.90 1.10 172.00 (60.20%) 95.80 15.30 45 859
2017 Illumina Squash [34] (Cucurbita maxima) Cultivar ‘Rimu’ 20 271.40 40.70 3.70 211.40 (77.90%) 201.20 (74.14%) 95.50 40.30 32 076
2017 Illumina Pumpkin [34] (Cucurbita moschata) Cultivar ‘Rifu’ 20 269.90 40.50 4.00 238.50 (88.40%) 221.30 (82.00%) 95.90 40.60 32 205
2018 Illumina HiSeq 2000 Zucchini [29] (Cucurbita pepo ssp. pepo) MU-CU-16 20 263.00 110.00 1.80 214.10 (81.40%) 92.10 37.80 27 870
2018 Illumina, SMRT Luo-han-guo [44] (Siraitia grosvenorii) Variety ‘Qingpiguo’ 14 469.50 432.38 89.20 51.14 30 565
2019 Illumina Melon [22] (Cucumis melo var. makuwa) ‘Chang Bougi’ 12 344.00 15.00 1.00 52.00 36 235
2019 Illumina Melon [22] (Cucumis melo var. Makuwa) SW3 12 354.00 25.00 1.60 54.00 38 173
2019 PacBio, BioNano, Hi-C and genetic maps Watermelon [32] (Citrullus lanatus) Inbred line 97 103 11 365.10 2300.00 21.90 362.70 (99.30%) 96.80 55.55 22 596
2019 Illumina Watermelon [33] (Citrullus lanatus ssp. vulgaris) ‘Charleston Gray’ 11 396.40 36.70 7.47 382.50 (96.50%) 379.20 (95.66%) 91.80 49.17 22 546
2019 Illumina, SMRT sequencing and Hi-C Melon [24] (Cucumis melo var. inodorus) ‘Payzawat’ 12 386.00 2860.00 380.79 (98.53%) 363.76 (95.53%) 92.78 49.80 22 924
2019 Illumina, Illumina MiSeq, and PacBio RS II Silver-seed gourd [35] (Cucurbita argyrosperma ssp. argyrosperma) SMH-JMG-627 20 228.80 463.40 0.62 93.20 34.10 28 298
2019 PacBio, 10X Genomics, and Hi-C technologies Cucumber [16] (Cucumis sativus var. sativus) ‘Chinese long’ inbred line 9930 7 226.20 8900.00 11.50 211.00 (93.30%) 36.43 24 317
2019 Illumina and PacBio Wax gourd [37] (Benincasa hispida) Inbred line B227 12 913.00 68.50 3.40 859.00 (94.10%) 91.00 75.50 27 467
2020 Illumina and PacBio Cucumber [19] (Cucumis sativus var. sativus) B10 7 342.29 858.00 91.30 27 271
2020 PacBio Melon [20] (Cucumis melo) DHL92 12 357.64 714.00 343.00 (96.00%) 94.80 29 980
2020 ONT, Bionano optical map, Illumina HiSeq, mate pair, and linkage map information Melon [21] (Cucumis melo var. reticulatus) ‘Harukei-3’ 12 378.00 8600.00 17.50 95.30 55.82 33 829
2020 Nanopore and Hi-C Snake gourd [45] (Trichosanthes anguina) 11 919.76 20110.00 82.12 918.80 (99.89%) 95.38 80.03 22 874
2020 PacBio and Hi-C Sponge gourd [38] (Luffa cylindrica) 13 669.00 5000.00 53.00 91.60 62.18 31 661
2020 Illumina, PacBio, 10× Genomics and Hi-C Sponge gourd [39] (L. cylindrica) Inbred line P93075 13 656.19 8800.00 48.76 63.81 25 508
2020 SMRT and Chicago/Hi-C Ridge gourd [39] (Luffa acutangula) Inbred line AG-4 13 734.60 0.79 92.70 62.17 32 233
2020 SMRT and Chicago/Hi-C Sponge gourd [39] (Luffa cylindrica) Inbred line SO-3 13 689.80 0.58 93.00 56.78 43 828
2020 PacBio and Hi-C Bitter gourd [41] (M. charantia) OHB3-1 11 302.99 9898.00 25.37 291.70 (96.27%) 96.40 52.52
2020 Illumina Bitter gourd [42] (M. charantia) ‘Dali-11’ 11 293.60 62.60 3.30 251.30 (85.50%) 96.70 41.50 26 427
2020 Illumina Bitter gourd [42] (M. charantia) TR 11 296.30 16.10 0.60 97.10 39.90 28 827
2021 Illumina and PacBio Silver-seed gourd [36] (Cucurbita argyrosperma ssp. sororia) 20 255.20 1205.50 12.10 98.80 92.80 30 592
2021 Illumina and PacBio Silver-seed gourd [36] (C. argyrosperma ssp. argyrosperma) SMH-JMG-627 20 231.60 447.00 11.70 99.97 93.20 27 998
2021 Illumina Cucumis hystrix [26] 12 297.50 220.95 14.06 268.90 (90.4%) 262.40 (88.2%) 93.50 23 864
2021 Illumina, SMRT, Hi-C, and BioNano optical mapping Cucumis × hytivus [27] 19 540.75 6596.00 27.20 525.78 (97.23%) 490.71 (93.33%) 50.98 45 687
2021 SMRT and Hi-C Cucumis metuliferus [28] CM27 (PI 482460) 12 329.00 2900.00 316.82 (97.99%) 316.82 (97.99%) 93.54 42.63 29 214
2021 SMRT and Hi-C Melon [28] (Cucumis melo ssp. agrestis) IVF77 12 364.00 490.00 339.72 (96.31%) 339.72 (96.31%) 93.26 51.11 27 073
2021 Illumina, SMRT Bottle gourd [8] (Lagenaria siceraria) ‘Hangzhou Gourd’ 11 297.00 11200.00 28.40 95.50 44.99 23 541
2021 Nanopore and Hi-C Chayote [46] (Sechium edule) 14 608.17 8400.00 46.56 606.42 (99.71%) 598.48 (98.41%) 95.62 65.94 28 237

BUSCOs, Benchmarking Universal Single-Copy Orthologs; SMRT, single-molecule real-time; Hi-C, high-throughput chromosome conformation capture.

According to the reported syntenic relationships among genomes of cucurbits, including melon (n = 12), cucumber (n = 7), wax gourd (n = 12), bottle gourd (n = 11), watermelon (n = 11), and pumpkin (n = 20), it is inferred that the ancestral cucurbit protochromosome number was 15 and the most ancestral state is preserved in the wax gourd genomes among these species [37]. Collinearity analysis showed that the seven chromosomes of melon are derived directly from the ancestral ones, which is the most preserved cucurbit ancestral karyotype after that of the wax gourd, and the bottle gourd genome is the third best preserved ancestral cucurbit karyotype after those of the wax gourd and melon [30, 37]. Researchers have suggested that the bottle gourd chromosomes derived from the ancestral Cucurbitaceae karyotypes through 19 chromosomal fissions and 20 fusions [30]. The modern 11-chromosome structure of watermelon evolved from the ancestral Cucurbitaceae karyotype (12 chromosomes) through 27 fissions and 28 fusions, which indicates that the watermelon genome has undergone more rearrangements than that of the bottle gourd [30]. An extensive chromosomal rearrangement has also occurred in the zucchini genome [29], and the complicated syntenic patterns have unveiled the great complexity of chromosomal evolution and rearrangements in important Cucurbitaceae crops. In pumpkin, six chromosomes remained in the ancestral karyotype state, whereas all chromosomes of cucumber and watermelon were formed through many fusions and fissions [37]. The collinearity analysis showed that, of the five chromosomes in cucumber, each had a one-to-two syntenic relationship with 10 of the melon chromosomes [12]. Four chromosomes in snake gourd each virtually had a one-to-one syntenic relationship with sponge gourd chromosomes [45], which indicates that they are closely related to each other. Overall, this information is of fundamental importance for comparative genomics in cucurbits.

Genome evolution

Besides the whole-genome triplication event (gamma) shared by all eudicots, it seems that at least four additional whole-genome duplication (WGD) events occurred during the evolution of Cucurbitaceae plants [2]. An early large-scale duplication event (CucWGD1), a cucurbit-common tetraploidization at the origin of the Cucurbitaceae family, has been identified, which occurred shortly after the gamma event (115–130 Mya) [2, 46, 47]. Moreover, three relatively recent WGDs have been identified within three tribes [2]. The tribe Cucurbiteae probably experienced one WGD at its origin (CucWGD2) [34]. Several studies have shown that zucchini (Cucurbita pepo), pumpkin (Cucurbita moschata and Cucurbita maxima), and silver-seed gourd (Cucurbita argyrosperma) from the tribe Cucurbiteae underwent WGD events [29, 34, 35]. In addition, several members of the tribe Sicyoeae also exhibit evidence for one WGD event (CucWGD3) [2]. One recent WGD event occurred in chayote (Sechium edule) of the tribe Sicyoeae at about 25 ± 4 Mya [46]. CucWGD4 is likely shared by the members of Hemsleya in the Gomphogyneae tribe [2].

According to the evolutionary relationship among Cucurbitaceae [26, 28, 30, 34, 37, 39, 42, 4446], we summarized the phylogeny of 17 sequenced Cucurbitaceae species (Cucumis × hytivus is not included) by integrating relevant information (Fig. 1). Phylogenetic analysis indicates that a variety of fruit and vegetable Cucurbitaceae crops emerged with different shapes due to species differentiation after the first shared WGD event. Among these 17 species, the first divergent species appears to be luo-han-guo followed by bitter gourd [2, 3437]. The Sicyoeae branch containing sponge gourd, ridge gourd, snake gourd, and chayote diverged sequentially [45, 46]. Species belonging to Benincaseae and Cucurbiteae form the sister clades. The Benincaseae tribe is represented by the four successively divergent genera of Cucumis, Benincasa, Lagenaria, and Citrullus [34, 37], and the Cucurbiteae tribe is represented by four Cucurbita species with the sister pairs of Cucurbita moschata and Cucurbita argyrosperma grouped with Cucurbita pepo and Cucurbita maxima in succession [34, 35] (Fig. 1). Although the divergence times among these cucurbit crops have been estimated using a Bayesian method [2], the exact divergence time of each Cucurbitaceae species remains unclear.

Figure 1.

Figure 1

Phylogenetic tree based on the reported phylogenetic relationships of 17 Cucurbitaceae species. Orange stars indicate cucurbit-specific WGD (CucWGD) events. Tribes are listed to the right of the tree.

The history of the speciation events has been reported in several studies on genomic research and sometimes there is a conflict of the estimated species divergence time [30, 3437, 45, 46], which may be affected by the species representativeness, method, fossils, and confidence interval used in estimating the time. For example, the divergence between cucumber and melon has been variously estimated at 8.4–11.8 (the median value is 10.1) Mya [23, 30, 37, 48]. However, according to the research of Ma et al. [45], Fu et al. [46], and Sun et al. [34], the two species (cucumber and melon) diverged ~5–12, 4–14, and 6.06–6.94 Mya, respectively. Therefore, estimates of the divergence between cucumber and melon range from about 4 to 14 Mya. This information is summarized in Table 2.

Table 2.

Reported estimates of divergence and evolution of members of the Cucurbitaceae.

Cucurbitaceae species Divergence time (Mya) Reference Comprehensive viewpoint (Mya)
Cucumis sativus and Cucumis melo 5–12 45 4–14
8.4–11.8 30
4–14 46
10.1 37
6.06–6.94 34
12.59 28
9 44
8.2–10.0 42
9.63 40
6.8 39
9.6 26
10 30
Cucumis hystrix and Cucumis sativus 4.5 26 4.5
Cucumis metuliferus and Cucumis melo 17.85 28 17.85
Citrullus lanatus and Lagenaria siceraria 14–27 45 10–30
10.4–14.6 30
12–30 46
13.8 37
18.34–19.75 34
16 28
13.0–15.7 42
15.11 40
14.9 39
21.4 26
10–14 30
B. hispida and Citrullus lanatus/Lagenaria siceraria 16.3 37 16.3–18.1
18.1 40
16.8 39
Siraitia grosvenorii and Citrullus lanatus 40.9 44 40.9
Momordica charantia and other 48–77 45 29.2–96
29.2–41 30
39–96 46
36.1 37
44.1 ± 14 35
34.94–37.24 34
33.3–40.4 42
44.06 40
41.6 39
49.8 26
Cucurbita maxima and Cucurbita moschata 3.04–3.84 34 3.04–7.3
5.3–7.3 42
4.81 40
Cucurbita maxima and Cucurbita pepo 13.9 28 13.9
Cucurbita moschata and Cucurbita pepo 3.9–5.4 42 3–16
3–16 46
3–13 45
6.3 39
Cucurbita moschata and Cucurbita argyrosperma 3.98 ± 1.7 35 3.98 ± 1.7
T. anguina and Sechium edule 27–45 46 27–45
Luffa acutangula and Luffa cylindrica 7.97 40 7.97
Luffa cylindrica and T. anguina 33–47 45 29–55
29–55 46

Genes associated with important agronomic traits

With the development of the whole-genome sequences of Cucurbitaceae, a large number of coding genes have been annotated and genes related to fruit and vegetable quality traits have begun to be identified. A wide range of important phenotypic and agronomic traits of Cucurbitaceae plants include pathogen resistance, fruit size, mass, color, texture, length, shape, rind form, ripening behavior, sugar content, bitterness, flavor and aroma, sex determination, and tendrils [12, 24, 32]. Population analysis and genome-wide association studies (GWAS) on diverse species accessions has contributed to the identification of a number of candidate genes controlling desirable fruit and vegetable traits [24]. This provides information for effective breeding strategies and is conducive to the development of high-quality, resilient elite cultivars of Cucurbitaceae species [23, 25].

Resistance genes

Plant resistance (R) genes are among the most important targets for plant breeding programs and have been the object of intense research. R genes can activate plant defense systems to restrict pathogen invasion and improve plant resistance against major diseases [49]. The major resistance genes have been identified in various Cucurbitaceae species. Among these genes, those encoding the nucleotide-binding site leucine-rich repeat (NBS-LRR) proteins are related to effector-triggered immunity, which is a significant component of plant disease resistance [50]. A total of 44 NBS-LRR (NLR) genes, consisting of 26 coiled-coil (CC)-NBS-LRR (CNL) and 18 Toll interleukin receptor (TIR)-NBS-LRR (TNL) genes, have been identified in watermelon [31]. The number of NLR genes identified in cucumber, melon, wax gourd, Cucurbita maxima, Cucurbita moschata, and bitter gourd are 74, 84, 82, 30, 57, and 78, respectively [34, 37].

Research on the aphid resistance of cucumber cultivar ‘EP6392’ showed that 8 of the 49 DEGs may be relevant to aphid resistance [51]. The volatile (E,Z)-2,6-nonadienal (NDE) is involved in resistance to a number of bacteria and fungi in cucumber [52]; several EIF4E and EIF4G genes were found to be resistant to plant RNA virus infections, and two At (glyoxylate aminotransferase) gene homologs conferring potential resistance to downy mildew have also been identified [12]. Interestingly, an EIF4E gene found in melon mediates recessive resistance against melon necrotic spot virus [5355], and the increased expression of two glyoxylate aminotransferase (At1 and At2) genes was found in wild melon genotypes, which may contribute to their resistance to downy mildew [56].

The most prevalent viruses that have a significant impact on the production of cucurbit crops are aphid-transmitted viruses in the Potyviridae family, including papaya ring-spot virus watermelon strain (PRSV-W), zucchini yellow mosaic virus (ZYMV), and watermelon mosaic virus (WMV) [5764]. Of these, PRSV-W is one of the most destructive viruses that infect cucurbits worldwide [6567]. The bottle gourd USVL5VR-Ls line is resistant to PRSV-W [68], and resistance is determined by Prs, an unidentified dominant monogenic locus [30]. An NBS-LRR gene (RGH10) was shown to confer PRSV resistance in melon [69]. Research has showed that ethylene signaling may participate in the PRSV resistance mechanism in cucurbits. AP2/ERF transcription factors (TFs) have been reported as the basis of plant defense mechanisms against a wide range of pathogens, including viruses, which makes the AP2/ERF gene family a feasible source of candidate genes for Prs [70]. In the snake gourd genome, five R genes potentially involved in the plant–pathogen interaction pathway have been identified [45]. Changes in their expression are associated with the changes in resistance during fruit ripening, which may possibly be related to the resistance of snake gourd to pathogens and insects [45].

Sex determination

In Cucurbitaceae, sex determination is closely related to fruit earliness, yield, and quality [71]. Ethylene stimulates femaleness and is regarded as the main regulatory factor of sex determination [7274]. Naturally occurring mutations in the genes encoding the corresponding enzymes in the ethylene biosynthesis pathway have a notable impact on sex determination in the Cucurbitaceae [75, 76]. For example, a loss-of-function mutation in a 1-aminocyclopropane-1-carboxylic acid synthase (ACS) gene in melon and cucumber leads to the enhancement of ‘maleness’ [75, 77]. There seem to be similar mechanisms at play in Cucurbita pepo [78]. In addition, the ACC oxidase gene CsACO2 is essential in female flower formation in cucumber and mutations in this gene confer androecy [79]. Studies have also shown that ethylene receptors are implicated in the regulation of zucchini sex determination [80, 81]. Cucumber and melon are often used to study sex expression in Cucurbitaceae plants [12, 23, 82]. Three major sex determination genes, M, F, and A, have been established in cucumber, and shown to be members of the aminocyclopropane-1-carboxylic acid synthase (ACS) gene family (CsACS1G for F, CsACS2 for M, and CsACS11 for A) [77, 79, 8386]. Cucumber has a distinctive genetic system for gynoecious sex expression and contains three genes: CsACS1, CsACS1G, and CsMYB [8790]. Study has revealed that the CsACS1G gene is responsible for production and development of female flowers in cucumber gynoecy conferred by the F locus [91]. However, this gynoecy expression system appears to be unstable, which may be due to unequal crossing over at the copy number variation (CNV)-based femaleness (F) locus [87]. The melon sex determination-related gene Cm-ACS7 and ACS11, and cucumber ortholog Cs-ACS2, as well as Cucurbita pepo ortholog CpACS27A, are crucial regulatory enzymes in the ethylene biosynthetic pathway [77, 9295]. These genes are vital to the suppression of male organs and development of the female flower [77, 9295]. In addition, the gynoecious locus CmWIP1 involved in occurrence of gynoecy in melon has also been found to be implicated in sex determination of cucurbits [9699]. It has two orthologous genes (CpWIP1A and CpWIP1b) identified in Cucurbita pepo [92]. In addition, auxin can regulate sex expression through stimulating ethylene generation [7274]. Research has suggested that six auxin-related genes and three short-chain reductase or dehydrogenase genes involved in sex determination have higher expression levels in unisexual flowers of cucumber [12]. The identification and functional analysis of these genes have provided valuable information for the study of sex expression in other Cucurbitaceae plants.

Fruit color

The diverse color of fruit is determined by the concentrations and compositions of various pigments, mainly chlorophylls and carotenoids, as well as flavonoids (especially chalcones and anthocyanins). Melon rinds have a variety of colors, including green, white, orange, yellow, variegated, and striped [100]. It is known that β-carotene accumulation can contribute to the orange color, and the accumulation of lutein and other carotenoids contributes mainly to the yellow color of fruit [101], while the carotenoid content of white-fleshed melon and watermelon can be low or negligible [102, 103]. A yellow flavonoid pigment, naringenin chalcone, was identified as the major pigment in mature rinds of ‘canary yellow’ type melons [100]. Similarly, the main carotenoids that accumulate in yellow-fleshed watermelon and zucchini are lutein and β-carotene [104].

The key genes known to be implicated in the carotenoid metabolic pathway play important roles in regulating carotenoid accumulation, leading to changes in pigmentation [105]. The CmPPR1 (EVM0014144) gene may affect carotenoid accumulation and flesh color in melon [106108]. The CmOr gene controls β-carotene accumulation, resulting in the orange flesh colors in melon fruit [109], while the identified MELO3C003097 gene may serve as a strong candidate for the Wf locus controlling white and green melon flesh [25]. Moreover, two peel-color-related candidate genes, MELO3C003375 and EVM0012228 (CmKFB), have been identified [25]; CmKFB genes negatively regulated the accumulation of naringenin chalcone determining the yellow color of melon rind [110]. The flesh color of Cucurbita moschata and Cucurbita maxima usually appears to be yellow and orange, while zucchini is mainly white and pale yellow [111114]. β-Carotene hydrolase (CHYB) and phytoene synthase (PSY) are two main genes affecting the formation of yellow-fleshed fruit of Cucurbita moschata, Cucurbita maxima, and Cucurbita pepo [113115], while the carotenoid cleavage dioxygenases 4 (CCD4) gene exerts an important function in the regulation of white pulp in Cucurbita pepo [112]. Ripe M. charantia fruits had higher carotenoid (mainly β-carotene) concentrations [116]. During fruit ripening, increased expression of phytoene synthase (McPSY) and phytoene desaturase (McPDS), associated with carotenoid synthesis, was observed, resulting in carotenoid accumulation in the pericarp and a change of peel color from green to orange [116, 117]. A study in Cucurbita pepo showed that the up-regulated expression of several structural genes involved in carotenoid metabolic pathways probably leads to the increased carotenoid accumulation in ripe fruit [92]. It is well known in tomato that PSY1 is a critically important enzyme that is induced during ripening [118, 119]. In ripening fruit of sweet watermelon, the PSY1 gene may be involved in the transition from pale-colored to red, orange, or yellow flesh through increasing total carotenoid accumulation [32]. Mutation in LCYB may lead to increased lycopene content, since artificial selection of the mutation was shown to be responsible for the red flesh color in most sweet watermelon cultivars [32]. Moreover, ClTST2, a sugar transporter gene, was credited with facilitating carotenoid accumulation in watermelon fruit flesh [32]. During flesh color formation, the up-regulated expression of gene ClPHT4;2 was closely related to increased carotenoid contents in watermelon flesh [107, 120]. In chayote fruit, a number of candidate genes regulating pigment accumulation have also been identified, such as HCAR (7-hydroxymethyl chlorophyll a reductase), regulating chlorophyll content, and β-carotene hydroxylase 2 (CHY2), CCD1, CCD4, and ZEP [46]. These genes may be involved in fruit color production [46]. The up-regulated expression of carotenoid accumulation-related genes may contribute to the increase of carotenoid content, making the fruit turn orange-red after ripening in snake gourd fruit [45].

Fruit size, shape, and texture

There are many factors that affect the formation of fruit shape, and their interaction and coordination eventually lead to differences in fruit shape. Various studies have reported a variety of classical and newly identified key genes related to fruit shape, mainly including SUN, OFP, WOX, YABBY, AP2, and auxin transporters [8, 121124]. Apart from these well-known genes, sugar signaling and metabolism have been suggested to be related to cell division and growth, which can influence organ shape [125]. Through GWAS analysis, a strong association signal related to fruit shape in watermelon was identified near the ClFS1 (Cla97C03G066390) gene controlling fruit elongation [126]. In addition, other genes or proteins related to fruit shape are also found in different plants, such as the TONNEAU1 recruiting motif protein (TRM5), the AP2/ERF transcription factor (AP2a) gene in tomato [127, 128], and the CAD1 gene belonging to the LRR-RLK family in peach [129].

The fruit shapes of Cucurbitaceae plants are diverse, and some genes controlling their shape variation have been identified. Quantitative trait locus (QTL) analysis for cucumber showed that the round fruit shape in WI7239 is controlled by two QTLs, FS2.1 and FS1.2, containing the tomato homologous genes SlTRM5 (CsTRM5) and SUN (CsSUN25–26-27a), respectively [128, 130]. The deletion of the rst exon of FS1.2 in cucumbers results in the formation of round fruits [130]. In another study, FS5.2 greatly influenced the formation of round fruit in WI7167 cucumber [130]. Watermelon fruits have three major shapes: elongate (OO), oval (Oo), and spherical (oo), controlled by a single, incompletely dominant gene [126]. A candidate gene, Cla011257, on chromosome 3 related to watermelon fruit shape (ClFS1) was identified and results suggested that Cla011257 might control spherical fruit shape and a deletion of 159 bp in Cla011257 may lead to elongated fruit in watermelon [126]. The wax gourd fruit shapes are mainly long cylindrical, cylindrical, and round [131]. During ovary formation, the expression levels of Bch02G016830 (designated BFS) in round wax gourd fruit are significantly higher than in long cylindrical fruits [131]. Therefore, BFS might be a candidate gene for fruit shape in wax gourds [131]. Variations in BFS might slow down cell division at the ovary formation stage and may contribute to the regulation of wax gourd fruit size [131]. In Cucurbita pepo, a single gene, Di, controls the disk fruit shape, which is dominant over spherical or pear-shaped fruit [132]. In Cucurbita moschata, the gene Bn controls butternut fruit shape and is dominant to bn for crookneck fruit shape [133]. In addition, sex expression has pleiotropic effects on cucumber and melon fruit shape [140, 134]. A 14-bp deletion in CsACS2, the candidate gene for the monoecious (m) locus in cucumber, resulted in elongated fruit shape in cucumber [95]. The pleiotropic effect of sex expression on fruit shape is also well established in melon [134].

Plant hormones have been showed to contribute to the regulation of fruit size and development [135]. Ethylene participates in many plant development processes and it serves as a triggering signal to initiate climacteric fruit ripening [136, 137]. The CpACS27A gene in Cucurbita pepo is the homologous gene of CmACS7 (MELO3C015444) in melon, which is involved in ethylene synthesis and sex determination and also influences fruit length [75, 108]. Auxin plays a critical role in cell expansion during fruit development stages [138140] and the role of the main regulators of auxin—auxin response factors (ARFs)—in cell division and growth have been well established [140, 141]. A total of 56 ARF genes were identified in bitter gourd [142], but in other families the number can vary considerably. It has been suggested that auxin-responsive GH3 family genes, auxin-responsive protein (IAA), and SAUR family proteins may be associated with chayote fruit enlargement [46], and the up-regulated expression of auxin-related genes may be involved in snake gourd fruit elongation [45]. SAUR was reported to be implicated in the regulation of plant growth and development through promoting cell expansion [143145], Bhi10G001538 and Bhi10G000196 may be important candidate genes contributing to large fruit during wax gourd domestication [37], and Bhi10G000196 is orthologous to the tomato gene SlFIN (Solyc11g064850) responsible for enlarged tomato fruit [146]. In addition, four WUSCHEL TFs have been identified in Cucurbita pepo [92], which affect fruit size [147, 148].

During fruit growth, development and ripening, there are many changes to cell wall structure and properties in cell wall biogenesis and modification, cell expansion, unidirectional elongation, and fruit softening [149, 150]. Numerous different types of cell-wall-modifying enzymes have been identified as being involved in the development and ripening processes of many fruits, including the pectin-modifying enzymes [polygalacturonase (PG), pectinesterase (PE), pectate lyase (PL), and β-galactosidase (β-GAL)] and the hemicellulose/cellulose-modifying enzymes [β-1,4-glucanase, xyloglucan transglycosylase/hydrolase (XTH) and expansin (EXP)], which together lead to changes in fruit texture by regulating the structure of cell wall polymers and influence fruit ripening [137, 151]. The increased expression of β-1,4-glucanase or enhanced enzyme activity is usually associated with fruit softening [75]. In addition, six genes (three pectinesterase genes, two gibberellin 20 oxidase 1-B-like genes, and one pectate lyase-like gene) involved in cell wall biosynthesis have been identified that may play important roles in determining epidermis thickness in the melon [24]. Regulation of the expression of many DEGs related to cell wall modification may be associated with fruit texture changes in snake gourd, including β-galactosidase 10/5-like, cellulose synthase-like protein, endoglucanase 10/11/17-like, expansin-A4/A10-like, β-glucosidase 18-like, and pectinesterase 53 [45]. Moreover, polygalacturonase, pectinesterase, and cellulose synthase-like protein B4 may affect cell wall properties and fruit texture during chayote development [46]. Expansins are cell wall proteins regulating cell size and fruit growth in plants, and are also highly expressed during fruit development and ripening [152, 153]. Although they have no catalytic activity, the expansins appear to induce loosening of bonds between cellulose and hemicellulose in the cell wall, leading to ‘polymer creep’ within the cell wall during growth, resulting in cell enlargement or shape change [137]. Also, expansins enable cell expansion and fruit softening by triggering the loosening of the cell wall [154]. Expansin-A12 is thought to be implicated in melon fruit size [24], and expansin-like B1, identified in the chayote fruit, may induce plant cell wall extension, with increased transcripts contributing to rapid fruit enlargement [46].

Other genes involved in cell division and cell cycle regulation can also directly influence the growth rate of plant tissues and determine the final size of plant organs [155]. A total of six DEGs related to the regulation of cell division and the cell cycle were identified in bottle gourd [156]. Furthermore, the study of melon showed that L-ascorbate oxidase (AAO) could play a role in the late stage of fruit development, associated with the change in fruit size [157]. Differential expression of the gene was also found in Cucurbita pepo [92], snake gourd [45], and chayote [46]. In Cucurbita pepo, up-regulated expression of the CpOVATE gene acting as a repressor of growth was observed in the small-fruit ‘Munchkin’, which showed that OVATE plays a key role in shorter fruit [158]. Similarly, the hexokinase (CpHXK-1) and CpFW2.2 genes were also found to contribute to a reduction in fruit size [158].

Fruit taste

There are three major components, including acidity, sugar, and volatile flavor compounds, that together contribute to the overall taste of fleshy fruit [159]. The PH gene (CmPH) identified in melon has an important regulatory effect on fruit acidity [159], and numerous genes involved in the citrate acid cycle that may influence the accumulation of organic acids have also been identified in melon [160]. The ClBt gene in watermelon and CsBt in cucumber regulate fruit bitterness [5, 32, 161] and volatile (E,Z)-2,6-nonadienal (NDE) confers on cucumber its ‘fresh green’ flavor [162], while CmTHAT1 (thiol acyltransferase, EVM0016460) affects fruit flavor [24,108].

Sugar accumulation is the main factor that contributes to the sweet taste, which is particularly important in the fruit ripening process of melon and watermelon. Two candidate genes, EVM0015625 and EVM0019658, have been suggested to be responsible for sugar accumulation in melon and the β-glucosidase and α-l-fucosidase 2 genes are related to the synthesis and transportation of sugars [24]. In melon fruit, a total of 63 genes may be involved in the sugar metabolism pathway [23], and enzymes considered to be involved in regulating sugar biosynthesis, unloading, transport, and metabolism processes during watermelon flesh development include neutral invertase, α-galactosidase, sucrose phosphate synthase, insoluble acid invertase, soluble acid invertase, UDP-glucose 4-epimerase, and UDP-galactose/glucose pyrophosphorylase [31]. An alkaline α-galactosidase gene (ClAGA2) was suggested to be related to the accumulation of sugar in watermelon pulp by promoting the metabolism of raffinose into glucose, fructose, and sucrose [32, 163165]. The roles of vacuolar sugar transporter ClVST1, hexose transporter ClSWEET3, and tonoplast sugar transporter ClTST2 in the sugar accumulation of watermelon fruit are well established [165]. ClVST1 is responsible for glucose and sucrose efflux and unloading in the watermelon fruit [166]. The key transporter protein ClTST2 contributes to the accumulation of sucrose, fructose, and glucose in the vacuole of watermelon fruit cells [167]. Their expression levels are positively correlated with watermelon fruit sugar content and their overexpression increased fruit sugar accumulation of watermelon flesh [168]. In addition, the overexpression of an ortholog of ClTST2 (CmTST2) in melon fruit could increase sugar content [168]. TF genes putatively implicated in sugar accumulation include a bZIP gene, namely Cla014572, which functions as a key regulatory factor of sugar accumulation during fruit development [31, 169]. Further work on the identification, differential expression, and functional analysis of these genes will contribute to the understanding of fruit flavor of Cucurbitaceae plants.

The catabolism of several amino acids plays a central role in the production of aroma compounds in melon [170]. Valine, leucine, and isoleucine are implicated in the biosynthesis of branched-chain esters [171], and tyrosine and phenylalanine participate in the biosynthesis of aromatic esters [172]. Ethylene can enhance the levels of these amino acids to promote synthesis of esters, thus affecting melon flavor [170], and ethylene may also enhance aminotransaminase (AT) activity by increasing the expression of CmBCAT1 and CmArAT1, whose gene products convert branched chain amino acids into aroma volatiles through amino acid aminotransferases [172, 173]. The key role of the two genes in the biosynthesis of melon aroma volatiles is well documented [172]. Sulfur-containing aroma volatiles make an important contribution to the distinctive aroma of melon and other fruits [173] and thioether esters greatly promote the fruity aroma of melon fruit [174, 175]. l-Methionine was postulated to be a precursor of aroma volatiles in melon fruit [175]. Two distinct parallel pathways for l-methionine catabolism, a transamination route involving the action of an l-methionine aminotransferase and a γ-lyase route involving the action of an l-methionine-γ-lyase activity encoded by melon gene CmMGL is involved in the formation of melon aroma volatiles [173]. In addition, sulfur-containing esters may also be synthesized from cysteine [170].

The cucurbitacins are plant triterpenoids that form the bitter compounds predominant in the Cucurbitaceae family and impart a bitter taste in cucumber, zucchini, melon, pumpkin, and other plant foods [5, 161, 176]. To date, many cucurbitacins, including cucurbitacins A–L, O–T, and several others, have been discovered in plants (https://en.wikipedia.org/wiki/Cucurbitacin). Several studies have shown that they exhibit wide-ranging pharmacological activities, such as cytotoxic, hepatoprotective, purgative, anti-inflammatory, anti-infectious, antidiabetic, antitumour and anticancer effects [177180]. In addition, cucurbitacin I can suppress cell motility through interfering indirectly with actin dynamics [181]; cucurbitacin B and cucurbitacin I could be beneficial in suppressing adipocyte differentiation and preventing metabolic diseases [182]; and the efficacy of cucurbitacin R and dihydrocucurbitacin B on the immune system has also been recognized [183].

The precursors of cucurbitane triterpenoids are synthesized through the mevalonate pathway [184] and cucurbitadienol is produced by cucurbitadienol synthase, forming the basic skeleton of cucurbitane triterpenoids [185] (Fig. 2). Cucurbitacins C (CuC), B (CuB), and E (CuE) are the main bitter substances isolated from cucumber [5], melon [186], and watermelon [187], respectively. The biosynthesis pathway of CuC has been described by Shang et al. [5]; nine CuC biosynthetic enzymes (CsBi, seven CYPs, and CsACT) were identified and four catalytic steps were elucidated. Eight CuB (CmBi, six CYPs, and CmACT) and 10 CuE biosynthetic enzymes (ClBi, 8 CYPs, and ClACT) have also been identified in melon and watermelon, respectively [161]. The cucurbitacin biosynthetic enzymes (Bi, eight CYPs, and ACT) have also been identified in Luffa acutangula and Luffa cylindrica [39]. The biosynthesis pathway of cucurbitane triterpenoid in bitter gourd was reported by Cui et al. [42]. The identification of these bitter genes has contributed to understanding the regulatory and biochemical variations of cucurbitacins and provided important information for molecular breeding for taste improvement.

Figure 2.

Figure 2

The cucurbitacin biosynthesis pathway in melon, cucumber, and watermelon.

Transcription factors involved in fruit growth and ripening

Many TF families have important effects on fruit development [188190]. Myeloblastosis (MYB) proteins are one of the largest TF families in plants and are widely involved in diverse plant-specific processes, such as plant organ development, signal transduction, secondary metabolism, and multiple stress responses [191194]. In cucumber, two MYB genes, CsMYB6 (Csa3G824850) and CsTRY (Csa5G139610), have been reported to negatively regulate fruit spine or trichome initiation [195]. Other research has shown that the CsTRY not only regulates fruit spine or trichome formation, but also plays a negative regulatory role in anthocyanin synthesis [196]. Moreover, CsMYB60 is a key regulatory gene that determines fruit spine color in cucumber, and is a good candidate for the B (black spine) gene controlling the black fruit-spine trait, which regulates the pigmentation of black spines [197, 198]. A total of 162 MYB genes have been identified in watermelon [199].

The GRAS family constitutes one of the major plant-specific TF families that are related to plant growth, development, cell signaling, and stress tolerance [200]. It has been reported that a total of 237 GRAS genes were identified in six Cucurbitaceae crop genomes. The number of GRAS genes was little different among these species, including Cucumis sativus (37), Cucumis melo (36), B. hispida (35), Citrullus lanatus (37), and Lagenaria siceraria (37) [201, 202], while the number present in Cucurbita moschata (55) was considerably greater. It is known that silencing the SlGRAS2 gene can reduce fruit weight during tomato fruit development [203]. The study proposed that several genes homologous to SlGRAS2 (CmoCh09G009100.1, CmoCh01G012140.1, MELO3C018144T1) among these GRAS genes might potentially function in fruit development [203].

The NAC domain genes are also one of the largest TF families in plants [204]. A total of 81 genes encoding 92 proteins of the NAC-domain family have been identified in the melon genome [204, 205]. They play an important part in the regulation of fruit ripening in different plants and CmNAC-NOR, a melon NAC gene family member, is a homolog of tomato Nor gene (SlNAC-NOR), involved in the climacteric fruit ripening process [136, 204, 206]. The NAC gene SlNAC4 can influence carotenoid accumulation and ethylene synthesis and is a positive regulator of fruit ripening in tomato [206]. The precise roles of the crucial tomato ripening ‘master regulators’, including MADS-RIN, NAC-NOR, and SPL-CNR, have been re-evaluated and it turns out that their severe ripening-inhibition phenotypes result from gain-of-function mutations [136]. Nevertheless, in the wild type, these regulators, plus Nor-like1 and other MADS and NAC genes, together with ethylene, play major roles in changes in color, flavor, texture, and ripening progression through promoting the full expression of related genes [206, 207]. MADS-box genes have been reported to regulate fruit expansion and ripening processes in melon [205, 208]. In addition, there are many other TFs involved in the regulation of fruit ripening, including the positive regulators TAGL1 [209] and LeHB-1 [210] and the negative regulators LeERF6 [211] and LeAP2a [212].

Transcriptomics

Transcriptome analysis has become an effective approach to understanding the gene networks that govern quality and developmental processes (Fig. 3) and can aid in identifying and exploiting superior cultivars with desirable traits, thus accelerating the Cucurbitaceae plant-breeding process. The transcriptome sequences of many Cucurbitaceae plants, including Cucurbita maxima [35], Cucurbita moschata [35], Cucurbita argyrosperma [35], bottle gourd [30], wax gourd [37], chayote [46], snake gourd [45], watermelon [31, 32], and zucchini [29], are available in the Sequence Read Archive (SRA) database of NCBI. These data provide important information on protein-coding gene prediction, new gene discovery, and gene functional annotation. In addition, transcriptome sequencing has been employed to investigate the molecular basis of the development of many fleshy fruits in Cucurbitaceae species, including cucumber [152, 213215], melon [216218], watermelon [219], bitter gourd [150], Momordica cochinchinensis [220], bottle gourd [156], zucchini [92, 158, 221], pumpkin [222, 223], wax gourd [224], snake gourd [45], and chayote [46]. Many DEGs related to fruit quality have been identified and Table 3 shows the integrated gene information derived from the transcriptome data of Cucurbitaceae plants.

Figure 3.

Figure 3

Genes involved in fruit quality in Cucurbitaceae plants

Table 3.

Identified DEGs related to fruit quality from transcriptome data in Cucurbitaceae plants.

Cucurbitaceae Color-related genes Texture-related genes Aroma-, flavor-, and taste-related genes Plant hormone-related genes Key transcription factors
Cucumber (Cucumis sativus) β-carotene hydroxylase (BCH), β-carotene 3-hydroxylase (BCH), β-carotene isomerase (BISO), carotenoid cleavage dioxygenase 7 (CCD7), 9-cis-epoxycarotenoid dioxygenase (NCED) β-galactosidase (β-Gal), β-glucosidase (β-Glu), β-amylase 1/3 (BMY1/3), cellulase 3 (CL3), pectin lyase (PL), pectinesterase (PE), pectin methylesterase 3 (PME3), pectinacetylesterase (PAE), xyloglucan:xyloglucosyl transferase (XET5), expansin (EXP), cellulose synthase (CeSA) terpene synthase 21 (TPS21), lupeol synthase (LUS), glutamate dehydrogenase 2 (GDH2), lipooxygenase (LOX), phenylalanine ammonia-lyase 2 (PAL2), alcohol dehydrogenase 1 (ADH1), sucrose synthase 4/5 (SUS4/5), β-amyrin synthase (β-AS) 1-aminocyclopropane-1-carboxylate synthase 8/10 (ACS8/10), gibberellin-responsive protein, histidine phosphotransfer protein (HP), AUX1, Aux/IAA, small auxin-up-regulated RNA (SAUR), ent-kaurenoic acid (KAO), gibberellin 20-oxidase (GA20ox), gibberellin 2-oxidase (GA2ox), gibberellin insensitive dwarf 1 (GID1), auxin-responsive GH3 family protein (GH3), gibberellin-regulated family protein (GASA), auxin-responsive protein, auxin-induced protein 13, ethylene insensitive 3 (EIN3), ethylene response factor (ERF), gibberellin 2-oxidase 8 (GA2ox8) AP2/ERF, GRAS, HSF, LFY, MADS, NAC, WRKY, YABBY, Zinc finger protein, bZIP, MYB, TCP, WD40, bHLH, SBP, NF-YA, AUX/IAA
Melon (Cucumis melo) phytoene synthase (PSY), carotenoid hydroxylases (CYP97A3), 9-cis-epoxycarotenoid dioxygenases (NCED), abscisic acid 8′-hydroxylase (CYP707A) carotenoid isomerase (CRTISO), carotenoid hydroxylases (CYP97A3), 9-cis-epoxycarotenoid dioxygenases (CCD), abscisic acid 8′-hydroxylase (CYP707A) sucrose phosphate synthase 2 (CmSPS2), sucrose synthases (CmSUS1, CmSUS2 and CmSUS-LIKE1), hexokinases (CmHK2 and CmHK3), fructokinase 3 (CmFK3), acid invertases 2 (CmAIN2), cell wall invertases (CmCIN2 and CmCIN3), phosphoglucose isomerase cyt (CmPGIcyt), α-galactosidase 2 (GAL2), phosphoenolpyruvate carboxykinases (PCK1, PCK2 and PCK3), cytosolic NADP+-dependent isocitrate dehydrogenases (cICDH1 and cICDH2), lipoamide dehydrogenase 1 (LPD1), malate dehydrogenase (MD1, MD2), aconitate 1 (ACO1), 2-oxoglutarate dehydrogenase (OGDH), pyruvate dehydrogenase (PD) No data AP2/EREBP, Constans-like zinc finger, C2H2 zinc finger, GOLDEN2-like, MYB, bHLH, WRKY
Watermelon (Citrullus lanatus) phytoene synthase 1 (PSY1), lycopene β-cyclase (LCYB), 9-cis-epoxycarotenoid dioxygenase 5 (NCED5), β-carotene hydroxylase (BCH), carotenoid β-ring hydroxylase (CYP97A3), carotenoid cleavage dioxygenase (CCD) pectin methylesterase (PME), pectinesterase (PE), proline-rich proteins (PRPs), fasciclin-like arabinogalactan proteins (FLAs), xyloglucan endotransglycosylases (XETs), early nodulin-like proteins (ENODs), S-adenosyl methionine decarboxylase (SAMDC), β-D-glucosidase (β-Glu), β-galactosidase (β-Gal), cellulose synthase (CeSA), endo-1,4-β-glucanase (EG), endoglucanase (EG) delta7 sterol C-5 desaturase (5-DES), phenylalanine ammonia lyase (PAL), pyruvate decarboxylase (PDC), malate dehydrogenase (MD), sucrose synthase (SUS), sucrose-phosphate synthase (SPS), ascorbate peroxidase (APX), squalene synthase (SQS) ethylene response factor 1 (ERF1), gibberellin 20-oxidase (GA20ox), Gibberellin regulated protein (GASA), auxin-repressed protein ARP1 (ARP1), abscisic acid response protein, Aux/IAA protein, auxin response factor 2 (ARF2), auxin-repressed protein (ARP) AP2/ERF, bZIP, MADS, MYB, NAC, WRKY
Bitter gourd (Momordica charantia) phytoene synthase (PSY), zeta-carotene desaturase (ZDS), 15-cis-phytoene desaturase (PDS), lycopene β-cyclase (LCYB), lycopene epsilon-cyclase (LCYE), prolycopene isomerase (CRTISO), carotene epsilon-monooxygenase (CYP97C1), zeaxanthin epoxidase (ZEP), violaxanthin de-epoxidase (VDE), zeta-carotene isomerase (ZISO), β-carotene 3-hydroxylase (BCH), xanthoxin dehydrogenase (ABA2), flavonol synthase (FLS), chalcone synthase (CHS) Pectinesterase (PE), pectate lyase (PL), β-galactosidase (β-Gal), β-amylase (BMY), β-glucosidase (β-Glu), α-mannosidase (MANA), polygalacturonase 2 (PG2) sucrose synthase (SUS), phosphoenolpyruvate carboxylase (PEPC), (3S)-linalool synthase (TPS14), lipoxygenase (LOX), alcohol dehydrogenase (ADH), glutamine synthetase (GS), glutamate decarboxylase (GAD), pyruvate decarboxylase (PDC), phenylalanine ammonia-lyase (PAL), β-amyrin synthase (β-AS) 1-aminocyclopropane-1-carboxylate synthase (ACS), gibberellin 2-oxidase (GA2ox), gibberellin 20-oxidase GA20ox, auxin-responsive protein IAA (IAA), auxin response factor (ARF), gibberellin receptor GID1 (GID1), abscisic acid receptor PYR/PYL (PYR/PYL), jasmonic acid-amino synthetase (JAR1), ethylene receptor (ETR), serine/threonine-protein kinase CTR1 (CTR1), ethylene-insensitive protein 2/3 (EIN2/3), ethylene-responsive transcription factor 2 (ERF2), small auxin up RNA (SAUR) AP2, AP2/ERF, Dof, NAC, WRKY
Bottle gourd (Lagenaria siceraria) No data endoglucanase (EG), expansin (EXP), galactoside 2-α-L-fucosyltransferase 2 (FUT2), galacturonosyltransferase-like 2 (GATL2), xyloglucan endotransglucosylase/hydrolase 2 (XTH2), xyloglucan glycosyltransferase 5 (CSLC5), xyloglucan endotransglucosylase/hydrolase protein (XTH) sucrose-phosphate synthase 4 (SPS4), sucrose synthase 5 (SUS5) auxin-binding protein ABP20, auxin-induced protein, auxin efflux carrier component 8, cytokinin dehydrogenase (CKX), gibberellin 3-β-dioxygenase 1 (GA3ox1) APRR2, bHLH, bZIP, ERF, MYB, MYC, RAD, TFIIB, WRKY
Zucchini (Cucurbitapepo) phytoene synthase (PSY), phytoene desaturase (PDS), β-carotene hydroxylase (BCH), carotenoid cleavage dioxygenase (CCD), carotenoid isomerase (CRTISO), zeta-carotene desaturase (ZDS), chalcone synthase 2 (CHS2), zeaxanthin epoxidase (ZEP) β-amylase 1 (BMY1), pectate lyase-like (PL), β-glucosidase (β-Glu), β-galactosidase (β-Gal), β-1,3-glucanase 3 (GLU3), expansin (EXP), endoxyloglucan transferase a2 (xrt2), pectinesterase (PE), pectate lyase 12 (PL12), polygalacturonase (PG), glucan endo-1,3-β-D-glucosidase (BGL2), cellulose synthase-like protein D5 (CSLD5) linoleate 9S-lipoxygenase 6 (LOX6), glutamate dehydrogenase 2 (GDH2), linoleate 13S-lipoxygenase 2–1 (LOX), alcohol dehydrogenase (ADH), granule-bound starch synthase (GBSS), sucrose synthase 5 (SUS5) 1-aminocyclopropane-1-carboxylate oxidase 3/4 (ACO3/4), ethylene insensitive 2/3 (EIN2/3), ethylene response 1/2 (ETR1/2), ethylene response factor (ERF), 1-aminocyclopropane-1-carboxylate synthase (ACS), gibberellin 2-oxidase (GA2ox), auxin-induced protein, auxin response factor (ARF), auxin-responsive protein IAA (IAA), gibberellin-regulated protein 9 (GASA9), auxin-induced protein AUX28, abscisic acid 8′-hydroxylase (CYP707A), auxin-responsive protein SAUR72, abscisic acid receptor PYL2, cytokinin hydroxylase-like MYB, bHLH, AUX/IAA, AP2, AP2/ERF, SBP, CAAT, HSF, MBF1, bZIP, NAC, MADS, GRF, WRKY
Pumpkin (Cucurbita maxima) phytoene synthase (PSY), 15-cis-phytoene desaturase (PDS), zeta-carotene isomerase (ZISO), prolycopene isomerase (CRTISO), lycopene epsilon-cyclase (LCYE), lycopene β-cyclase (LCYB), β-ring hydroxylase (CYP97A3), β-carotene 3-hydroxylase (BCH), carotene epsilon-monooxygenase (CYP97C1), zeaxanthin epoxidase (ZEP), violaxanthin de-epoxidase (VDE), chalcone synthase (CHS) endoglucanase (EG), expansin (EXP), xyloglucan endotransglucosylase/hydrolase protein (XTH), β-amylase 1 (BMY1), pectate lyase-like (PL), pectin acetylesterase (PAE), pectinesterase (PE), β-glucosidase (β-Glu), β-galactosidase (β-Gal), polygalacturonase (PG), glucan endo-1,3-β-glucosidase (BG), cellulose synthase (CeSA) alcohol dehydrogenase (ADH), β-fructofuranosidase (INV), fructokinase (FK), hexokinase (HK), glucose-6-phosphate isomerase (PGI), phosphoglucomutase (PGM), UTP-glucose-1-phosphate uridylyltransferase (UGPase), sucrose synthase (SUS), sucrose-phosphate synthase (SPS), ADP-sugar diphosphatase (NUDX14), glucose-1-phosphate adenylyltransferase (AGPase), 1,4-α-glucan branching enzyme (SBE), trehalose 6-phosphate synthase/phosphatase (TPS), α-amylase (amyA), 4-α-glucanotransferase (malQ), UDP-glucose 6-dehydrogenase (UGDH), UDP-glucuronate decarboxylase (UXS1), 1,4-β-D-xylan synthase (XS), β-D-xylosidase 4 (XYL4), UDP-glucuronate 4-epimerase (GAE), α-1,4-galacturonosyltransferase (GAUT) 1-aminocyclopropane-1-carboxylate oxidase (ACO), 1-aminocyclopropane-1-carboxylate synthase (ACS), ethylene insensitive 2/3 (EIN2/3), ethylene-responsive transcription factor (ERF), ethylene receptor 1 (ETR1), ethylene response sensor 1 (ERS1), gibberellin 2-oxidase (GA2ox), gibberellin 20-oxidase (GA20ox), auxin-induced protein, auxin response factor 6 (ARF6), auxin-responsive protein IAA (IAA), gibberellin-regulated protein (GASA), abscisic acid 8′-hydroxylase (CYP707A) bHLH, AP2/ERF, MYB, NAC, AUX/IAA, bZIP, WRKY
Snake gourd (T. anguina) phytoene synthase (PSY), 15-cis-phytoene desaturase (PDS), prolycopene isomerase (CRTISO), zeta-carotene desaturase (ZDS), lycopene β-cyclase (LCYB), lycopene epsilon cyclase (LCYE), 15-cis-zeta-carotene isomerase (ZISO), 9-cis-epoxycarotenoid dioxygenase NCED2 (NCED2), β-carotene 3-hydroxylase 1 (BCH1), carotene epsilon-monooxygenase (CYP97C1) cellulose synthase-like protein E1 (CSLE1), endoglucanase (EG), expansin (EXP), glucan 1,3-β-glucosidase (BGL), pectinesterase (PE), polygalacturonase (PG), β-galactosidase (β-Gal), glucan endo-1,3-β-D-glucosidase (BGL2), pectin acetylesterase (PAE), pectin methyltransferase (PMT), β-glucosidase (β-Glu), pectate lyase (PL), glucan endo-1,3-β-glucosidase (BG) linoleate 9S-lipoxygenase-like (LOX), linoleate 13S-lipoxygenase 2–1(LOX), alcohol dehydrogenase (ADH), glutamate dehydrogenase 1 (GDH1), phenylalanine ammonia-lyase (PAL), sucrose synthase (SUS) 1-aminocyclopropane-1-carboxylate oxidase (ACO), ethylene-responsive transcription factor (ERF), ethylene receptor 2 (ETR2), abscisic acid receptor PYR1-like, auxin-induced protein, auxin-responsive protein IAA, auxin response factor (ARF), abscisic acid receptor PYR1, auxin-responsive protein SAUR50-like, gibberellin-regulated protein (GASA), gibberellin receptor, abscisic acid 8-hydroxylase 4-like, gibberellin 2-β-dioxygenase 1-like, gibberellin 3-oxidase 3 (GA3ox3), auxin transporter-like protein (LAX), auxin-responsive protein SAUR71-like, ethylene insensitive 3 (EIN3), ethylene response sensor 1 (ERS1) AP2/ERF, MYC, ERF, ANT
Chayote (Sechium edule) phytoene synthase (PSY), 9-cis-epoxycarotenoid dioxygenase NCED2/3 (NCED2/3), zeaxanthin epoxidase (ZEP), carotenoid 9,10-cleavage dioxygenase 1 (CCD1), carotenoid cleavage dioxygenase 4/8 (CCD4/8), β-carotene hydroxylase 2 (BCH2), β-carotene isomerase D27 (D27), flavonol synthase (FLS), chalcone synthase 2 (CHS2) expansin (EXP), glucan endo-1,3-β-glucosidase (BG), polygalacturonase (PG), xyloglucan endotransglucosylase/hydrolase (XTH), β-galactosidase (β-Gal), β-glucosidase (β-Glu), cellulose synthase (CeSA), endoglucanase (EG), pectin acetylesterase (PAE), pectinesterase (PE), α-mannosidase (MANA), β-amylase 1 (BMY1), pectin methyltransferase (PMT), pectinesterase/pectinesterase inhibitor 25 (PME25), pectate lyase (PL) linoleate 13S-lipoxygenase 2–1 (LOX), linoleate 13S-lipoxygenase 3–1 (LOX), linoleate 9S-lipoxygenase (LOX), alcohol dehydrogenase-like 6 (ADH6), glutamate dehydrogenase 2 (GDH2), terpene synthase 10 (TPS10), sucrose-phosphate synthase 1 (SPS1), sucrose synthase (SUS), sucrose-phosphatase 1 (SPP1) 1-aminocyclopropane-1-carboxylate synthase (ACS),1-aminocyclopropane-1-carboxylate oxidase (ACO), ethylene-responsive transcription factor (ERF), abscisic acid 8-hydroxylase (CYP707A), auxin response factor (ARF), auxin-induced protein, auxin-responsive protein IAA, gibberellin receptor GID1B, abscisic acid receptor PYL8/9, auxin-responsive protein SAUR, gibberellin 20 oxidase 1 (GA20ox1), gibberellin 2-β-dioxygenase (GA2ox), gibberellin-regulated protein 14 (GASA14) AP2, MYB, NAC, WRKY, MYC, bHLH, SBP, AP2/ERF, bZIP, GRF

Importance of genome resequencing for the development of molecular breeding

Whole-genome resequencing technology has been used to investigate wide germplasm resources. Resequencing of multiple materials from different crop species has helped reveal the domestication history of cucurbit crops and candidate genes or loci influencing agronomic traits. Important cucurbit crops that have been resequenced include Citrullus lanatus [31, 32], Cucumis sativus [15], Cucumis melo [24, 25, 225], B. hispida [37], M. charantia [41], and Lagenaria siceraria [8]. Resequencing and provision of large-scale germplasm resources can be applied to population genomic analyses and GWAS to identify QTLs. Genome-wide single-nucleotide polymorphism (SNP) markers have been widely used in molecular breeding for mapping of important fruit quality trait genes and can contribute to the discovery of candidate loci or key genes and molecular markers associated with important traits in cucurbits for crop improvement (Table 4).

Table 4.

List of resequenced species of Cucurbitaceae plants.

Date Species Number of accessions Average sequencing depth SNPs Indels Reference
2012 Citrullus lanatus 20 5–16× 6 784 860 965 006 31
2013 Cucumis sativus 115 18.3× 3 305 010 336 081 15
2019 Cucumis melo 1175 4.71–18.92× 5 678 165 957 421 25
2019 Citrullus lanatus 414 14.5× 19 725 853 6 675 290 32
2019 B. hispida 146 15.68× 16 183 153 2 190 214 37
2019 Cucumis melo 50 18× 1 761 822 735 486 24
2020 M. charantia 60 - 6 135 286 41
2020 M. charantia 189 8.8–37.8 14 450 193 1 588 578 42
2021 Lagenaria siceraria 50 2 004 276 399 664 8
2021 Cucumis metuliferus 40 12× 1 688 333 247 969 28

A genome variation map for cucumber fruit was obtained through deep resequencing of 115 cucumber lines and a region containing a gene related to the loss of bitterness in cucumber fruit was identified [15]. The QTL mapping of cucumber also identified eight QTLs related to leaf size or fruit length [15]. Moreover, a natural genetic variant in a β-carotene hydroxylase 33 gene (CsaBCH1) that resulted in accumulation of β-carotene and formation of orange fruit endocarp was identified, which could be helpful in obtaining varieties with higher nutritional value [15]. In Payzawat melon, six structural gene variants potentially controlling the thickness of the epidermis were identified by analyzing the QTLs related to epidermis thickness [24]. In addition, Zhao et al. [25] reported a comprehensive map of the melon genomic variation that originated from the resequencing of 1175 accessions, and GWAS studies for 16 agronomic traits identified 208 loci markedly related to fruit quality, mass, and morphological characters. This study proposed that the strong differentiation between Cucumis melo and Cucumis agrestis may contribute to breeding. Watermelon breeding has mainly focused on fruit quality traits, particularly, sweetness, flesh color, and rind pattern, which has led to the narrow genetic base of watermelon [32]. In 2013, Guo et al. [31] resequenced 20 watermelon accessions and identified many disease-resistance genes that had been lost during domestication. Thus, improving resistance to pathogens is an ongoing goal of sweet watermelon breeding programs. Interestingly, Citrullus amarus, Citrullus colocynthis, and Citrullus mucosospermus have been used for breeding studies to find new sources of disease and insect resistance to improve sweet watermelon. Whole-genome resequencing of 414 accessions identified genomic regions associated with critical fruit quality traits and using GWAS identified a total of 43 association signals, which provided useful information for watermelon breeding [32].

Bitter gourd is an important vegetable and medicinal plant in the Cucurbitaceae family. The bitter taste of bitter gourd is due to the existence of cucurbit triterpenoid compounds cucurbitacins [42] and it has the potential for further improvement [41]. A total of 1507 marker loci were genotyped by using restriction-associated DNA tag sequencing (RAD-seq) analysis, resulting in an improved linkage map [6]. A total of 255 scaffolds were assigned to the linkage map through anchoring RAD tag markers [6]. Interspecific crosses play a vital part in Cucurbita breeding for transferring favorable traits between species [34], and 40 transcriptomes assembled for 11 species of the Cucurbita genus could serve as a valuable source of molecular markers [29]. In addition, research on the resequencing of 146 wax gourd accessions mapped nine QTLs for fruit-size-associated traits, and 11 candidate domestication genes and a number of genomic regions putatively related to the determination of wax gourd fruit size have been identified using GWAS [37]. These resequencing results and the genome sequences presented will provide a basis for DNA marker development, gene identification, and molecular breeding of these Cucurbitaceae species.

Future prospects

New-generation sequencing technologies and breeding techniques, together with bioinformatics tools, have greatly promoted the progress of plant breeding, although further bioinformatics analysis of whole-genome sequencing data of Cucurbitaceae crops is still required in order to accelerate Cucurbitaceae crop breeding improvement. The integration of multiomics data with genetic and phenotypic data will help to identify genes related to important traits and accelerate the process of plant breeding [71].

Genetic transformation and genome editing of Cucurbitaceae plants have a significant development potential for obtaining new cucurbit phenotypes with ideal traits. A reverse genetic approach, Targeting Induced Local Lesions in Genomes (TILLING), can be applied to the breeding of Cucurbitaceae crops and help to improve agronomic traits [226]. Different DNA mutant TILLING libraries have been set up in cucurbits [227231]. This approach has provided a resource for plant breeding programs and future functional genomics study. Genome editing technology is attracting attention and breeding efficiency can be rapidly improved through combining the genomic and variomic information on crops [232]. Developing efficient and reliable genetic transformation technology for the target crops will contribute to the wide application of this approach in Cucurbitaceae crops. CRISPR/Cas9 is a common and efficient technique for genome editing and has been used for Cucurbitaceae crops to knock out target genes and obtain crop materials with desirable agronomic traits [233, 234], such as cucumber [235, 236], watermelon [237, 238], and pumpkin [239], and has become a precision-breeding approach for modifying traits in plants species [240]. In the future, a wide range of genome analysis and editing research is expected to expand our understanding and implementation for Cucurbitaceae plant breeding programs.

Acknowledgements

This work was supported by grants from the Beijing Municipal Science and Technology Commission (Z191100008619004 and Z191100004019010), the National Natural Science Foundation of China (31772022 and 32072284), the Key Project ‘Science and Technology Boost the Economy 2020’, China Agriculture Research System of MOF and MARA (CARS-23), Special innovation ability construction fund of Beijing Academy of Agricultural and Forestry Sciences (20200427 and 20210437), Collaborative innovation center of Beijing Academy of Agricultural and Forestry Sciences (201915). The author thanks Professor Zhangjun Fei in Cornell University, Ithaca, NY, USA for reviewing this manuscript in his busy schedule and putting forward valuable guidance.

Conflict of interests

The authors declare that they have no conflict of interests.

References

  • 1. Schaefer H, Renner SS. Phylogenetic relationships in the order Cucurbitales and a new classification of the gourd family (Cucurbitaceae). Taxon. 2011;60:122–38. [Google Scholar]
  • 2. Guo J, Xu W, Hu Yet al. Phylotranscriptomics in Cucurbitaceae reveal multiple whole-genome duplications and key morphological and molecular innovations. Mol Plant2020;13:1117–33. [DOI] [PubMed] [Google Scholar]
  • 3. Schaefer H, Heibl C, Renner SS. Gourds afloat: a dated phylogeny reveals an Asian origin of the gourd family (Cucurbitaceae) and numerous oversea dispersal events. Proc R Soc B. 2009;276:843–51. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 4. Sui X, Nie J, Li Xet al. Transcriptomic and functional analysis of cucumber (Cucumis sativus L.) fruit phloem during early development. Plant J. 2018;96:982–96. [DOI] [PubMed] [Google Scholar]
  • 5. Shang Y, Ma Y, Zhou Yet al. Biosynthesis, regulation, and domestication of bitterness in cucumber. Science2014;346:1084–8. [DOI] [PubMed] [Google Scholar]
  • 6. Urasaki N, Takagi H, Natsume Set al. Draft genome sequence of bitter gourd (Momordica charantia), a vegetable and medicinal plant in tropical and subtropical regions. DNA Res. 2017;24:51–8. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 7. Devi N. Medicinal values of Trichosanthus cucumerina L. (snake gourd) - a review. Br J Pharm Res. 2017;16:1–10. [Google Scholar]
  • 8. Xu P, Wang Y, Sun Fet al. Long-read genome assembly and genetic architecture of fruit shape in the bottle gourd. Plant J. 2021;107:956–968. [DOI] [PubMed] [Google Scholar]
  • 9. Kersey PJ. Plant genome sequences: past, present, future. Curr Opin Plant Biol. 2019;48:1–8. [DOI] [PubMed] [Google Scholar]
  • 10. Arabidopsis Genome Initiative. Analysis of the genome sequence of the flowering plant Arabidopsis thaliana. Nature. 2000;408:796–815. [DOI] [PubMed] [Google Scholar]
  • 11. International Rice Genome Sequencing Project. The map-based sequence of the rice genome. Nature. 2005;436:793–800. [DOI] [PubMed] [Google Scholar]
  • 12. Huang S, Li R, Zhang Zet al. The genome of the cucumber, Cucumis sativus L. Nat Genet. 2009;41:1275–81. [DOI] [PubMed] [Google Scholar]
  • 13. Zhu C, Li X, Zheng J. Transcriptome profiling using Illumina-and SMRT based RNA-seq of hot pepper for in-depth understanding of genes involved in CMV infection. Gene. 2018;666:123–33. [DOI] [PubMed] [Google Scholar]
  • 14. D’Amore R, Johnson J, Haldenby Set al. SMRT gate: a method for validation of synthetic constructs on Pacific Biosciences sequencing platforms. Bio Techniques. 2017;63:13–20. [DOI] [PubMed] [Google Scholar]
  • 15. Qi J, Liu X, Shen Det al. A genomic variation map provides insights into the genetic basis of cucumber domestication and diversity. Nat Genet. 2013;45:1510–5. [DOI] [PubMed] [Google Scholar]
  • 16. Li Q, Li H, Huang Wet al. A chromosome-scale genome assembly of cucumber (Cucumis sativus L.). GigaScience. 2019;8:giz072. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 17. Wóycicki R, Witkowicz J, Gawroński Pet al. The genome sequence of the north-European cucumber (Cucumis sativus L.) unravels evolutionary adaptation mechanisms in plants. PLoS One. 2011;6:e22728. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 18. Yang L, Koo D-H, Li Yet al. Chromosome rearrangements during domestication of cucumber as revealed by high-density genetic mapping and draft genome assembly. Plant J2012;71:895–906. [DOI] [PubMed] [Google Scholar]
  • 19. Osipowski P, Pawełkowicz M, Wojcieszek Met al. A high-quality cucumber genome assembly enhances computational comparative genomics. Mol Gen Genomics. 2020;295:177–93. [DOI] [PubMed] [Google Scholar]
  • 20. Castanera R, Ruggieri V, Pujol Met al. An improved melon reference genome with single-molecule sequencing uncovers a recent burst of transposable elements with potential impact on genes. Front Plant Sci. 2020;10:1815. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 21. Yano R, Ariizumi T, Nonaka Set al. Comparative genomics of muskmelon reveals a potential role for retrotransposons in the modification of gene expression. Commun Biol. 2020;3:432. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 22. Shin AY, Koo N, Kim Set al. Draft genome sequences of two oriental melons, Cucumis melo L. var. makuwa. Sci Data. 2019;6:220. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 23. Garcia-Mas J, Benjak A, Sanseverino Wet al. The genome of melon (Cucumis melo L.). Proc Natl Acad Sci USA. 2012;109:11872–7. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 24. Zhang H, Li X, Yu Het al. A high-quality melon genome assembly provides insights into genetic basis of fruit trait improvement. iScience2019;22:16–27. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 25. Zhao G, Lian Q, Zhang Zet al. A comprehensive genome variation map of melon identifies multiple domestication events and loci influencing agronomic traits. Nat Genet. 2019;51:1607–15. [DOI] [PubMed] [Google Scholar]
  • 26. Qin X, Zhang Z, Lou Qet al. Chromosome-scale genome assembly of Cucumis hystrix—a wild species interspecifically cross-compatible with cultivated cucumber. Hortic Res. 2021;8:40. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 27. Yu X, Wang P, Li Jet al. Whole-genome sequence of synthesized allopolyploids in Cucumis reveals insights into the genome evolution of allopolyploidization. Adv Sci. 2021;8:2004222. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 28. Ling J, Xie X, Gu Xet al. High-quality chromosome-level genomes of Cucumis metuliferus and Cucumis melo provide insight into Cucumis genome evolution. Plant J. 2021;107:136–48. [DOI] [PubMed] [Google Scholar]
  • 29. Montero-Pau J, Blanca J, Bombarely Aet al. De novo assembly of the zucchini genome reveals a whole-genome duplication associated with the origin of the Cucurbita genus. Plant Biotechnol J. 2018;16:1161–71. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 30. Wu S, Shamimuzzaman M, Sun Het al. The bottle gourd genome provides insights into Cucurbitaceae evolution and facilitates mapping of a papaya ring-spot virus resistance locus. Plant J. 2017;92:963–75. [DOI] [PubMed] [Google Scholar]
  • 31. Guo S, Zhang J, Sun Het al. The draft genome of watermelon (Citrullus lanatus) and resequencing of 20 diverse accessions. Nat Genet. 2013;45:51–8. [DOI] [PubMed] [Google Scholar]
  • 32. Guo S, Zhao S, Sun Het al. Resequencing of 414 cultivated and wild watermelon accessions identifies selection for fruit quality traits. Nat Genet. 2019;51:1616–23. [DOI] [PubMed] [Google Scholar]
  • 33. Wu S, Wang X, Reddy Uet al. Genome of ‘Charleston Gray’, the principal American watermelon cultivar, and genetic characterization of 1,365 accessions in the U.S. National Plant Germplasm System watermelon collection. Plant Biotechnol J. 2019;17:2246–58. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 34. Sun H, Wu S, Zhang Get al. Karyotype stability and unbiased fractionation in the paleoallotetraploid Cucurbita genomes. Mol Plant. 2017;10:1293–306. [DOI] [PubMed] [Google Scholar]
  • 35. Barrera-Redondo J, Ibarra-Laclette E, Vázquez-Lobo Aet al. The genome of Cucurbita argyrosperma (silver-seed gourd) reveals faster rates of protein-coding gene and long noncoding RNA turnover and neofunctionalization within Cucurbita. Mol Plant. 2019;12:506–20. [DOI] [PubMed] [Google Scholar]
  • 36. Barrera-Redondo J, Sánchez-de la Vega G, Aguirre-Liguori JAet al. The domestication of Cucurbita argyrosperma as revealed by the genome of its wild relative. Hortic Res. 2021;8:109. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 37. Xie D, Xu Y, Wang Jet al. The wax gourd genomes offer insights into the genetic diversity and ancestral cucurbit karyotype. Nat Commun. 2019;10:5158. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 38. Zhang T, Ren X, Zhang Zet al. Long-read sequencing and de novo assembly of the Luffa cylindrica (L.) Roem. genome. Mol Ecol Resour. 2020;20:511–9. [DOI] [PubMed] [Google Scholar]
  • 39. Wu H, Zhao G, Gong Het al. A high-quality sponge gourd (Luffa cylindrica) genome. Hortic Res. 2020;7:128. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 40. Pootakham W, Sonthirod C, Naktang Cet al. De novo assemblies of Luffa acutangula and Luffa cylindrica genomes reveal an expansion associated with substantial accumulation of transposable elements. Mol Ecol Resour. 2021;21:212–25. [DOI] [PubMed] [Google Scholar]
  • 41. Matsumura H, Hsiao M-C, Lin Y-Pet al. Long-read bitter gourd (Momordica charantia) genome and the genomic architecture of nonclassic domestication. Proc Natl Acad Sci USA. 2020;117:14543–51. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 42. Cui J, Yang Y, Luo Set al. Whole-genome sequencing provides insights into the genetic diversity and domestication of bitter gourd (Momordica spp.). Hortic Res. 2020;7:85. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 43. Itkin M, Davidovich-Rikanati R, Cohen Set al. The biosynthetic pathway of the nonsugar, high-intensity sweetener mogroside V from Siraitia grosvenorii. Proc Natl Acad Sci USA2016;113:E7619–28. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 44. Xia M, Han X, He Het al. Improved de novo genome assembly and analysis of the Chinese cucurbit Siraitia grosvenorii, also known as monk fruit or luo-han-guo. GigaScience. 2018;7:1–9. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 45. Ma L, Wang Q, Mu Jet al. The genome and transcriptome analysis of snake gourd provide insights into its evolution and fruit development and ripening. Hortic Res. 2020;7:199. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 46. Fu A, Wang Q, Mu Jet al. Combined genomic, transcriptomic, and metabolomic analyses provide insights into chayote (Sechium edule) evolution and fruit development. Hortic Res. 2021;8:35. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 47. Wang J, Sun P, Li Yet al. An overlooked paleotetraploidization in Cucurbitaceae. Mol Biol Evol. 2018;35:16–26. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 48. Sebastian P, Schaefer H, Telford IRHet al. Cucumber (Cucumis sativus) and melon (C. melo) have numerous wild relatives in Asia and Australia, and the sister species of melon is from Australia. Proc Natl Acad Sci USA. 2010;107:14269–73. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 49. Pandolfi V, Neto JRCF, Silva MDet al. Resistance (R) genes: applications and prospects for plant biotechnology and breeding. Curr Protein Pept Sci2017;18:1–12. [DOI] [PubMed] [Google Scholar]
  • 50. Jones JDG, Vance RE, Dangl JL. Intracellular innate immune surveillance devices in plants and animals. Sci ence2016;354:1117. [DOI] [PubMed] [Google Scholar]
  • 51. Liang DN, Liu M, Hu Qet al. Identification of differentially expressed genes related to aphid resistance in cucumber (Cucumis sativus L.). Sci Rep2015;5:9645. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 52. Cho MJ, Buescher RW, Johnson Met al. Inactivation of pathogenic bacteria by cucumber volatiles (E,Z)-2,6-nonadienal and (E)-2-nonenal. J Food Prot. 2004;67:1014–6. [DOI] [PubMed] [Google Scholar]
  • 53. Nieto C, Morales M, Orjeda Get al. An eIF4E allele confers resistance to an uncapped and nonpolyadenylated RNA virus in melon. Plant J2006;48:452–62. [DOI] [PubMed] [Google Scholar]
  • 54. Nieto C, Piron F, Dalmais Met al. EcoTILLING for the identification of allelic variants of melon eIF4E, a factor that controls virus susceptibility. BMC Plant Biol. 2007;7:1–9. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 55. Rodríguez-Hernández AM, Gosalvez B, Sempere RNet al. Melon RNA interference (RNAi) lines silenced for cm-eIF4E show broad virus resistance. Mol Plant Pathol2012;13:755–63. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 56. Taler D, Galperin M, Benjamin Iet al. Plant eR genes that encode photorespiratory enzymes confer resistance against disease. Plant Cell2004;16:172–84. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 57. Ali A, Abdalla O, Bruton Bet al. Occurrence of viruses infecting watermelon, other cucurbits, and weeds in the parts of southern United States. Plant Heal Prog2012;13:9. [Google Scholar]
  • 58. Turechek WW, Kousik CS, Adkins S. Distribution of four viruses in single and mixed infections within infected watermelon plants in Florida. Phytopathology. 2010;100:1194–203. [DOI] [PubMed] [Google Scholar]
  • 59. Orfanidou C, Maliogka VI, Katis NI. First report of Cucurbit chlorotic yellows virus in cucumber, melon, and watermelon in Greece. Plant Dis2014;98:1446–6. [DOI] [PubMed] [Google Scholar]
  • 60. Lecoq H, Desbiez C. Viruses of cucurbit crops in the Mediterranean region: an ever-changing picture. Adv Virus Res. 2012;84:67–126. [DOI] [PubMed] [Google Scholar]
  • 61. Desbiez C, Wipf-Scheibel C, Millot Pet al. Distribution and evolution of the major viruses infecting cucurbitaceous and solanaceous crops in the French Mediterranean area. Virus Res. 2020;286:198042. [DOI] [PubMed] [Google Scholar]
  • 62. Desbiez C, Verdin E, Moury Bet al. Prevalence and molecular diversity of the main viruses infecting cucurbit and solanaceous crops in Azerbaijan. Eur J Plant Pathol2019;153:359–69. [Google Scholar]
  • 63. Mnari-Hattab M, Gauthier N, Zouba A. Biological and molecular characterization of the Cucurbit aphid-borne yellows virus affecting cucurbits in Tunisia. Plant Dis2009;93:1065–72. [DOI] [PubMed] [Google Scholar]
  • 64. Kassem MA, Sempere RN, Juárez Met al. Cucurbit aphid-borne yellows virus is prevalent in field-grown cucurbit crops of southeastern Spain. Plant Dis. 2007;91:232–8. [DOI] [PubMed] [Google Scholar]
  • 65. Gonsalves D, Tripathi S, Carr JBet al. Papaya ringspot virus. Plant Health Instr2010;10:1094. [Google Scholar]
  • 66. Yap YK, Duangjit J, Panyim S. N-terminal of Papaya ringspot virus type-W (PRSV-W) helper component proteinase (HC-pro) is essential for PRSV systemic infection in zucchini. Virus Genes. 2009;38:461–7. [DOI] [PubMed] [Google Scholar]
  • 67. Barbosa GDS, Lima JADA, Queiróz MADet al. Identification and effects of mixed infection of potyvirus isolates with Cucumber mosaic virus in cucurbits. Rev Caatinga. 2016;29:1028–35. [Google Scholar]
  • 68. Ling K-S, Levi A, Adkins Set al. Development and field evaluation of multiple virus-resistant bottle gourd (Lagenaria siceraria). Plant Dis. 2013;97:1057–62. [DOI] [PubMed] [Google Scholar]
  • 69. Brotman Y, Normantovich M, Goldenberg Zet al. Dual resistance of melon to Fusarium oxysporum races 0 and 2 and to Papaya ring-spot virus is controlled by a pair of head-to-head-oriented NB-LRR genes of unusual architecture. Mol Plant. 2013;6:235–8. [DOI] [PubMed] [Google Scholar]
  • 70. Phukan UJ, Jeena GS, Tripathi Vet al. Regulation of Apetala2/ethylene response factors in plants. Front Plant Sci. 2017;8:150. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 71. Su J, Jiang J, Zhang Fet al. Current achievements and future prospects in the genetic breeding of chrysanthemum: a review. Hortic Res2019;6:109. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 72. Tanurdzic M, Banks JA. Sex-determining mechanisms in land plants. Plant Cell2004;16:S61–71. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 73. Rudich J, Halevy AH, Kedar N. Ethylene evolution from cucumber plants as related to sex expression. Plant Physiol1972;49:998–9. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 74. Takahashi H, Jaffe MJ. Further studies of auxin and ACC induced feminization in the cucumber plant using ethylene inhibitors. Phyton (B Aires)1984;44:81–6. [PubMed] [Google Scholar]
  • 75. Boualem A, Fergany M, Fernandez Ret al. A conserved mutation in an ethylene biosynthesis enzyme leads to andromonoecy in melons. Science2008;321:836–8. [DOI] [PubMed] [Google Scholar]
  • 76. García A, Aguado E, Parra Get al. Phenomic and genomic characterization of a mutant platform in Cucurbita pepo. Front Plant Sci2018;9:1049. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 77. Boualem A, Troadec C, Camps Cet al. A cucurbit androecy gene reveals how unisexual flowers develop and dioecy emerges. Science2015;350:688–91. [DOI] [PubMed] [Google Scholar]
  • 78. Martínez C, Manzano S, Megías Zet al. Molecular and functional characterization of CpACS27A gene reveals its involvement in monoecy instability and other associated traits in squash (Cucurbita pepo L.). Planta. 2014;239:1201–15. [DOI] [PubMed] [Google Scholar]
  • 79. Chen H, Sun J, Li Set al. An ACC oxidase gene essential for cucumber carpel development. Mol Plant2016;9:1315–27. [DOI] [PubMed] [Google Scholar]
  • 80. García A, Aguado E, Martínez Cet al. The ethylene receptors CpETR1A and CpETR2B cooperate in the control of sex determination in Cucurbita pepo. J Exp Bot. 2020;71:154–67. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 81. Schilling S, McCabe PF, Melzer R. Love is in the air: ethylene and sex determination in Cucurbita pepo. J Exp Bot2020;71:4–6. [DOI] [PubMed] [Google Scholar]
  • 82. Martin A, Troadec C, Boualem Aet al. A transposon-induced epigenetic change leads to sex determination in melon. Nature. 2009;461:1135–8. [DOI] [PubMed] [Google Scholar]
  • 83. Malepszy S, Niemirowicz-Szczytt K. Sex determination in cucumber (Cucumis sativus) as a model system for molecular biology. Plant Sci1991;80:39–47. [Google Scholar]
  • 84. Trebitsh T, Staub JE, O’Neill SD. Identification of a 1-aminocyclopropane-1-carboxylic acid synthase gene linked to the female (F) locus that enhances female sex expression in cucumber. Plant Physiol1997;113:987–95. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 85. Mibus H, Tatlioglu ET. Molecular characterization and isolation of the F/f gene for femaleness in cucumber (Cucumis sativus L.). Theor Appl Genet2004;109:1669–76. [DOI] [PubMed] [Google Scholar]
  • 86. Li Z, Huang S, Liu Set al. Molecular isolation of the M gene suggests that a conserved-residue conversion induces the formation of bisexual flowers in cucumber plants. Genetics2009;182:1381–5. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 87. Li Z, Han Y, Niu Het al. Gynoecy instability in cucumber (Cucumis sativus L.) is due to unequal crossover at the copy number variation-dependent femaleness (F) locus. Hortic Res. 2020;7:32. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 88. Knopf RR, Trebitsh T. The female-specific CS-ACS1G gene of cucumber. A case of gene duplication and recombination between the non-sex-specific 1-aminocyclopropane-1-carboxylate synthase gene and a branched-chain amino acid transaminase gene. Plant Cell Physiol2006;47:1217–28. [DOI] [PubMed] [Google Scholar]
  • 89. Wu T, Qin Z, Feng Zet al. Functional analysis of the promoter of a female-specific cucumber CsACS1G gene. Plant Mol Biol Rep2012;30:235–41. [Google Scholar]
  • 90. Huiming C, Liang X, Xiangming L. Analysis on CsACS1G gene determining gynoecious and location in cucumber. Mol Plant Breed2005;3:520–4. [Google Scholar]
  • 91. Zhang H, Li S, Yang Let al. Gain-of-function of the 1-aminocyclopropane-1-carboxylate synthase gene ACS1G induces female flower development in cucumber gynoecy. Plant Cell2021;33:306–21. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 92. Xanthopoulou A, Montero-Pau J, Picó Bet al. A comprehensive RNA-Seq-based gene expression atlas of the summer squash (Cucurbita pepo) provides insights into fruit morphology and ripening mechanisms. BMC Genomics. 2021;22:341. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 93. Tan J, Tao Q, Niu Het al. A novel allele of monoecious (m) locus is responsible for elongated fruit shape and perfect flowers in cucumber (Cucumis sativus L.). Theor Appl Genet. 2015;128:2483–93. [DOI] [PubMed] [Google Scholar]
  • 94. Kamachi S, Mizusawa H, Matsuura Set al. Expression of two 1-aminocyclopropane-1-carboxylate synthase genes, CS-ACS1 and CS-ACS2, correlated with sex phenotypes in cucumber plants (Cucumis sativus L.). Plant Biotechnol. 2000;17:69–74. [Google Scholar]
  • 95. Li Z, Wang S, Tao Qet al. A putative positive feedback regulation mechanism in CsACS2 expression suggests a modified model for sex determination in cucumber (Cucumis sativus L.). J Exp Bot. 2012;63:4475–84. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 96. Eleblu JSY, Haraghi A, Mania Bet al. The gynoecious CmWIP1 transcription factor interacts with CmbZIP48 to inhibit carpel development. Sci Rep2019;9:15443. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 97. Becker C, Weigel D. Epigenetic variation: origin and transgenerational inheritance. Curr Opin Plant Biol2012;15:562–7. [DOI] [PubMed] [Google Scholar]
  • 98. Boualem A, Lemhemdi A, Sari M-Aet al. The andromonoecious sex determination gene predates the separation of Cucumis and Citrullus genera. PLoS One2016;11:e0155444. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 99. Roldan MVG, Izhaq F, Verdenaud Met al. Integrative genome-wide analysis reveals the role of WIP proteins in inhibition of growth and development. Commun Biol. 2020;3:1–12. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 100. Tadmor Y, Burger J, Yaakov Iet al. Genetics of flavonoid, carotenoid, and chlorophyll pigments in melon fruit rinds. J Agric Food Chem2010;58:10722–8. [DOI] [PubMed] [Google Scholar]
  • 101. Yuan H, Zhang J, Nageswaran Det al. Carotenoid metabolism and regulation in horticultural crops. Hortic Res. 2015;2:1–11. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 102. Burger Y, Paris HS, Cohen Ret al. Genetic diversity of Cucumis melo. Hortic Rev2010;36:165–98. [Google Scholar]
  • 103. Lv P, Li N, Liu Het al. Changes in carotenoid profiles and in the expression pattern of the genes in carotenoid metabolisms during fruit development and ripening in four watermelon cultivars. Food Chem. 2015;174:52–9. [DOI] [PubMed] [Google Scholar]
  • 104. Azevedo-Meleiro CH, Rodriguez-Amaya DB. Qualitative and quantitative differences in carotenoid composition among Cucurbita moschata, Cucurbita maxima, and Cucurbita pepo. J Agric Food Chem. 2007;55:4027–33. [DOI] [PubMed] [Google Scholar]
  • 105. Hermanns AS, Zhou X, Xu Qet al. Carotenoid pigment accumulation in horticultural plants. Hortic Plant J. 2020;6:343–60. [Google Scholar]
  • 106. Subburaj S, Tu L, Lee Ket al. A genome-wide analysis of the pentatricopeptide repeat (PPR) gene family and PPR-derived markers for flesh color in watermelon (Citrullus lanatus). Genes. 2020;11:1125. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 107. Wang C, Qiao A, Fang Xet al. Fine mapping of lycopene content and flesh color related gene and development of molecular marker-assisted selection for flesh color in watermelon (Citrullus lanatus). Front Plant Sci. 2019;10:1240. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 108. Galpaz N, Gonda I, Shem-Tov Det al. Deciphering genetic factors that determine melon fruit-quality traits using RNA-Seq-based high-resolution QTL and eQTL mapping. Plant J. 2018;94:169–91. [DOI] [PubMed] [Google Scholar]
  • 109. Tzuri G, Zhou X, Chayut Net al. A ‘golden’ SNP in CmOr governs fruit flesh color of melon (Cucumis melo). Plant J. 2015;82:267–79. [DOI] [PubMed] [Google Scholar]
  • 110. Feder A, Burger J, Gao Set al. A Kelch domain-containing F-box coding gene negatively regulates flavonoid accumulation in Cucumis melo L. Plant Physiol. 2015;169:1714–26. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 111. Ge Y, Li X, Yang X Xet al. Genetic analysis and gene mapping of rind and flesh color of Cucurbita maxima. Acta Bot Boreali-Occid Sin. 2015;35:1524–9. [Google Scholar]
  • 112. González-Verdejo CI, Obrero Á, Román Bet al. Expression profile of carotenoid cleavage dioxygenase genes in summer squash (Cucurbita pepo L.). Plant Foods Hum Nutr. 2015;70:200–6. [DOI] [PubMed] [Google Scholar]
  • 113. Nakkanong K, Yang JH, Zhang MF. Carotenoid accumulation and carotenogenic gene expression. J Agric Food Chem. 2012;60:5936–44. [DOI] [PubMed] [Google Scholar]
  • 114. Obrero Á, González-Verdejo CI, Die JVet al. Carotenogenic gene expression and carotenoid accumulation in three varieties of Cucurbita pepo during fruit development. J Agric Food Chem. 2013;61:6393–409. [DOI] [PubMed] [Google Scholar]
  • 115. Wyatt LE, Strickler SR, Mueller LAet al. Comparative analysis of Cucurbita pepo metabolism throughout fruit development in acorn squash and oilseed pumpkin. Hortic Res. 2016;3:16045. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 116. Tuan PA, Kim JK, Park NIet al. Carotenoid content and expression of phytoene synthase and phytoene desaturase genes in bitter melon (Momordica charantia). Food Chem. 2011;126:1686–92. [DOI] [PubMed] [Google Scholar]
  • 117. Cuong DM, Arasu MV, Jeon Jet al. Medically important carotenoids from Momordica charantia and their gene expressions in different organs. Saudi J Biol Sci. 2017;24:1913–9. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 118. Luo Z, Zhang J, Li Jet al. A STAY-GREEN protein SlSGR1 regulates lycopene and β-carotene accumulation by interacting directly with SlPSY1 during ripening processes in tomato. New Phytol. 2013;198:442–52. [DOI] [PubMed] [Google Scholar]
  • 119. Fray RG, Grierson D. Identification and genetic analysis of normal and mutant phytoene synthase genes of tomato by sequencing, complementation and co-suppression. Plant Mol Biol1993;22:589–602. [DOI] [PubMed] [Google Scholar]
  • 120. Zhang J, Guo S, Ren Yet al. High-level expression of a novel chromoplast phosphate transporter ClPHT4;2 is required for flesh color development in watermelon. New Phytol. 2017;213:1208–21. [DOI] [PubMed] [Google Scholar]
  • 121. Liu J, Van Eck J, Cong Bet al. A new class of regulatory genes underlying the cause of pear-shaped tomato fruit. Proc Natl Acad Sci USA. 2002;99:13302–6. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 122. Xiao H, Jiang N, Schaffner EKet al. Retrotransposon-mediated gene duplication underlies morphological variation of tomato fruit. Science2008;319:1527–30. [DOI] [PubMed] [Google Scholar]
  • 123. Knaap E, Chakrabarti M, Chu YHet al. What lies beyond the eye: the molecular mechanisms regulating tomato fruit weight and shape. Front Plant Sci2014;5:227. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 124. Zhao JY, Jiang L, Che Get al. A functional allele of CsFUL1 regulates fruit length through inhibiting CsSUP and auxin transport in cucumber. Plant Cell. 2019;31:1289–307. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 125. Wang L, Ruan YL. Regulation of cell division and expansion by sugar and auxin signaling. Front Plant Sci2013;4:163. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 126. Dou JL, Zhao S, Lu Xet al. Genetic mapping reveals a candidate gene (ClFS1) for fruit shape in watermelon (Citrullus lanatus L.). Theor Appl Genet2018;131:947–58. [DOI] [PubMed] [Google Scholar]
  • 127. Karlova R, Rosin FM, Busscher-Lange Jet al. Transcriptome and metabolite profiling show that APETALA2a is a major regulator of tomato fruit ripening. Plant Cell. 2011;23:923–41. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 128. Wu S, Zhang B, Keyhaninejad Net al. A common genetic mechanism underlies morphological diversity in fruits and other plant organs. Nat Commun2018;9:1–12. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 129. Cao K, Zhou Z, Wang Qet al. Genome-wide association study of 12 agronomic traits in peach. Nat Commun2016;7:13246. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 130. Pan YP, Liang X, Gao Met al. Round fruit shape in WI7239 cucumber is controlled by two interacting quantitative trait loci with one putatively encoding a tomato SUN homolog. Theor Appl Genet. 2017;130:573–86. [DOI] [PubMed] [Google Scholar]
  • 131. Cheng, Z., Liu Z, Xu Yet al. Fine mapping and identification of the candidate gene BFS for fruit shape in wax gourd (Benincasa hispida). Theor Appl Genet. 2021;134:3983–95. [DOI] [PubMed] [Google Scholar]
  • 132. Paris HS, Brown RN. The genes of pumpkin and squash. HortScience. 2005;40:1620–30. [Google Scholar]
  • 133. Pan YP, Wang Y, McGregor Cet al. Genetic architecture of fruit size and shape variation in cucurbits: a comparative perspective. Theor Appl Genet. 2020;133:1–21. [DOI] [PubMed] [Google Scholar]
  • 134. Loy JB. Fruit size in melon in monoecious and andromonoecious isolines. Cucurbit Genetics Cooperative Report. 2006;28–29:12–3. [Google Scholar]
  • 135. Kumar R, Khurana A, Sharma A. Role of plant hormones and their interplay in development and ripening of fleshy fruits. J Exp Bot2013;65:4561–75. [DOI] [PubMed] [Google Scholar]
  • 136. Li S, Chen K, Grierson D. Molecular and hormonal mechanisms regulating fleshy fruit ripening. Cell. 2021;10:1136. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 137. Tucker G, Yin X, Zhang Aet al. Ethylene and fruit softening. Food Qual Saf. 2017;1:253–67. [Google Scholar]
  • 138. Sundberg E, Østergaard L. Distinct and dynamic auxin activities during reproductive development. Cold Spring Harb Perspect Biol2009;1:a001628. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 139. Yuan Y, Xu X, Gong Zet al. Auxin response factor 6A regulates photosynthesis, sugar accumulation, and fruit development in tomato. Hortic Res. 2019;6:85–116. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 140. Pattison R, Csukasi F, Catalá C. Mechanisms regulating auxin action during fruit development. Physiol Plant. 2014;151:62–72. [DOI] [PubMed] [Google Scholar]
  • 141. Zhang Y, Zeng Z, Chen Cet al. Genome-wide characterization of the auxin response factor (ARF) gene family of litchi (Litchi chinensis Sonn.): phylogenetic analysis, miRNA regulation and expression changes during fruit abscission. PeerJ. 2019;7:e6677. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 142. Shukla A, Singh VK, Bharadwaj DRet al. De novo assembly of bitter gourd transcriptomes: gene expression and sequence variations in gynoecious and monoecious lines. PLoS One. 2015;10:e0128331. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 143. Ren H, Gray WM. SAUR proteins as effectors of hormonal and environmental signals in plant growth. Mol Plant. 2015;8:1153–64. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 144. Stortenbeker N, Bemer M. The SAUR gene family: the plant’s toolbox for adaptation of growth and development. J Exp Bot. 2019;70:17–27. [DOI] [PubMed] [Google Scholar]
  • 145. Zhang H, Yu Z, Yao Xet al. Genome-wide identification and characterization of small auxin-up RNA (SAUR) gene family in plants: evolution and expression profiles during normal growth and stress response. BMC Plant Biol. 2021;21:1–14. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 146. Xu C, Liberatore KL, CA MAet al. A cascade of arabinosyltransferases controls shoot meristem size in tomato. Nat Genet. 2015;47:784–95. [DOI] [PubMed] [Google Scholar]
  • 147. Jha P, Ochatt SJ, Kumar V. WUSCHEL: a master regulator in plant growth signaling. Plant Cell Rep. 2020;39:431–44. [DOI] [PubMed] [Google Scholar]
  • 148. Lian G, Ding Z, Wang Qet al. Origins and evolution of WUSCHEL-related homeobox protein family in plant kingdom. Sci World J. 2014;2014 10.1155/2014/534140. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 149. Bashline L, Lei L, Li Set al. Cell wall, cytoskeleton, and cell expansion in higher plants. Mol Plant. 2014;7:586–600. [DOI] [PubMed] [Google Scholar]
  • 150. Forlani S, Masiero S, Mizzotti C. Fruit ripening: the role of hormones, cell wall modifications, and their relationship with pathogens. J Exp Bot2019;70:2993–3006. [DOI] [PubMed] [Google Scholar]
  • 151. Wang D, Yeats TH, Uluisik Set al. Fruit softening: revisiting the role of pectin. Trends Plant Sci2018;23:302–10. [DOI] [PubMed] [Google Scholar]
  • 152. Ando K, Carr KM, Grumet R. Transcriptome analyses of early cucumber fruit growth identifies distinct gene modules associated with phases of development. BMC Genomics. 2012;13:518. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 153. Harmer S, Orford SJ, Timmis JNet al. Characterisation of six α-expansin genes in Gossypium hirsutum (upland cotton). Mol Gen Genomics2002;268:1–9. [DOI] [PubMed] [Google Scholar]
  • 154. Marowa P, Ding A, Kong Y. Expansins: roles in plant growth and potential applications in crop improvement. Plant Cell Rep. 2016;35:949–65. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 155. Anwar R, Fatima S, Mattoo AKet al. Fruit architecture in polyamine-rich tomato germplasm is determined via a medley of cell cycle, cell expansion, and fruit shape genes. Plants (Basel). 2019;8:387. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 156. Zhang H, Tan J, Zhang Met al. Comparative transcriptomic analysis of two bottle gourd accessions differing in fruit size. Genes2020;11:359. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 157. Chatzopoulou F, Sanmartin M, Mellidou Iet al. Silencing of ascorbate oxidase results in reduced growth, altered ascorbic acid levels and ripening pattern in melon fruit. Plant Physiol Biochem. 2020;156:291–303. [DOI] [PubMed] [Google Scholar]
  • 158. Xanthopoulou A, Ganopoulos I, Psomopoulos Fet al. De novo comparative transcriptome analysis of genes involved in fruit morphology of pumpkin cultivars with extreme size difference and development of EST-SSR markers. Gene. 2017;622:50–66. [DOI] [PubMed] [Google Scholar]
  • 159. Cohen S, Itkin M, Yeselson Yet al. The PH gene determines fruit acidity and contributes to the evolution of sweet melons. Nat Commun2014;5:4026. [DOI] [PubMed] [Google Scholar]
  • 160. Zhang H, Wang H, Yi Het al. Transcriptome profiling of Cucumis melo fruit development and ripening. Hortic Res. 2016;3:16014. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 161. Zhou Y, Ma Y, Zeng Jet al. Convergence and divergence of bitterness biosynthesis and regulation in Cucurbitaceae. Nat Plants2016;2:16183. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 162. Buescher RH, Buescher RW. Production and stability of (E,Z)-2,6-nonadienal, the major flavor volatile of cucumbers. J Food Sci. 2001;66:357–61. [Google Scholar]
  • 163. Gong C, Zhu H, Lu Xet al. An integrated transcriptome and metabolome approach reveals the accumulation of taste-related metabolites and gene regulatory networks during watermelon fruit development. Planta2021;254:1–12. [DOI] [PubMed] [Google Scholar]
  • 164. Jayakodi M, Schreiber M, Mascher M. Sweet genes in melon and watermelon. Nat Genet2019;51:1572–3. [DOI] [PubMed] [Google Scholar]
  • 165. Ren Y, Li M, Guo Set al. Evolutionary gain of oligosaccharide hydrolysis and sugar transport enhanced carbohydrate partitioning in sweet watermelon fruits. Plant Cell. 2021;33:1554–73. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 166. Ren Y, Sun H, Zong Met al. Localization shift of a sugar transporter contributes to phloem unloading in sweet watermelons. New Phytol2020;227:1858–71. [DOI] [PubMed] [Google Scholar]
  • 167. Ren Y, Guo S, Zhang Jet al. A tonoplast sugar transporter underlies a sugar accumulation QTL in watermelon. Plant Physiol2018;176:836–50. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 168. Cheng J, Wen S, Xiao Set al. Overexpression of the tonoplast sugar transporter CmTST2 in melon fruit increases sugar accumulation. J Exp Bot2018;69:511–23. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 169. Yang Y, Li J, Li Het al. The bZIP gene family in watermelon: genome-wide identification and expression analysis under cold stress and root-knot nematode infection. PeerJ. 2019;7:e7878. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 170. Chen G, Hackett R, Walker Det al. Identification of a specific isoform of tomato lipoxygenase (TomLoxC) involved in the generation of fatty acid-derived flavour compounds. Plant Physiol. 2004;136:2641–51. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 171. Wyllie SG, Fellman JK. Formation of volatile branched chain esters in bananas (Musa sapientum L.). J Agric Food Chem2000;48:3493–6. [DOI] [PubMed] [Google Scholar]
  • 172. Gonda I, Bar E, Portnoy Vet al. Branched-chain and aromatic amino acid catabolism into aroma volatiles in Cucumis melo L. fruit. J Exp Bot2010;61:1111–23. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 173. Gonda I, Lev S, Bar Eet al. Catabolism of L-methionine in the formation of sulfur and other volatiles in melon (Cucumis melo L.) fruit. Plant J. 2013;74:458–72. [DOI] [PubMed] [Google Scholar]
  • 174. Wyllie SG, Leach DN. Sulfur-containing compounds in the aroma volatiles of melons (Cucumis melo). J Agric Food Chem. 1992;40:253–6. [Google Scholar]
  • 175. Jordan MJ, Shaw PE, Goodner KL. Volatile components in aqueous essence and fresh fruit of Cucumis melo cv. Athena (muskmelon) by GC-MS and GC-O. J Agric Food Chem. 2001;49:5929–33. [DOI] [PubMed] [Google Scholar]
  • 176. Kaushik U, Aeri V, Mir SR. Cucurbitacins – an insight into medicinal leads from nature. Pharmacogn Rev2015;9:12–8. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 177. Chen JC, Chiu MH, Nie RLet al. Cucurbitacins and cucurbitane glycosides: structures and biological activities. Nat Prod Rep. 2005;22:386–99. [DOI] [PubMed] [Google Scholar]
  • 178. Tannin-Spitz T, Grossman S, Dovrat Set al. Growth inhibitory activity of cucurbitacin glucosides isolated from Citrullus colocynthis on human breast cancer cells. Biochem Pharmacol. 2007;73:56–67. [DOI] [PubMed] [Google Scholar]
  • 179. Thoennissen NH, Iwanski GB, Doan NBet al. Cucurbitacin B induces apoptosis by inhibition of the JAK/STAT pathway and potentiates antiproliferative effects of gemcitabine on pancreatic cancer cells. Cancer Res2009;69:5876–84. [DOI] [PubMed] [Google Scholar]
  • 180. Chen X, Bao J, Guo Jet al. Biological activities and potential molecular targets of cucurbitacins: a focus on cancer. Anti-Cancer Drugs. 2012;23:777–87. [DOI] [PubMed] [Google Scholar]
  • 181. Knecht DA, LaFleur RA, Kahsai AWet al. Cucurbitacin I inhibits cell motility by indirectly interfering with actin dynamics. PLoS One. 2010;5:e14039. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 182. Seo CR, Yang DK, Song N-Jet al. Cucurbitacin B and cucurbitacin I suppress adipocyte differentiation through inhibition of STAT3 signaling. Food Chem Toxicol. 2014;64:217–24. [DOI] [PubMed] [Google Scholar]
  • 183. Rios JL. Effects of triterpenes on the immune system. J Ethnopharmacol2010;128:1–14. [DOI] [PubMed] [Google Scholar]
  • 184. Phillips DR, Rasbery JM, Bartel Bet al. Biosynthetic diversity in plant triterpene cyclization. Curr Opin Plant Biol2006;9:305–14. [DOI] [PubMed] [Google Scholar]
  • 185. Shibuya M, Adachi S, Ebizuka Y. Cucurbitadienol synthase, the first committed enzyme for cucurbitacin biosynthesis, is a distinct enzyme from cycloartenol synthase for phytosterol biosynthesis. Tetrahedron. 2004;60:6995–7003. [Google Scholar]
  • 186. Lester G. Melon (Cucumis melo L.) fruit nutritional quality and health functionality. HortTechnology1997;7:222–7. [Google Scholar]
  • 187. Matsuo K, DeMilo AB, Schroder RFWet al. Rapid high-performance liquid chromatography method to quantitate elaterinide in juice and reconstituted residues from a bitter mutant of Hawkesbury watermelon. J Agric Food Chem1999;47:2755–9. [DOI] [PubMed] [Google Scholar]
  • 188. Karlova R, Chapman N, David Ket al. Transcriptional control of fleshy fruit development and ripening. J Exp Bot2014;65:4527–41. [DOI] [PubMed] [Google Scholar]
  • 189. Zhang Z, Li X. Genome-wide identification of AP2/ERF superfamily genes and their expression during fruit ripening of Chinese jujube. Sci Rep2018;8:15612. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 190. Wang X, Zeng W, Ding Yet al. Peach ethylene response factor PpeERF2 represses the expression of ABA biosynthesis and cell wall degradation genes during fruit ripening. Plant Sci. 2019;283:116–26. [DOI] [PubMed] [Google Scholar]
  • 191. Liu C, Long J, Zhu Ket al. Characterization of a citrus R2R3-MYB transcription factor that regulates the flavonol and hydroxycinnamic acid biosynthesis. Sci Rep. 2016;6:25352. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 192. Phan HA, Li SF, Parish RW. MYB80, a regulator of tapetal and pollen development, is functionally conserved in crops. Plant Mol Biol2012;78:171–83. [DOI] [PubMed] [Google Scholar]
  • 193. Cheng H, Song S, Xiao Let al. Gibberellin acts through jasmonate to control the expression of MYB21, MYB24, and MYB57 to promote stamen filament growth in Arabidopsis. PLoS Genet. 2009;5:e1000440. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 194. Seo PJ, Park CM. MYB96-mediated abscisic acid signals induce pathogen resistance response by promoting salicylic acid biosynthesis in Arabidopsis. New Phytol2010;186:471–83. [DOI] [PubMed] [Google Scholar]
  • 195. Yang S, Cai Y, Liu Xet al. A CsMYB6-CsTRY module regulates fruit trichome initiation in cucumber. J Exp Bot2018;69:1887–902. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 196. Zhang L, Pan J, Wang Get al. Cucumber CsTRY negatively regulates anthocyanin biosynthesis and trichome formation when expressed in tobacco. Front Plant Sci2019;10:1232. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 197. Liu M, Zhang C, Duan Let al. CsMYB60 is a key regulator of flavonols and proanthocyanidans that determine the colour of fruit spines in cucumber. J Exp Bot2019;70:69–84. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 198. Walters SA, Shetty NV, Wehner TC. Segregation and linkage of several genes in cucumber. J Am Soc Hortic Sci2001;126:442–50. [Google Scholar]
  • 199. Xu Q, He J, Dong JHet al. Genomic survey and expression profiling of the MYB gene family in watermelon. Hortic Plant J. 2018;4:1–15. [Google Scholar]
  • 200. Tian F, Yang D-C, Meng Y-Qet al. PlantRegMap: charting functional regulatory maps in plants. Nucleic Acids Res. 2020;48:D1104–13. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 201. Sidhu NS, Pruthi G, Singh Set al. Genome-wide identification and analysis of GRAS transcription factors in the bottle gourd genome. Sci Rep. 2020;10:14338. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 202. Zhang, Q., He J, Xu Yet al. Comparative analysis of GRAS genes in six Cucurbitaceae species provides novel insights into their evolution and function. 2021; 10.21203/rs.3.rs-258961/v1. [DOI]
  • 203. Li M, Wang X, Li Cet al. Silencing GRAS2 reduces fruit weight in tomato. J Integr Plant Biol. 2018;60:498–513. [DOI] [PubMed] [Google Scholar]
  • 204. Puranik S, Sahu PP, Srivastava PSet al. NAC proteins: regulation and role in stress tolerance. Trends Plant Sci. 2012;17:369–81. [DOI] [PubMed] [Google Scholar]
  • 205. Ríos P, Argyris J, Vegas Jet al. ETHQV6.3 is involved in melon climacteric fruit ripening and is encoded by a NAC domain transcription factor. Plant J. 2017;91:671–83. [DOI] [PubMed] [Google Scholar]
  • 206. Zhu M, Chen G, Zhou Set al. A new tomato NAC (NAM/ATAF1/2/CUC2) transcription factor, SlNAC4, functions as a positive regulator of fruit ripening and carotenoid accumulation. Plant Cell Physiol2014;55:119–35. [DOI] [PubMed] [Google Scholar]
  • 207. Li S, Zhu B, Pirrello Jet al. Roles of RIN and ethylene in tomato fruit ripening and ripening-associated traits. New Phytol. 2020;226:460–75. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 208. Ezura H, Owino WO. Melon, an alternative model plant for elucidating fruit ripening. Plant Sci2008;175:121–9. [Google Scholar]
  • 209. Itkin M, Seybold H, Breitel Det al. TOMATO AGAMOUS-LIKE 1 is a component of the fruit ripening regulatory network. Plant J2009;60:1081–95. [DOI] [PubMed] [Google Scholar]
  • 210. Lin Z, Zhong S, Grierson D. Recent advances in ethylene research. J Exp Bot2009;60:3311–36. [DOI] [PubMed] [Google Scholar]
  • 211. Lee JM, Joung J-G, McQuinn Ret al. Combined transcriptome, genetic diversity and metabolite profiling in tomato fruit reveals that the ethylene response factor SlERF6 plays an important role in ripening and carotenoid accumulation. Plant J2012;70:191–204. [DOI] [PubMed] [Google Scholar]
  • 212. Chung MY, Vrebalov J, Alba Ret al. A tomato (Solanum lycopersicum) APETALA2/ERF gene, SlAP2a, is a negative regulator of fruit ripening. Plant J. 2010;64:936–47. [DOI] [PubMed] [Google Scholar]
  • 213. Jiang L, Yan S, Yang Wet al. Transcriptomic analysis reveals the roles of microtubule-related genes and transcription factors in fruit length regulation in cucumber (Cucumis sativus L.). Sci Rep2015;5:8031. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 214. Xu X, Pan J, He Met al. Transcriptome profiling reveals key genes related to astringency during cucumber fruit development. 3 Biotech. 2019;9:390. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 215. Wang L, Cao C, Zheng Set al. Transcriptomic analysis of short-fruit 1 (sf1) reveals new insights into the variation of fruit-related traits in Cucumis sativus. Sci Rep. 2017;7:2950. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 216. Blanca J, Esteras C, Ziarsolo Pet al. Transcriptome sequencing for SNP discovery across Cucumis melo. BMC Genomics2012;13:280. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 217. Shin AY, Kim Y-M, Koo Net al. Transcriptome analysis of the oriental melon (Cucumis melo L. var. makuwa) during fruit development. PeerJ. 2017;5:e2834. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 218. Chayut N, Yuan H, Saar Yet al. Comparative transcriptome analyses shed light on carotenoid production and plastid development in melon fruit. Hortic Res. 2021;8:112. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 219. Guo S, Liu J, Zheng Yet al. Characterization of transcriptome dynamics during watermelon fruit development: sequencing, assembly, annotation and gene expression profiles. BMC Genomics2011;12:454. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 220. Hyun TK, Rim Y, Jang H-Jet al. De novo transcriptome sequencing of Momordica cochinchinensis to identify genes involved in the carotenoid biosynthesis. Plant Mol Biol. 2012;79:413–27. [DOI] [PubMed] [Google Scholar]
  • 221. Blanca J, Cañizares J, Roig Cet al. Transcriptome characterization and high throughput SSRs and SNPs discovery in Cucurbita pepo (Cucurbitaceae). BMC Genomics. 2011;12:104. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 222. Wu TQ, Luo S, Wang Ret al. The first Illumina-based de novo transcriptome sequencing and analysis of pumpkin (Cucurbita moschata Duch.) and SSR marker development. Mol Breeding2014;34:1437–47. [Google Scholar]
  • 223. Huang HX, Yu T, Li J-Xet al. Characterization of Cucurbita maxima fruit metabolomic profiling and transcriptome to reveal fruit quality and ripening gene expression patterns. J Plant Biol. 2019;62:203–16. [Google Scholar]
  • 224. Jiang B, Xie D, Liu Wet al. De novo assembly and characterization of the transcriptome, and development of SSR markers in wax gourd (Benicasa hispida). PLoS One. 2013;8:e71054. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 225. Liu S, Gao P, Zhu Qet al. Resequencing of 297 melon accessions reveals the genomic history of improvement and loci related to fruit traits in melon. Plant Biotechnol J. 2020;18:2545–58. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 226. Baloglu MC. Genomics of cucurbits. In: Rout GR, Peter KV, eds. Genetic Engineering of Horticultural Crops. Academic Press, 2018,413–32. [Google Scholar]
  • 227. González M, Xu M, Esteras Cet al. Towards a TILLING platform for functional genomics in Piel de Sapo melons. BMC Res Notes. 2011;4:289. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 228. Vicente-Dólera N, Troadec C, Moya Met al. First TILLING platform in Cucurbita pepo: a new mutant resource for gene function and crop improvement. PLoS One. 2014;9:e112743. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 229. Boualem A, Fleurier S, Troadec Cet al. Development of a Cucumis sativus TILLinG platform for forward and reverse genetics. PLoS One. 2014;9:e97963. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 230. Dahmani-Mardas F, Troadec C, Boualem Aet al. Engineering melon plants with improved fruit shelf life using the TILLING approach. PLoS One2010;5:e15776. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 231. Fraenkel R, Kovalski I, Troadec Cet al. Development and evaluation of a cucumber TILLING population. BMC Res Notes2014;7:846. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 232. Hao N, Han D, Huang Ket al. Genome-based breeding approaches in major vegetable crops. Theor Appl Genet. 2020;133:1739–52. [DOI] [PubMed] [Google Scholar]
  • 233. Khan MZ, Zaidi SS, Amin Iet al. A CRISPR way for fast-forward crop domestication. Trends Plant Sci. 2019;24:293–6. [DOI] [PubMed] [Google Scholar]
  • 234. Chen K, Wang Y, Zhang Ret al. CRISPR/Cas genome editing and precision plant breeding in agriculture. Annu Rev Plant Biol2019;70:667–97. [DOI] [PubMed] [Google Scholar]
  • 235. Hu B, Li D, Liu Xet al. Engineering non-transgenic gynoecious cucumber using an improved transformation protocol and optimized CRISPR/Cas9 system. Mol Plant. 2017;10:1575–8. [DOI] [PubMed] [Google Scholar]
  • 236. Xin T, Zhang Z, Li Set al. Genetic regulation of ethylene dosage for cucumber fruit elongation. Plant Cell2019;31:1063–76. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 237. Tian S, Jiang L, Gao Qet al. Efficient CRISPR/Cas9-based gene knockout in watermelon. Plant Cell Rep2017;36:399–406. [DOI] [PubMed] [Google Scholar]
  • 238. Tian S, Jiang L, Cui Xet al. Engineering herbicide-resistant watermelon variety through CRISPR/Cas9-mediated base-editing. Plant Cell Rep2018;37:1353–6. [DOI] [PubMed] [Google Scholar]
  • 239. Huang Y, Cao H, Yang Let al. Tissue-specific respiratory burst oxidase homologue-dependent H2O2 signaling to the plasma membrane H+-ATPase confers potassium uptake and salinity tolerance in Cucurbitaceae. J Exp Bot. 2019;70:5879–93. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 240. Ma X, Zhu Q, Chen Yet al. CRISPR/Cas9 platforms for genome editing in plants: developments and applications. Mol Plant. 2016;9:961–74. [DOI] [PubMed] [Google Scholar]

Articles from Horticulture Research are provided here courtesy of Oxford University Press

RESOURCES