Home Life Sciences Selection and functional identification of Dof genes expressed in response to nitrogen in Populus simonii × Populus nigra
Article Open Access

Selection and functional identification of Dof genes expressed in response to nitrogen in Populus simonii × Populus nigra

  • Shenmeng Wang , Ruoning Wang and Chengjun Yang EMAIL logo
Published/Copyright: July 13, 2022

Abstract

In plants, Dof transcription factors are involved in regulating the expression of a series of genes related to N uptake and utilization. Therefore, the present study investigated how DNA-binding with one finger (Dof) genes are expressed in response to nitrogen (N) form and concentration to clarify the role of Dof genes and their functions in promoting N assimilation and utilization in poplar. The basic characteristics and expression patterns of Dof genes in poplar were analyzed by the use of bioinformatics methods. Dof genes expressed in response to N were screened, after which the related genes were cloned and transformed into Arabidopsis thaliana; the physiological indexes and the expression of related genes were subsequently determined. The function of Dof genes was then verified in Arabidopsis thaliana plants grown in the presence of different N forms and concentrations. Forty-four Dof genes were identified, most of which were expressed in the roots and young leaves, and some of the Dof genes were expressed under ammonia- and nitrate-N treatments. Three genes related to N induction were cloned, their proteins were found to localize in the nucleus, and PnDof30 was successfully transformed into Arabidopsis thaliana for functional verification. On comparing Arabidopsis thaliana with WT Arabidopsis thaliana plants, Arabidopsis thaliana plants overexpressing the Dof gene grew better under low N levels; the contents of soluble proteins and chlorophyll significantly increased, while the soluble sugar content significantly decreased. The expressions of several AMT, NRT, and GS genes were upregulated, while the expressions of several others were downregulated, and the expression of PEPC and PK genes significantly increased. In addition, the activity of PEPC, PK, GS, and NR enzymes significantly increased. The results showed that overexpression of PnDof30 significantly increased the level of carbon and N metabolism and improved the growth of transgenic Arabidopsis thaliana plants under low-N conditions. The study revealed the biological significance of poplar Dof transcription factors in N response and regulation of related downstream gene expression and provided some meaningful clues to explain the huge difference between poplar and Arabidopsis thaliana transformed by exogenous Dof gene, which could promote the comprehensive understanding of the molecular mechanism of efficient N uptake and utilization in trees.

1 Introduction

Nitrogen (N) is one of the several nutrient elements required for the process of plant growth and development [1] and is also one of the most abundant elements in plants. N is present within approximately 70% of the nutrients plants obtain from the environment [2]. N is assimilated into substances that can be used directly or in enzymatic reactions. N is used in various physiological and metabolic reactions, including synthesizing nucleic acids, proteins, coenzyme factors, and molecules involved in signal transmission and storage. However, in nature, the N concentration in the soil is generally low, and N nutrition is often the main factor limiting the growth of plants, including trees [3,4,5]. Fertilization can effectively alleviate N deficiency in an environment in a short period, but it can also reduce the nutrient absorption function of roots [6]; this reduction is not conducive to late plant growth and excessive reliance on chemical fertilizers brings about great pressure to the environment.

DNA-binding with one finger (Dof) transcription factors are plant-specific transcription factors. The Dof family has many members and is part of the zinc-finger protein superfamily. Dof proteins generally range from 200 to 400 amino acids (AAs) in length and contain both a highly conserved N-terminus and a less conserved C-terminus. The N-terminus contains a highly conserved Dof domain of 52 AAs in which the CX2CX21CX2C motif forms a single zinc-finger structure [7]. To date, except for the pumpkin Dof protein AOBP, which recognizes AGTA sequences [8], other Dof proteins recognize AAAG sequences or their reverse complementary sequence CTTT [9,10,11]. The diversity of the C-terminal region of Dofs may be related to the role of different regulatory signals. Combining with different regulatory proteins or other signaling factors to regulate the transcription of target genes is the basis of the functional diversity of Dof transcription factors.

Dof transcription factor-encoding genes generally are members of larger gene families. To date, 37 Dof family members have been identified in Arabidopsis thaliana [10], 46 have been identified in maize [12], 30 in rice [13], 26 in barley [11], 28 in soybean [14], 46 in carrot [15], 38 in pea [16], 42 in Tribulus alfalfa [17], and 29 in eggplant [18]. Dof transcription factors are largely involved in the regulation of photosynthesis, the synthesis of seed storage proteins, seed development and germination, dormancy, flowering time, cell wall synthesis, the development of vascular bundles, fruit maturity, the accumulation of starch, and other plant-specific biological processes [10,19,20,21,22,23]. In addition, Dof transcription factors also participate in the expression of genes related to the regulation of carbon (C) metabolism [24], the cell cycle [25], abiotic stress tolerance [26,27], and N absorption and utilization [28,29].

In 2000, Yanagisawa performed instantaneous expression in maize protoplasts and used electrophoretic mobility shift assays to identify the target genes regulated by maize Dof1 and found that ZmDof1 could bind the promoter of the phosphoenolpyruvate carboxylase (PEPC) (C4-type PEPC) gene [30]. In 2004, Yanagisawa et al. transformed ZmDof1 into the C3 plant species Arabidopsis thaliana, which improved the N uptake and assimilation efficiency of the transgenic Arabidopsis thaliana plants. Under low-N conditions, the transgenic Arabidopsis thaliana plants grew better than the wild-type (WT) Arabidopsis thaliana plants, and the free AA content significantly increased in the former [31]. Similar results were obtained in 2011 when Kurai transformed ZmDof1 genes into rice: the N-use efficiency and growth index of the transgenic rice plants significantly improved [32]. In 2008, when Rueda used pine protoplasts to study downstream genes whose expression was regulated by PpDof5 in Pinus pinaster, it was found that the transcription factor could activate the expression of the glutamine synthase gene GS1b and inhibit the expression of GS1a [33]. In 2015, Rueda-López et al. transformed PpDof5 into Arabidopsis thaliana and found that compared with those of the WT plants, the lignin content and C and N metabolism of the transgenic plants significantly increased [34]. OsDof25, a functional homolog of ZmDof1, was isolated from rice by Santos in 2012, and its expression was determined to be regulated by N. After OsDof25 was transformed into Arabidopsis thaliana, it was found that the C and N metabolism levels improved, and compared with that of the WT plants, the AA content of the transgenic plants significantly increased [35]. Similarly, Wang transformed the AtDof1 gene into tobacco in 2013 and found that the activities of the PEPC, pyruvate kinase (PK), glutamine synthetase (GS), and nitrate reductase (NR) enzymes increased significantly in the transgenic tobacco compared with the WT [36].

In 2013, however, Lin et al. transformed ZmDof1 into poplar, but there was no significant change in C and N metabolism or growth between the transgenic poplar and WT poplar [37]. This experiment showed that the N use efficiency of transgenic poplar did not improve under low-N conditions, whether in the culture flask or in the greenhouse, and genes involved in N metabolism and N absorption and utilization, such as PEPC, PK, Asparagine synthetase (AS), GS, NADP-malate dehydrogenase, isocitrate dehydrogenase, and other expression levels did not increase. The promoter regions of the above C/N metabolism-related genes all have Dof binding domain sequence AAAG, indicating that Dof transcription factors recognize these genes. It is possible that the Dof transcription factor in poplar may be involved in regulating the C/N balance pathway, and this regulatory mechanism may be different from that of maize and Arabidopsis thaliana function, and screened out the Dof transcription factor that regulates C/N balance in poplar. Therefore, in this study, to identify Dof members in poplar that can improve C and N metabolism and plant growth at low-N levels, we identified the members of the Dof transcription factor family in poplar, identified the genes expressed in response to N through various N treatments, and screened the candidate genes. Sequence analysis was performed, and the subcellular localization of the gene products was subsequently determined. Afterward, the candidate genes were transformed into Arabidopsis thaliana growing under low-N levels for functional identification.

In the recent 5 years, various studies on the Dof gene are still emerging one after another. In 2018, Wang screened and identified 24 Dof genes in the Dof genomes of physic nut, and divided them into three categories based on phylogenetic inference. The genome comparison discovered that the expansion of the Dof gene family in physic nut mainly resulted from segmental duplication, and this expansion was mainly subjected to positive selection. Furthermore, many JcDof genes were significantly responsive to the salt and drought treatments [38]. Syed identified Dof transcription factors in pineapple and characterized their expression profiles. Expression analysis using real-time quantitative PCR (qRT-PCR) of pineapple Dof genes family under different abiotic stress (cold, heat, salt, and drought) showed a dynamic response of Dof genes. Thus, we can see that Dof genes expression during abiotic stress reveals their vital role in pineapple growth and development, which could be utilized agronomically [39].

In 2019, Liu et al. investigated the role of PbDof9.2 in flowering regulation in Pyrus bretschneideri. It is concluded that the PbDof9.2 suppressed the flowering time regulator FT and could repress flowering time by promoting the activity of PbTFL1a and PbTFL1b promoters. These results suggest that Dof transcription factors have conserved functions in plant development [40]. On the contrary, Tokunaga et al. found that the overexpression of DOF-type transcription factors can enhance lipid synthesis in Chlorella vulgaris. Under N-deficient conditions, the transformant CvDOF#3 showed approximately 1.5-fold higher neutral lipid content per cell compared to the original strain and also showed a His-tagged DOF candidate protein expression of 0.6%. Microscopic observations revealed that CvDOF#3 cells were larger. The findings suggested that the overexpression of the endogenous DOF-type transcription factor can be used for improving the lipid content in Chlorella vulgaris [41].

In 2020, Waqas conducted a systematic genome-wide analysis of Dof family members in selected cotton species and identified 58, 55, 89, and 110 Dof genes in G. arboreum, G. raimondii, G. hirsutum, and G. barbadense, respectively. The combined phylogeny analysis among the GaDof, GrDof, GhDof, GbDof, and AtDof proteins showed orthologous genes among cotton Dofs. This proved the evolution of polyploid cotton from diploid cotton species [42]. In 2021, Neeta analyzed the Dof gene in Brassica napus and concluded that based on the orthology, synteny, and evolutionary analysis, the calculated divergence times indicated that the divergence of the Brassica and Arabidopsis genus (∼17 Mya), the whole-genome triplication event (9–15 Mya), and the formation of Brassica napus (7,500 years ago) drove the expansion of the BnaDof gene family. Synteny analysis also highlighted that most of the Dof genes in Brassica napus with known chromosomal locations were not translocated. Tissue-specific expression highlighted the role of BnaDofs in organ development and other developmental processes. Most of the BnaDofs were responsive to temperature fluctuations and were differentially regulated, particularly by cold stress [43].

2 Materials and methods

2.1 Plant material

Tissue culture-generated seedlings of Populus simonii × Populus nigra were grown in a growth chamber at 23°C, under 16 h of light and 8 h of darkness, and under a light intensity of 100 μmol m−2 s−1. Hydroponic cultivation was performed at 25°C, under 16 h of light and 8 h of darkness, and under a light intensity of 120 μmol m−2 s−1. Populus simonii × Populus nigra seedlings were cultured in lactate aqueous solution supplemented with 1 mM ammonium-nitrate for 1 week. After 3 days of N being withheld, the N supply was restored for 2 or 48 h, and then the N was withheld again for another 2 or 48 h. The solutions were replaced every 3 days. Samples were collected at each time point and frozen in liquid N for further study.

Seeds of Arabidopsis thaliana plants were disinfected in 75% ethanol comprising 0.05% Triton X-100 for 15 min, washed with absolute ethanol, dried, and germinated on 1/2-strength Murashige and Skoog (MS) solid media. An additional 25 mg L−1 kanamycin was used for screening the transgenic lines. When the seedlings had developed their first pair of true leaves, they were transplanted into the soil for the eventual harvesting of their seeds. To carry out the hydroponic experiment, a special device was first constructed. The two ends of a 1.5 mL centrifuge tube were removed, and the middle part was filled with 6 g L−1 agar and placed on a rectangular plastic plate. An Arabidopsis thaliana seed was placed in the center of the agar to germinate, and the device was placed in the liquid nutrient solution, which was an improved version of Hoagland solution (pH = 5.8). Each hydroponic container was placed in 4 L of nutrient solution, which was changed every 3 days. The N concentrations of the nutrient solutions were 0.15, 0.3, and 3 mM (NH4NO3 and KNO3 were at the same molar ratio). Images were collected, and the growth data were statistically significant after 25 and 45 days. Whole plants were then frozen in liquid N for further study.

2.2 Gene cloning and vector construction

On the basis of the sequence of Populus trichocarpa, we designed the following gene-specific primers for cloning: for PnDof19, 5′-AACCAATACTCACTCCTCCAACA-3′ and 5′-AGGGCACATAAAGTAACCAAATC-3′; for PnDof20, 5′-AAAGATGATTCAAGAACTCTTAGGA-3′ and 5′-AATTGTTCTTAAGGATATGCACC-3′; and for PnDof30, 5′-ACCTGGTCTTTGTCTGTTTACTCTT-3′ and 5′-CTTCCACACCTGTCTTATACCCTTG-3′. Fragment amplification and vector construction involved the use of a KOD Plus Neo high-fidelity DNA polymerase (Toyobo), a pEASY cloning vector (TransGen Biotech), and Escherichia coli Trans1-T1 sensitive cells (TransGen Biotech).

The plant transient expression vector pBS-GFP was used for subcellular localization. The primers used were as follows: for pBS-PnDof19, 5′-agggtaccATGCCGGCAGAATTA-3′ and 5′-tgactagtTTTAAGACCATTCCC-3′; for pBS-PnDof20, 5′-agggtaccATGATTCAAGAACTC-3′ and 5′- tgactagtAGGATATGCACCATT-3′; and for pBS-PnDof30, 5′-agggtaccATGATTCCTTCGAGA-3′ and 5′-tgactagtAAGAAGTACTGAAGA-3′. The 5′ end of each pair of primers was added to the KpnI and SpeI restriction sites (New England Biolab). Arabidopsis plants were genetically transformed using the plant expression vector pROK2, and the primers used for pROK2-PnDof30 included 5′-actctagaATGATTCCTTCGAGA-3′ and 5′-tgggtaccTCAAAGAAGTACTGA-3′. The 5′ end of the primers was added to the KpnI and SpeI restriction sites (New England Biolab). After restriction enzyme digestion, T4 DNA ligase (TransGen Biotech) was used for fragment ligation. The primers were synthesized and the sequencing was performed by Harbin Boshi Biotechnology.

2.3 Genome-wide analysis of the Dof transcription factor family members in Populus trichocarpa

The Phytozome (https://phytozome.jgi.doe.gov) [44] and Plant Transcription Factor Database (PlantTFDB) (http://planttfdb.cbi.pku.edu.cn) websites were queried to obtain genome-wide information concerning the Dof transcription factor family in Populus trichocarpa [45]. The website of the ProtParam tool (http://web.expasy.org/protparam/) was used to analyze the protein physicochemical properties [46], and the gene structure was analyzed via the Gene Structure Display Server 2.0 website (http://gsds.cbi.pku.edu.cn/) [47]. The WoLF PSORT website was used for subcellular localization predictions (http://www.genscript.com/psort/wolf_psort.html). The chromosome localization of the Dof genes was performed based on data from the Phytozome database and from the complete genome sequence of Populus trichocarpa, which was obtained in 2006 [48]. ClustalX software was used for multiple comparisons of protein sequences [49], and MEGA 5 software was used to construct phylogenetic trees [50]. Selecton software was used to analyze the evolutionary selection pressure (http://selecton.tau.ac.il) [51], and the mechanistic–empirical combination model was used for the analysis [52]. The AspenDB (http://aspendb.uga.edu/), poplar EFP browser (http://bar.utoronto.ca/efppop/cgi-bin/efpWeb.cgi), and Phytozome websites were used to analyze gene expression patterns, and the MeV 4.7.4 website was used to construct heat maps. The protein sequences of the cloned genes were analyzed with BioEdit multiple comparison software.

2.4 PCR, RNA extraction, and qRT-PCR

The RNA extraction reagent used in this experiment was pBIOZOL (Beijing Biomars-Technology). In addition, a PrimeScriptTM RT Reagent Kit (TaKaRa) was used for reverse transcription, an SYBR Green Real-time Quantitative Kit (CWBIO) was used to quantify the reagents, and the quantitative PCR instrument used was an ABI 7500 system. The reactions and steps were performed according to the manufacturers’ instructions.

2.5 Subcellular localization

PDS-1000 was used for subcellular localization. Microcarriers were bombarded into lower onion epidermal cells. One day after dark culture, the cells were examined via confocal laser microscopy. The nuclei were stained with 4′,6-diamidino-2-phenylindole (DAPI) reagent and imaged under fluorescent light, and combined fields.

2.6 Genetic transformation of Arabidopsis thaliana

After transforming the pROK2-PnDof30 vector into Agrobacterium tumefaciens GV3101, the gene was transformed into the genotype of Arabidopsis thaliana ecotype Col-0 plants by the floral-dip method [53]. The seeds of Arabidopsis thaliana homozygous lines were screened on 1/2-strength MS plates supplemented with 25 mg L−1 kanamycin for 3 continuous generations.

2.7 Measurements of physiological parameters

The chlorophyll, soluble protein, and soluble sugar contents and the activities of PEPC, PK, GS, and NR were determined via standard kits (Suzhou Comin Biotechnology) in accordance with the product instructions.

3 Results

3.1 Identification and bioinformatic analysis of the Dof transcription factor family members in Populus trichocarpa

Dof members usually have a conserved Dof domain. To identify all the Dof members in Populus trichocarpa, we searched the Populus trichocarpa V3.0 database for the conserved protein sequence from the Phytozome website and ultimately obtained 45 candidate sequences. After comparing the candidate sequences with the Populus trichocarpa Dof members in the PlantTFDB, the redundant sequences were removed, and 44 genes that might encode Dof transcription factors were ultimately identified. According to their chromosomal location information, these members were named PtDof01–44 (Table 1). Among these members, 20 had no introns, 21 had 1 intron, and 3 had 2 introns. The proteins encoded by these genes were 159–506 AAs in length, had a molecular weight ranging from 17.73 to 55.26 kDa, and had an isoelectric point ranging from 4.46 to 10.86. Subcellular localization prediction showed that all the members were localized in the nucleus, except PtDof19 which was localized in the mitochondria.

Table 1

Members of the Dof gene family in Populus trichocarpa

Dof gene Locus name Chromosome location Amino acids Intron number Subcellular location Mass (kDa) Pi
PtDof01 Potri.001G086400 Chr01:6823402… 6825528 285 1 Nucl 31.39 8.27
PtDof02 Potri.001G238400 Chr01:24960012… 24961681 332 1 Nucl 35.59 9.85
PtDof03 Potri.002G070700 Chr02:4885437…4886585 301 0 Nucl 34.18 4.46
PtDof04 Potri.002G129600 Chr02:9717125…9718494 306 1 Nucl 32.22 5.64
PtDof05 Potri.002G174300 Chr02:13296746…13298822 263 0 Nucl 28.90 8.93
PtDof06 Potri.003G034200 Chr03:4308607…4309970 235 0 Nucl 25.14 8.94
PtDof07 Potri.003G144500 Chr03:16086386…16088714 279 1 Nucl 30.68 8.46
PtDof08 Potri.004G038800 Chr04:2946651…2948757 304 0 Nucl 33.91 8.55
PtDof09 Potri.004G046100 Chr04:3494264…3496215 325 0 Nucl 35.55 9.20
PtDof10 Potri.004G046600 Chr04:3543422…3545548 391 1 Nucl 43.13 8.76
PtDof11 Potri.004G056900 Chr04:4529874…4530691 159 0 Nucl 17.73 9.28
PtDof12 Potri.004G121800 Chr04:11516992…11520601 503 1 Nucl 55.06 5.26
PtDof13 Potri.005G131600 Chr05:10630910…10632126 253 0 Nucl 25.99 8.58
PtDof14 Potri.005G134200 Chr05:10904734…10906329 331 1 Nucl 35.46 9.77
PtDof15 Potri.005G149100 Chr05:13304087…13305115 342 0 Nucl 37.00 8.62
PtDof16 Potri.005G188900 Chr05:20644099…20645183 301 0 Nucl 34.04 4.56
PtDof17 Potri.006G084200 Chr06:6374602…6376438 326 1 Nucl 34.63 9.02
PtDof18 Potri.006G202500 Chr06:21732370…21733236 288 0 Nucl 31.91 6.77
PtDof19 Potri.007G036400 Chr07:2818037…2819446 248 0 Mito 25.50 8.37
PtDof20 Potri.007G038100 Chr07:2988100…2989665 323 0 Nucl 34.16 8.83
PtDof21 Potri.007G058200 Chr07:6232778…6234228 344 0 Nucl 37.08 8.09
PtDof22 Potri.008G055100 Chr08:3254228…3256078 345 2 Nucl 36.90 9.09
PtDof23 Potri.008G087800 Chr08:5489825…5493164 500 1 Nucl 54.06 6.95
PtDof24 Potri.009G029500 Chr09:4021928…4024120 326 1 Nucl 34.65 9.43
PtDof25 Potri.010G167600 Chr10:17052686…17055696 496 1 Nucl 54.19 7.28
PtDof26 Potri.010G205400 Chr10:19627248…19629172 356 2 Nucl 37.53 9.51
PtDof27 Potri.011G047500 Chr11:4061272…4063211 305 0 Nucl 33.89 8.26
PtDof28 Potri.011G054300 Chr11:4716002…4717173 325 0 Nucl 35.71 9.35
PtDof29 Potri.011G054400 Chr11:4740021…4740708 170 0 Nucl 19.61 10.86
PtDof30 Potri.011G055600 Chr11:4863648…4865957 357 1 Nucl 39.20 8.88
PtDof31 Potri.011G065900 Chr11:6038802…6040217 165 1 Nucl 18.29 8.82
PtDof32 Potri.012G018700 Chr12:1730440…1732794 297 1 Nucl 32.79 7.89
PtDof33 Potri.012G063800 Chr12:7683888…7685455 329 0 Nucl 36.23 6.86
PtDof34 Potri.012G081300 Chr12:10796766…10799160 312 1 Nucl 34.20 6.90
PtDof35 Potri.013G066700 Chr13:5186342…5190516 494 1 Nucl 53.79 6.98
PtDof36 Potri.014G036600 Chr14:2984228…2985710 261 0 Nucl 27.53 6.38
PtDof37 Potri.014G100900 Chr14:7895311…7897106 229 1 Nucl 25.08 9.21
PtDof38 Potri.015G009300 Chr15:611499…613350 255 1 Nucl 28.12 8.69
PtDof39 Potri.015G048300 Chr15:5093143…5095147 321 0 Nucl 35.23 7.61
PtDof40 Potri.015G077100 Chr15:10212436…10214743 314 1 Nucl 34.79 6.65
PtDof41 Potri.016G069300 Chr16:5006439…5007305 225 2 Nucl 25.21 6.72
PtDof42 Potri.017G084600 Chr17:10185927…10189438 506 1 Nucl 55.26 5.36
PtDof43 Potri.019G040700 Chr19:4633699…4637172 493 1 Nucl 53.50 5.37
PtDof44 Potri.T146000 scaffold_455:16693…17845 274 0 Nucl 30.87 4.80

Multiple alignments of the Dof domain sequences of the Dof transcription factor members showed that 43 Dof domains were conserved; these domains comprised 54–55 AAs and especially included cysteines at sites 3, 6, 28, and 31. These four cysteines are essential for the formation of zinc-finger structures (Figure 1). Notably, the Dof domain of the PtDof29 protein is incomplete and does not have a typical C2–C2 zinc-finger structure.

Figure 1 
                  Multiple alignment of 44 conserved PtDof protein sequences. Two sets of four cysteines highlighted in yellow form a zinc-finger structure, and the underlined area is the Dof domain.
Figure 1

Multiple alignment of 44 conserved PtDof protein sequences. Two sets of four cysteines highlighted in yellow form a zinc-finger structure, and the underlined area is the Dof domain.

To study the evolutionary relationships among members of the Dof family, we constructed a neighbor-joining (NJ) phylogenetic tree of the Dof protein sequences via MEGA 5 software (Figure 2). The results showed that the Dof members of poplar could be divided into 4 subfamilies (I, II, III, and IV) that contained 15, 11, 8, and 9 Dof members, respectively. Although the two branches of the fourth subfamily were not on the same trunk, the evolutionary relationships were very similar between each other, and the gene structures were similar; thus, they were combined into one subfamily. PtDof29 is relatively independent and does not belong to any subfamily. The gene exon and intron structure maps effectively support the classification results of the subfamily.

Figure 2 
                  Phylogenetic relationships and gene structure of Dof family members in Populus trichocarpa. The NJ phylogenetic tree on the left was constructed via MEGA 5; the tree comprises the aligned protein sequences of 44 PtDof members, and the 4 subfamilies are named I, II, III, and IV. The gene exon/intron structure is shown on the right. The blue lines represent untranslated regions, the yellow lines represent coding areas, and the thin lines represent introns.
Figure 2

Phylogenetic relationships and gene structure of Dof family members in Populus trichocarpa. The NJ phylogenetic tree on the left was constructed via MEGA 5; the tree comprises the aligned protein sequences of 44 PtDof members, and the 4 subfamilies are named I, II, III, and IV. The gene exon/intron structure is shown on the right. The blue lines represent untranslated regions, the yellow lines represent coding areas, and the thin lines represent introns.

The coding DNA sequence (CDS) of the Populus trichocarpa Dof gene was input into the Selecton server, and a selection pressure map of each site was obtained (Figure 3). In the maps, yellow represents a positive selection site with less distribution; there was no positive selection site distributed within the Dof domain. White to purple represents negative selection sites; almost all AAs in the Dof domain are associated with a negative selection site, and most of them are dark purple on the map, representing substantial purifying selection. These results indicate that the Dof domain is under a substantial amount of purifying selection, maintaining a high degree of evolutionary rigor; most of the non-Dof domains are under neutral selection. Compared with the other sites, these sites are less conserved and have a higher probability of mutation. This conclusion also explains why the sequences of the Dof family members are quite different.

Figure 3 
                  Evolution pressure analysis of PtDofs in Populus trichocarpa. The black underlined area represents the Dof domain.
Figure 3

Evolution pressure analysis of PtDofs in Populus trichocarpa. The black underlined area represents the Dof domain.

Based on the chromosomal location information of the 44 Dof genes whose sequence is on the Phytozome website and the homologous recombination map published in 2006, we mapped the chromosomal location map of the Populus trichocarpa Dof genes (Figure 4); notably, PtDof29 was unable to be mapped because it belonged to the scaffold structure. The results showed that the Dof genes were distributed throughout the chromosomes, indicating that Dof genes might be ancient. In general, genes within the homologous recombination region of a chromosome may originate from the same ancestor gene. After combining these results with the results of the phylogenetic tree, we identified nine pairs of genes that may have been generated by homologous recombination events of chromosomes in recent evolutionary years: PtDof1/PtDof7, PtDof2/PtDof24, PtDof4/PtDof36, PtDof5/PtDof37, PtDof11/PtDof31, PtDof22/PtDof26, PtDof23/PtDof25, PtDof32/PtDof38, and PtDof34/PtDof40. These nine gene pairs were present not only in the homologous recombination region of the same chromosome but also in the same branch of the evolutionary tree. Among the Dof genes in Populus trichocarpa, these nine pairs are most homologous. Therefore, we speculate that these nine pair of genes may have originated from homologous recombination events throughout the evolution.

Figure 4 
                  Chromosome mapping of Dof family members in Populus trichocarpa. The same colored regions represent chromosomal homologous recombination regions.
Figure 4

Chromosome mapping of Dof family members in Populus trichocarpa. The same colored regions represent chromosomal homologous recombination regions.

After searching the AspenDB website for sequences of gene probes for the Populus trichocarpa Dof family members, we searched the expression data of each gene in the EFP database. In total, 34 Dof gene expression data points were ultimately identified and used to construct an expression map (Figure 5). The results showed that the expression of the Dof genes was mostly downregulated in the mature leaves. However, in the young leaves, the expressions of 9 genes was downregulated, and the expressions of the other 25 genes were upregulated. The expressions of the PtDof10, PtDof18, PtDof20, and PtDof41 genes were relatively high, and the expression of the PtDof19 gene was the highest. In the roots, the expressions of 7 genes was downregulated, and that of 27 genes were upregulated, of which the expressions of PtDof5, PtDof14, and PtDof20 were the highest. In the young leaves of plants growing in the darkness, the expressions of 15 genes was upregulated, and that of 19 genes were downregulated. In the young leaves of plants growing in darkness but then exposed to light for 3 h, the expressions of 12 genes was upregulated, and that of 22 genes were downregulated. In seedlings subjected to continuous light, the expressions of 17 genes were upregulated, and that of 17 genes were downregulated. In the female flowers, the expressions of 13 genes were upregulated, and that of 21 genes were downregulated, and in the male flowers, the expressions of 20 genes were upregulated, and that of 14 genes were downregulated. In the xylem, the expressions of 11 genes was upregulated, and that of 33 genes were downregulated. In conclusion, most of the members of the Populus trichocarpa Dof gene family were expressed in young leaves and roots, and the expression patterns in other plant parts were more complex, which indicated that the function of Dof genes might be substantially different.

Figure 5 
                  Expression patterns of 34 Populus trichocarpa Dof members based on EFP data. ML: Mature leaves; YL: Young leaves; Rt: Roots; Ds: Dark-grown seedlings; Ds3h: Dark-grown seedlings exposed to light for 3 h; Cls: Continuous light-grown seedlings; fc: Female catkins; mc: Male catkins; xy: Xylem.
Figure 5

Expression patterns of 34 Populus trichocarpa Dof members based on EFP data. ML: Mature leaves; YL: Young leaves; Rt: Roots; Ds: Dark-grown seedlings; Ds3h: Dark-grown seedlings exposed to light for 3 h; Cls: Continuous light-grown seedlings; fc: Female catkins; mc: Male catkins; xy: Xylem.

The fragments per kilobase of transcript per million mapped reads (FPKM) data of the Populus trichocarpa Dof genes were obtained from the Phytozome website, and the expression data were used to construct a gene expression heat map (Figure 6). The results showed that the expressions of PtDof10 and PtDof30 were higher in the leaves (early stage of female floral buds) than in the other organs. In the leaves (immature ones), the most highly expressed genes were PtDof30 and PtDof19. In the young leaves, PtDof10 and PtDof30 were highly expressed; in the roots, the PtDof5 gene was expressed the most. The most highly expressed genes in the root tips were PtDof10 and PtDof30. In the stems (internodes), the most highly expressed genes were PtDof32 and PtDof39; in the stem nodes, the most highly expressed genes were PtDof19, PtDof10, PtDof38, and PtDof39. The expression of PtDof05, PtDof10, PtDof19, PtDof20, PtDof23, PtDof30, and PtDof39 in various tissues was significantly higher than that of the other studied genes.

Figure 6 
                  Expression patterns of 44 Populus trichocarpa Dof members based on FPKM data obtained from the Phytozome website. Lf: Leaves (early-stage female floral buds); Li: Leaves (immature); Ly: Leaves (young); Rt: Roots; RtT: Root tips; Si: Stems (internodes); Sn: Stems (nodes).
Figure 6

Expression patterns of 44 Populus trichocarpa Dof members based on FPKM data obtained from the Phytozome website. Lf: Leaves (early-stage female floral buds); Li: Leaves (immature); Ly: Leaves (young); Rt: Roots; RtT: Root tips; Si: Stems (internodes); Sn: Stems (nodes).

3.2 Screening of Dof genes in response to N changes in Populus simonii × Populus nigra

Two groups of tissue culture-generated seedlings were subjected to N-treatment experiments under LA hydroponic solution. In the first group, 1 mM ammonium-nitrate was used as the sole N source (Figure 7). After 1 week of cultivation, the seedlings were subjected to an N-deficient solution for 3 days. Afterward, they were subjected to an N-sufficient solution for 2 or 48 h, after which the N was withheld again for 2 or 48 h. Samples were taken at each of these time points. Populus simonii × Populus nigra seedlings were grown under N-deficient conditions in vitro for 3 days as controls and were provided different forms of N in vivo only to maintain biological activity. After the N was absorbed under in vivo conditions, the N treatments were carried out in vitro to obtain information on Dof genes induced in response to N in Populus simonii × Populus nigra. qRT-PCR was used to measure the expression of the Dof genes (the primers used are listed in Table 1). The internal reference gene used was cell division control protein 2 (CDC2), and the 2−ΔΔCt method was used to calculate the relative expression. After several rounds of designing primers and performing quantitative experiments, the expression data of 38 genes were ultimately obtained. The results showed that the expression patterns of these genes in the leaves, stems, and roots were very different under different treatments, showing different degrees of tissue specificity. Compared with that in the control group, the expressions of 31 Dof genes in the leaves of the treatment group increased 2 h after the N supply was restored. After the N supply was restored for 48 h, the expression levels were lower than that after the N supply was restored for 2 h, and the expression levels of most of the Dof genes were lower after N was withheld than when the N was supplied for 2 and 48 h. It could be concluded that a short N supply induces the expressions of most Dof genes in the leaves but that a prolonged N supply inhibits the expressions of some of these genes. In the stems, 32 Dof genes presented higher expression levels when N was withheld for 48 h but lower expression levels when N was resupplied for 2 and 48 h. These results suggest that low N contents in vivo induced the expressions of Dof genes in the absence of an N supply in vitro. In the roots, the expressions of 34 Dof genes decreased 2 h after the N was resupplied, while the expressions of 17 Dof genes increased slightly within the 48 h during which the N was resupplied and withheld again. This may indicate that the expressions of some Dof genes was induced under low N levels in the roots.

Figure 7 
                  Relative expression levels of the Populus simonii × Populus nigra PnDof genes under sufficient and deficient N supplies, as revealed by qRT-PCR (L: leaves; S: stems; R: roots).
Figure 7

Relative expression levels of the Populus simonii × Populus nigra PnDof genes under sufficient and deficient N supplies, as revealed by qRT-PCR (L: leaves; S: stems; R: roots).

In the second N treatment experiment, seedlings were treated with ammonium and nitrate N for 2 weeks at concentrations of 0.1, 1, or 10 mM (Figure 8). The results showed that the expressions of ten genes (Dof4, Dof10, Dof11, Dof13, Dof21, Dof28, Dof30, Dof32, Dof42, and Dof43) were induced in the leaves under low ammonium concentrations, that of eight genes (Dof4, Dof9, Dof19, Dof26, Dof30, Dof36, Dof37, and Dof40) were induced in the stems, and that of seven genes (Dof7, Dof9, Dof10, Dof21, Dof25, Dof36, and Dof42) were induced in the roots. Under low-nitrate conditions, the expressions of nine genes were induced in leaves, including Dof4, Dof10, Dof13, Dof21, Dof23, Dof30, Dof32, Dof33, and Dof36; the expressions of four genes were induced in stems, including Dof26, Dof28, Dof36, and Dof43; and the expressions of 19 genes were induced in the roots. In conclusion, the expressions of the following genes was induced under both low levels of N at the same time: Dof4, Dof10, Dof13, Dof21, Dof30, and Dof32 in leaves; Dof36 in the stems; and Dof25, Dof36, and Dof42 in the roots.

Figure 8 
                  Relative expression levels of Populus simonii × Populus nigra PnDof genes under two kinds of N supplied at three different concentrations, as revealed by qRT-PCR (L: Leaves; S: Stems; R: Roots).
Figure 8

Relative expression levels of Populus simonii × Populus nigra PnDof genes under two kinds of N supplied at three different concentrations, as revealed by qRT-PCR (L: Leaves; S: Stems; R: Roots).

3.3 Cloning of the PnDof19, PnDof20, and PnDof30 genes from Populus simonii × Populus nigra

We cloned the CDSs of the PnDof19 (MK796000), PnDof20 (MK7960001), and PnDof30 (MK789595) genes from Populus simonii × Populus nigra cDNA (Figure 9), which were 747, 972, and 1,074 bp long, respectively, and encoded 248, 323, and 357 AAs, respectively.

Figure 9 
                  Gene coding sequences of PnDof19, PnDof20, and PnDof30 alongside their translated protein sequences in Populus simonii × Populus nigra.
Figure 9

Gene coding sequences of PnDof19, PnDof20, and PnDof30 alongside their translated protein sequences in Populus simonii × Populus nigra.

To explore whether the transcription factors encoded by the PnDof19, PnDof20, and PnDof30 genes cloned from Populus simonii × Populus nigra are involved in regulating C and N metabolism, we compared the sequences of these three proteins with the sequences of three functional proteins that specifically regulate C and N metabolism (Figure 10). The results showed that all six proteins had a complete Dof domain and four highly conserved cysteines, and all of them had a nuclear localization signal specific to Dof transcription factor family members (B1 and B2 regions).

Figure 10 
                  Multiple alignment of protein sequences. The black box area is the conserved Dof domain, and B1 and B2 represent Dof-specific nuclear localization signals.
Figure 10

Multiple alignment of protein sequences. The black box area is the conserved Dof domain, and B1 and B2 represent Dof-specific nuclear localization signals.

Moreover, we identified Dof genes with different functions (both AtDof1/NM_104048.4 and AtOBP1/OAP05220.1 in Arabidopsis thaliana and OsDof12/AAL84292.1 in rice) and constructed phylogenetic trees comprising the CDSs of both the three protein-coding genes cloned by us and the known Dof protein-coding genes (Figure 11). Because of the poor conservation of Dof protein sequences of non-Dof-domain regions, we used only conserved domain sequences to construct phylogenetic trees to determine their evolutionary relationships more accurately. The results showed that PnDof20 and PnDof30 branched together with the N metabolism regulatory genes AtDof1 and PpDof5, suggesting that PnDof20 and PnDof30 may also be N metabolism regulatory genes; PnDof19 and the cell cycle regulatory gene AtOBP1 were branched together on one branch, and thus, we speculated that PnDof19 might be related to cell cycle regulation. OsDof12 is a flowering regulatory gene that is independent of the other genes in the phylogenetic tree.

Figure 11 
                  Phylogenetic tree comprising PnDof19, PnDof20, and PnDof30 in Populus simonii × Populus nigra and several specific functional Dof genes in other species.
Figure 11

Phylogenetic tree comprising PnDof19, PnDof20, and PnDof30 in Populus simonii × Populus nigra and several specific functional Dof genes in other species.

3.4 Subcellular localization of the PnDof19, PnDof20, and PnDof30 proteins

To determine whether the three Populus simonii × Populus nigra Dof proteins have characteristics of general transcription factors, i.e., the localization of the protein in the nucleus, we fused the open reading frame of the PnDof19, PnDof20, and PnDof30 genes to the GFP gene within a PBS-GFP vector; onion subepidermal cells were subsequently transformed via gene gun bombardment. Cells displaying green fluorescence were observed via scanning laser confocal microscopy. The nuclei were stained with DAPI reagent and then observed and imaged under a microscope (Figure 12). The results showed that the PnDof19, PnDof20, and PnDof30 proteins localized to the nucleus, which is consistent with the localization of Dof proteins in other reported species [54,55].

Figure 12 
                  Subcellular localization of the PnDof19, PnDof20, and PnDof30 proteins. a: Fluorescence field; b: Bright field; and c: Superimposition of the fluorescence and bright fields.
Figure 12

Subcellular localization of the PnDof19, PnDof20, and PnDof30 proteins. a: Fluorescence field; b: Bright field; and c: Superimposition of the fluorescence and bright fields.

3.5 Functional analysis of PnDof30-overexpressing Arabidopsis thaliana lines

The PnDof30 gene was inserted into the genome of Arabidopsis thaliana ecotype Col-0 plants by the floral-dip method, and 24 independent transgenic lines were selected for extraction of their genomic DNA. The transgenic lines were identified via PCR (Figure 13). After identification, 11 transgenic lines were randomly selected to determine the expression level of their PnDof30 gene (Figure 14). The results showed that the expression level and stability in each line were substantially different. We chose three stable expression lines, L1, L2, and L15, for functional analysis and then screened and identified the homozygotes.

Figure 13 
                  PCR-based identification of genomic DNA in transgenic Arabidopsis thaliana lines.
Figure 13

PCR-based identification of genomic DNA in transgenic Arabidopsis thaliana lines.

Figure 14 
                  Relative expression of PnDof30 in transgenic Arabidopsis thaliana lines.
Figure 14

Relative expression of PnDof30 in transgenic Arabidopsis thaliana lines.

The seeds of the WT and homozygous lines were vernalized and planted in several unique hydroponic devices for germination and growth. The hydroponic solutions used were improved versions of Hoagland solution, consisting of 0.15, 0.3, or 3 mM N. The phenotypes were evaluated after 25 days of plant growth (Figure 15). The results showed that the phenotypes of the growth-related parameters of the PnDof30-overexpressing Arabidopsis thaliana lines, such as leaf size, lotus leaf diameter, and leaf number, were significantly better than those of WT plants under all three different N concentrations, especially under low-N conditions.

Figure 15 
                  Phenotypic changes in PnDof30 transgenic Arabidopsis thaliana lines under N treatment for 25 days.
Figure 15

Phenotypic changes in PnDof30 transgenic Arabidopsis thaliana lines under N treatment for 25 days.

Because the plant size was too small for further determination of growth-related parameters, we cultured Arabidopsis thaliana plants in liquid media for 45 days and observed their phenotype (Figure 16). The results showed that when the culture period was extended to 45 days, the overexpression plants under low-N conditions still grew better than WT plants, and the leaf color was greener than that of WT plants. Under 3 mM N, the difference in growth between the overexpression lines and WT narrowed; nonetheless, the L2 transgenic lines were significantly better than the WT, and L1 and L3 transgenic lines were slightly better than the WT.

Figure 16 
                  Phenotypic changes in PnDof30 transgenic Arabidopsis thaliana lines under N treatment for 45 days.
Figure 16

Phenotypic changes in PnDof30 transgenic Arabidopsis thaliana lines under N treatment for 45 days.

The diameter of the leaves, the number of leaves, and fresh weight were further determined (Figure 17). Compared with that of the WT plants, the diameter of the lotus leaves of the L1 plants increased significantly at a 0.3 mM N concentration, and the diameter of the lotus leaves of the L2 and L3 plants increased significantly at all three N concentrations, which was consistent with the observed phenotypes. The number of leaves in the L2 and L3 plants was significantly different from that of WT plants at the 0.15 mM N concentration, and the L2 plants had significantly more leaves than the WT plants at the 0.3 and 3 mM N concentrations. Although the number of leaves was not significantly different between the L1 and L3 plants and the WT plants, we observed that the fresh weight of the three transformed lines was significantly higher than that of the WT plants at the three concentrations, and the difference between L2 and the WT was the most significant. Taken together, these results indicated that the PnDof30 gene could improve the growth index of Arabidopsis thaliana, especially under low-N levels.

Figure 17 
                  Rosette leaf diameter, leaf number, and fresh weight of Arabidopsis transgenic lines and WT plants subjected to 3 different N concentrations for 45 days (p < 0.05).
Figure 17

Rosette leaf diameter, leaf number, and fresh weight of Arabidopsis transgenic lines and WT plants subjected to 3 different N concentrations for 45 days (p < 0.05).

To further evaluate the effect of overexpression of the PnDof30 gene on the growth of Arabidopsis thaliana, we measured the contents of soluble sugars, soluble protein, and chlorophyll (Figure 18). The results showed that the soluble protein contents in the three transgenic lines were higher than those in the WT at the 0.15, 0.3, and 3 mM N concentrations, while the contents in the L3 plants at the 0.15 mM N concentration and in the L1 and L3 plants at 0.3 mM N concentrations were not significantly different from those in the WT, but they were slightly higher. Under the 0.15 mM N concentration, the soluble sugar content in the three transgenic lines decreased significantly; under the 0.3 mM N concentration, the content in the L1 plants decreased slightly, whereas in the L2 and L3 plants, the contents decreased significantly. However, there was no significant difference between the transgenic and WT plants at the 3 mM N concentration. Under the three N concentrations, the chlorophyll content in the transgenic lines was significantly higher than that in the WT plants. The increase in soluble protein content and the decrease in soluble sugar content indicated that the efficiency of N utilization improved and that C skeletons were consumed in the transgenic Arabidopsis thaliana lines, which resulted in a decrease in the soluble sugar content and an increase in both the soluble protein content and the chlorophyll content, which effectively promoted photosynthesis and C/N metabolism. Taken together, these results indicate that the PnDof30 gene can increase the C/N metabolic level in transgenic Arabidopsis thaliana, especially under low N levels.

Figure 18 
                  Contents of soluble sugars, soluble proteins, and chlorophyll in Arabidopsis transgenic lines and WT plants subjected to 3 different N concentrations for 45 days (p < 0.05).
Figure 18

Contents of soluble sugars, soluble proteins, and chlorophyll in Arabidopsis transgenic lines and WT plants subjected to 3 different N concentrations for 45 days (p < 0.05).

The activities of the PEPC, PK, GS, and NR enzymes were subsequently determined (Figure 19). According to the PEPC enzyme activity results, the activities in the three transformed lines were significantly higher than that in the WT plants at 0.15 mM N. Under 0.3 mM N, the activity in the L1 plants was slightly higher than that in the WT plants, and that in the L2 and L3 plants was significantly higher than that in the WT plants. Under 3 mM N, the enzyme activity in the L2 plants was significantly higher than that in the WT plants, while the enzyme activity in the L1 and L3 plants did not significantly differ from that in the WT plants.

Figure 19 
                  Enzyme activity of PEPC, PK, GS, and NR of Arabidopsis transgenic lines and WT under 3 concentrations of N treatment for 45 days (p < 0.05).
Figure 19

Enzyme activity of PEPC, PK, GS, and NR of Arabidopsis transgenic lines and WT under 3 concentrations of N treatment for 45 days (p < 0.05).

With respect to PK enzyme activity, the results showed that at 0.15 mM N concentration, the activities in the L1 and L2 plants were slightly higher and that in the L3 plants was significantly higher than that in the WT plants. Under 0.3 mM N, compared with the WT plants, the three transformed lines presented significantly higher PK activity. Under 3 mM N, the activity in the L1 plants was significantly higher than that in the WT plants, and the activities in the L2 and L3 were not significantly different from that in the WT plants.

In terms of GS enzyme activity, the results showed that the three transformed lines did not significantly differ from the WT plants at the 0.15 mM N concentration but that the activity in the former was significantly higher than that in the WT plants at the 0.3 and 3 mM N concentrations.

In terms of NR enzyme activity, the results showed that the activities of the three transgenic lines were significantly higher than those in the WT plants at all three N concentrations.

PEPC and PK are important enzymes involved in the process of C metabolism. The activities of the PEPC and PK enzymes in the transgenic lines increased under low-N conditions, indicating that the metabolism of C increased. Reactions involving GS constitute the first step of ammonium assimilation, and reactions involving NR constitute the first step of nitrate assimilation. The enzyme activities in all transgenic lines improved under low-N conditions. The results of the enzyme activity assay showed that overexpression of PnDof30 in Arabidopsis thaliana could promote C/N assimilation efficiency under low-N conditions.

We selected 13 genes that play a major role in the C/N pathways and measured their relative expression (Figure 20). Under 0.15 mM N, the expression of the PEPC1 gene in the three transgenic lines was not significantly different from that in the WT plants, but the expression of PEPC2 significantly increased. Under 0.3 and 3 mM N, the expressions of PEPC1 and PEPC2 in the transgenic lines increased. The expressions of PK1 and PK2 in the three transgenic lines significantly increased under all three N concentrations. The expression changes of the PEPC and PK genes were consistent with the results of the PEPC and PK enzyme activities.

Figure 20 
                  Relative expression levels of the PEPC and PK genes in Arabidopsis transgenic lines and WT plants subjected to 3 different concentrations of N for 45 days (p < 0.05).
Figure 20

Relative expression levels of the PEPC and PK genes in Arabidopsis transgenic lines and WT plants subjected to 3 different concentrations of N for 45 days (p < 0.05).

The ammonium transporter (AMT) protein data showed that the expression of 3 genes in the 3 transgenic lines significantly increased under 3 mM N (Figure 21). Under low-N conditions, the expressions of AMT1.1 and AMT1.2 were downregulated, and that of AMT1.3 was upregulated. All three genes were major functional genes involved in the high-affinity ammonium transport system in Arabidopsis thaliana; however, the contribution of each gene was unknown, so it was uncertain whether ammonium uptake increased in general.

Figure 21 
                  Relative expression levels of AMT genes in Arabidopsis transgenic lines and WT plants subjected to 3 different concentrations of N for 45 days (p < 0.05).
Figure 21

Relative expression levels of AMT genes in Arabidopsis transgenic lines and WT plants subjected to 3 different concentrations of N for 45 days (p < 0.05).

The nitrate transporter (NRT) data showed that the expression of two NRT genes increased in all three transgenic lines under 3 mM N, but the expression of NRT1.1 was downregulated and that of NRT2.1 was upregulated under low-N conditions (Figure 22). Like with the AMT gene, it was unclear whether the nitrate transport level improved.

Figure 22 
                  Relative expression levels of NRT genes in Arabidopsis transgenic lines and WT plants subjected to 3 different concentrations of N for 45 days (p < 0.05).
Figure 22

Relative expression levels of NRT genes in Arabidopsis transgenic lines and WT plants subjected to 3 different concentrations of N for 45 days (p < 0.05).

The GS data showed that the expression of GS1.1 in the three transgenic lines was significantly higher than that in the WT plants at all three N concentrations (Figure 23). The expressions of GS1.2, GS1.3, and GS2 were upregulated under 0.3 and 3 mM N; however, the expression of GS1.2 was upregulated while the expressions of both GS1.3 and GS2 were downregulated at 0.15 mM N. Under 0.15 mM N, GS1.2 expression in the transgenic lines was not significantly different from that in the WT plants, and GS1.3 and GS2 expression was downregulated. We speculate that 0.15 mM N is the key concentration responsible for GS1.2, GS1.3, and GS2 gene expressions. The change in GS gene expression under low-N levels is consistent with the change in GS enzyme activity.

Figure 23 
                  Relative expression levels of GS genes in Arabidopsis transgenic lines and WT plants subjected to 3 different concentrations of N for 45 days (p < 0.05).
Figure 23

Relative expression levels of GS genes in Arabidopsis transgenic lines and WT plants subjected to 3 different concentrations of N for 45 days (p < 0.05).

4 Discussion

Dof transcription factor family members are specific to plants. Yanagisawa and Izui first identified the gene whose encoded protein contains the Dof domain in 1993 [56]. Dof transcription factors have a variety of functions. Studies have shown that Dof transcription factors regulate the expression of many genes involved in C/N metabolic pathways, promote plant growth, and improve N use efficiency under low-N conditions. However, most of the recent studies on the ability of Dof transcription factors to improve plant N use efficiency have focused on model organisms or crop species with simple genomic backgrounds, such as maize, rice, and Arabidopsis thaliana, while the function of Dof transcription factors in forest tree species has rarely been investigated.

In 2006, Yang and Tuskan identified 41 Dof genes from Populus trichocarpa based on the V1.0 database [57] (Appendix Table A1). With the release of the Populus trichocarpa V2.2 database and the development of bioinformatics technology, Wang re-identified the Dof family members in Populus trichocarpa in 2017. Although the number of members identified was still 41, the content and depth of the study were improved, but the focus was mainly on osmotic stress [58]. Based on the new Populus trichocarpa V3.0 database, 44 members of the Dof gene family were identified in this study, three more than the total number previously identified. Except for the PtDof29, all the members contain a highly conserved Dof domain that includes four cysteines that constitute a zinc-finger structure, which is an important component of transcription factors. We divided all the members into four subfamilies according to the evolutionary relationships between the genes. We identified nine pairs of genes originating from homologous recombination events by performing a phylogenetic analysis and using chromosome mapping data. Gene expression pattern analysis revealed that most of the Dof genes were expressed in young leaves, stems, and roots.

In this study, the Dof genes of Populus trichocarpa after N treatment were screened via qRT-PCR. Finally, we screened three genes, PnDof19, PnDof20, and PnDof30, which responded to changes in N and whose expression changed with changes in N concentration.

We cloned the CDSs of the PnDof19, PnDof20, and PnDof30 genes from Populus trichocarpa cDNA and fused the sequences to the GFP gene for subcellular localization experiments. The results showed that all three transcription factors were localized in the nucleus, which was consistent with the localization of Dof proteins in other species [54,55].

N treatment was carried out on homozygous lines of PnDof30 transgenic Arabidopsis thaliana. The results showed that the transgenic Arabidopsis thaliana plants grew better than the WT plants under low-N conditions; the soluble protein and chlorophyll contents significantly increased, while the soluble sugar content significantly decreased. These results were consistent with those of Yanagisawa et al.’s research [31]. PEPC catalyzes the reaction of phosphoenolpyruvic acid with HCO3 to form oxaloacetic acid, a supplemental substrate of the tricarboxylic acid cycle [59]. PK catalyzes the production of pyruvic acid from phosphoenolpyruvate. These two enzymes are key enzymes involved in the process of C assimilation. Both the gene expression and enzyme activity of PEPC and PK significantly increased in the transgenic lines, indicating that overexpression of Dof genes increased the level of C metabolism. The reduction of nitrate to nitrite catalyzed by NR is the first step in nitrate assimilation, and NR gene expression and enzyme activity significantly increased. AMT and NRT are important transporters of inorganic N absorbed by plant roots. The expression of some of the major genes encoding both AMT and NRT was upregulated, and the expression of some was downregulated. Although it was unclear which gene contributes more to the uptake of inorganic N, a higher soluble protein content meant that transgenic Arabidopsis might have a higher overall N uptake efficiency. GS1.3 and GS2 expressions and enzyme activities were downregulated under 0.15 mM N, but the opposite results were observed under 0.3 mM N. We speculated that the 0.15 mM N concentration was the key regulatory concentration for GS1.3 and GS2.

The innovation of this study is that, first, we re-identified the Dof family of Populus trichocarpa, which has three more members than in previous studies; second, we cloned three Dof members in Populus simonii: PnDof19, PnDof20, and PnDof30; and third, in 2013, Lin transformed maize ZmDof1 into poplar and found that plant growth indicators and N assimilation were not improved at low-N levels. It is speculated that maize ZmDof1 is not suitable for poplar, a forest plant. The gene PnDof30, which can improve the growth index and C/N metabolism-related physiological index of Arabidopsis under low-N levels, was cloned by itself. This gene can be used as an important alternative tool to improve the growth state of poplar under a low-N environment.

5 Conclusion

This study focused on the expression of Dof gene in roots and leaves by N treatment of Populus nigra, and whether they were induced by N. According to the Dof gene expression heat map constructed from the EFP database in the bioinformatics chapter, we found 13 genes with high expression levels in leaves and roots: PtDof4/10/12/14/19/20/27/28/32/34/36/38/43; On the heat map made by the FPKM database, there are six genes with high expression in roots and leaves at the same time PtDof5/10/19/20/30/32; up to 16 genes were induced by N in leaves in experiments with supply interruption and restoration: PtDof6/8/10/11/20/21/23/24/26/27/28/30/32/35/37/40, while there are only two genes induced by N in roots: PtDof3/16, which indicates that the expression of Dof gene that plays a role in leaves is mostly induced by nitrogen, while the expression of Dof gene in roots is composed of type expression.

Based on the above expression data, we selected six genes, PtDof10/19/30 (subfamily I), PtDof32 (subfamily II), PtDof12 (subfamily III), and PtDof20 (subfamily IV), as candidate genes for functional studies. Through repeated gene cloning experiments, we finally cloned the three genes PnDof19, PnDof20, and PnDof30 for subsequent experiments. The Dof gene of P. nigra that responds to N was screened out by quantitative PCR. These three genes responded to changes in N, and their expression varied with N concentrations.

Through the method of molecular biology, we cloned PnDof19, PnDof20, and PnDof30 genes. The length of the open reading frame of the PnDof19 gene is 744 bp, encoding 248 AAs; the length of the ORF region of the PnDof20 gene is 969 bp, encoding 323 AAs; The ORF region of the PnDof30 gene is 1,071 bp in length and encodes 357 AAs. After cloning, we first fused these genes with the green fluorescent protein gene GFP and expressed them in the onion lower epidermis by gene gun transient transformation to explore the subcellular localization of these three transcription factors localized in the nucleus.

To further explore whether these three Dofs can promote N assimilation and regulate C/N metabolism, considering that Populus simonii × Populus nigra belongs to the Aigeiros segment, the transformation is difficult, so we constructed these three genes into the plant overexpression vector pROK2, and tried to transform Arabidopsis thaliana. After many attempts, only PnDof30 was successfully transformed into Arabidopsis. After repeated verification and screening, a homozygous line for follow-up research was successfully obtained.

According to the above reviews, the overexpression of PnDof30, a member of the Dof family in Populus simonii × Populus nigra, could promote the growth of Arabidopsis and increase the level of C/N metabolism under low-N conditions. Therefore, the Dof gene in Populus simonii × Populus nigra may be used as an important candidate to improve the growth of poplar under low-N conditions.

6 Author summary

In this study, we treated Populus simonii × Populus nigra with N, and focused on observing the expression of Dof gene in roots and leaves, and testing whether they are induced by N. Next we cloned the three genes that respond to N and constructed them into a plant expression vector. Finally, one of the genes was successfully expressed. In the subsequent low-N treatment, it was also confirmed that this gene can promote plant growth and increase the level of C/N metabolism.

Compared with the predecessors, the results of this study show that the genes of the Dof family of Populus trichocarpa have increased significantly. In addition, we also cloned three of the genes with good results. In addition, we cloned the gene PnDof30 from the small black poplar itself that can improve the growth indicators of Arabidopsis thaliana and the physiological indicators related to C/N metabolism under low-N levels. This gene can be used as a way to improve the growth status of poplars in low-N environments.

Acknowledgements

We would like to thank L.G.J. (Liu Guanjun), a teacher from the State Key Laboratory of Tree Genetics and Breeding, Northeast Forestry University, for his selfish help and financial support for this experiment.

  1. Funding information: This work was supported by the Fundamental Research Funds for the Central Universities (No. 2572019BA08).

  2. Author contributions: Y.C.J. and W.S.M. conceived and designed the experiments; W.R.N. and W.S.M. performed the experiments and wrote the article; W.R.N. analyzed the data; L.G.J. contributed reagents/materials/analysis tools.

  3. Conflict of interest: Authors state no conflict of interest.

  4. Data availability statement: The datasets generated during and/or analyzed during the current study are available from the corresponding author on reasonable request.

Appendix

Table A1

Dof family members of Populus trichocarpa in this study and their members identified in 2017

Name Phytozome Member in 2017 Phytozome
PtDof01 Potri.001G086400 PtrDof36 POPTR_0001s11130
PtDof02 Potri.001G238400 PtrDof32 POPTR_0001s24540
PtDof03 Potri.002G070700 PtrDof16 POPTR_0002s07150
PtDof04 Potri.002G129600 PtrDof17 POPTR_0002s13100
PtDof05 Potri.002G174300 PtrDof33 POPTR_0002s17490
PtDof06 Potri.003G034200 PtrDof30 POPTR_0003s02890
PtDof07 Potri.003G144500 PtrDof31 POPTR_0003s14450
PtDof08 Potri.004G038800 PtrDof27 POPTR_0004s03900
PtDof09 Potri.004G046100 PtrDof2 POPTR_0004s04590
PtDof10 Potri.004G046600
PtDof11 Potri.004G056900 PtrDof1 POPTR_0004s05580
PtDof12 Potri.004G121800 PtrDof22 POPTR_0004s12120
PtDof13 Potri.005G131600 PtrDof34 POPTR_0005s13990
PtDof14 Potri.005G134200 PtrDof4 POPTR_0005s14080
PtDof15 Potri.005G149100 PtrDof5 POPTR_0005s19310
PtDof16 Potri.005G188900
PtDof17 Potri.006G084200 PtrDof35 POPTR_0006s08440
PtDof18 Potri.006G202500 PtrDof18 POPTR_0006s21700
PtDof19 Potri.007G036400 PtrDof23 POPTR_0007s11790
PtDof20 Potri.007G038100 PtrDof6 POPTR_0007s11620
PtDof21 Potri.007G058200 PtrDof26 POPTR_0007s09520
PtDof22 Potri.008G055100 PtrDof39 POPTR_0008s05520
PtDof23 Potri.008G087800 PtrDof37 POPTR_0008s08740
PtDof24 Potri.009G029500 PtrDof40 POPTR_0009s03490
PtDof25 Potri.010G167600 PtrDof38 POPTR_0010s17480
PtDof26 Potri.010G205400 PtrDof14 POPTR_0010s21240
PtDof27 Potri.011G047500 PtrDof28 POPTR_0011s04730
PtDof28 Potri.011G054300 PtrDof12 POPTR_0011s05410
PtDof29 Potri.011G054400
PtDof30 Potri.011G055600 PtrDof13 POPTR_0011s05450
PtDof31 Potri.011G065900 PtrDof7 POPTR_0011s07400
PtDof32 Potri.012G018700 PtrDof20 POPTR_0012s02570
PtDof33 Potri.012G063800 PtrDof19 POPTR_0012s12670
PtDof34 Potri.012G081300 PtrDof21 POPTR_0012s08280
PtDof35 Potri.013G066700 PtrDof24 POPTR_0013s06290
PtDof36 Potri.014G036600 PtrDof15 POPTR_0014s03590
PtDof37 Potri.014G100900 PtrDof29 POPTR_0014s09640
PtDof38 Potri.015G009300 PtrDof41 POPTR_0015s01160
PtDof39 Potri.015G048300 PtrDof9 POPTR_0015s03520
PtDof40 Potri.015G077100 PtrDof10 POPTR_0015s08810
PtDof41 Potri.016G069300 PtrDof11 POPTR_0016s07000
PtDof42 Potri.017G084600 PtrDof25 POPTR_0017s12080
PtDof43 Potri.019G040700 PtrDof8 POPTR_0019s05720
PtDof44 Potri.T146000 PtrDof3 POPTR_0005s21130

References

[1] Kraiser T, Gras DE, Gutiérrez AG, González B, Gutiérrez RA. A holistic view of nitrogen acquisition in plants. J Exp Botany. 2011;62:1455–66.10.1093/jxb/erq425Search in Google Scholar

[2] Lanquar V, Frommer WB. Adjusting ammonium uptake via phosphorylation. Plant Signal & Behav. 2010;5:736–8.10.4161/psb.5.6.11696Search in Google Scholar

[3] Dickson RE. Carbon and nitrogen allocation in trees. Annales Des Sci Forestières. 1989;46:347–50.10.1051/forest:198905ART0142Search in Google Scholar

[4] Britto DT, Siddiqi MY, Glass AD, Kronzucker HJ. Futile transmembrane NH4(+) cycling: a cellular hypothesis to explain ammonium toxicity in plants. Proc Natl Acad Sci U S Am. 2001;98:4255–8.10.1073/pnas.061034698Search in Google Scholar

[5] Robertson GP, Vitousek PM. Nitrogen in agriculture: balancing the cost of an essential resource. Soc Sci Electron Publ. 2009;34:97–125.10.1146/annurev.environ.032108.105046Search in Google Scholar

[6] Tuomi J, Niemelä P, Sirén S. The panglossian paradigm and delayed inducible accumulation of foliar phenolics in mountain birch. Oikos. 1990;59:399–410.10.2307/3545152Search in Google Scholar

[7] Shimofurutani N, Kisu Y, Suzuki M, Esaka M. Functional analyses of the Dof domain, a zinc finger DNA-binding domain, in a pumpkin DNA-binding protein AOBP. Febs Lett. 1998;430:251–6.10.1016/S0014-5793(98)00670-XSearch in Google Scholar

[8] Kisu Y, Ono T, Shimofurutani N, Suzuki M, Esaka M. Characterization and expression of a new class of zinc finger protein that binds to silencer region of ascorbate oxidase gene. Plant Cell Physiol. 1998;39:1054–64.10.1093/oxfordjournals.pcp.a029302Search in Google Scholar

[9] Yanagisawa S, Schmidt RJ. Diversity and similarity among recognition sequences of Dof transcription factors. Plant J. 1999;17:209–14.10.1046/j.1365-313X.1999.00363.xSearch in Google Scholar

[10] Yanagisawa S. The Dof family of plant transcription factors. Trends plant Sci. 2002;7:555–60.10.1016/S1360-1385(02)02362-2Search in Google Scholar

[11] Moreno-Risueno MÁ, Martínez M, Vicente-Carbajosa J, Carbonero P. The family of DOF transcription factors: from green unicellular algae to vascular plants. Mol Genet Genomics. 2007;277:379–90.10.1007/s00438-006-0186-9Search in Google Scholar PubMed

[12] Chen Y, Cao J. Comparative analysis of Dof transcription factor family in maize. Plant Mol Biol Report. 2015;33:1245–58.10.1007/s11105-014-0835-9Search in Google Scholar

[13] Lijavetzky D, Carbonero P, Vicente-Carbajosa J. Genome-Wide comparative phylogenetic analysis of the rice and Arabidopsis Dof gene families. BMC Evolut Biol. 2003;3:17.10.1186/1471-2148-3-17Search in Google Scholar PubMed PubMed Central

[14] Wang HW, Zhang B, Hao YJ, Huang J, Tian AG, Liao Y, et al. The soybean Dof‐type transcription factor genes, GmDof4 and GmDof11, enhance lipid content in the seeds of transgenic Arabidopsis plants. Plant J. 2007;52:716–29.10.1111/j.1365-313X.2007.03268.xSearch in Google Scholar PubMed

[15] Xu ZS, Tan HW, Wang F, Hou XL, Xiong AS. CarrotDB: a genomic and transcriptomic database for carrot. Database. 2014;2014:bau096.10.1093/database/bau096Search in Google Scholar PubMed PubMed Central

[16] Malviya N, Gupta S, Singh V, Yadav M, Bisht N, Sarangi B, et al. Genome wide in silico characterization of Dof gene families of pigeonpea (Cajanus cajan (L) Millsp.). Mol Biol Rep. 2015;42:535–52.10.1007/s11033-014-3797-ySearch in Google Scholar PubMed

[17] Shu Y, Song L, Zhang J, Liu Y, Guo C. Genome-wide identification and characterization of the Dof gene family in medicago truncatula. Genet Mol Res. 2015;14:10645–57.10.4238/2015.September.9.5Search in Google Scholar PubMed

[18] Wei Q, Wang W, Hu T, Hu H, Mao W, Zhu Q, et al. Genome-wide identification and characterization of Dof transcription factors in eggplant (Solanum melongena L.). Peerj. 2018;6:e4481.10.7717/peerj.4481Search in Google Scholar PubMed PubMed Central

[19] Wei PC, Tan F, Gao XQ, Zhang XQ, Wang GQ, Xu H, et al. Overexpression of AtDof4.7, an Arabidopsis Dof family transcription factor, induces floral organ abscission deficiency in Arabidopsis. Plant Physiol. 2010;153:1031–45.10.1104/pp.110.153247Search in Google Scholar PubMed PubMed Central

[20] Cai X, Zhang Y, Zhang C, Zhang T, Hu T, Ye J, et al. Genome‐wide analysis of plant‐specific Dof transcription factor family in tomato. J Integr plant Biol. 2013;55:552–66.10.1111/jipb.12043Search in Google Scholar PubMed

[21] Feng BH, Han YC, Xiao YY, Kuang JF, Fan ZQ, Chen JY, et al. The banana fruit Dof transcription factor MaDof23 acts as a repressor and interacts with MaERF9 in regulating ripening-related genes. J Exp Botany. 2016;67:2263–75.10.1093/jxb/erw032Search in Google Scholar PubMed PubMed Central

[22] Qi X, Li S, Zhu Y, Zhao Q, Zhu D, Yu J. ZmDof3, a maize endosperm-specific Dof protein gene, regulates starch accumulation and aleurone development in maize endosperm. Plant Mol Biol. 2017;93:1–14.10.1007/s11103-016-0543-ySearch in Google Scholar PubMed

[23] Wu Q, Liu X, Yin D, Yuan H, Xie Q, Zhao X, et al. Constitutive expression of OsDof4, encoding a C2-C2 zinc finger transcription factor, confesses its distinct flowering effects under long- and short-day photoperiods in rice (Oryza sativa L.). Bmc Plant Biol. 2017;17:166.10.1186/s12870-017-1109-0Search in Google Scholar PubMed PubMed Central

[24] Yang CQ, Fang X, Wu XM, Mao YB, Wang LJ, Chen XY. Transcriptional regulation of plant secondary metabolism. J Integr plant Biol. 2012;54:703–12.10.1111/j.1744-7909.2012.01161.xSearch in Google Scholar PubMed

[25] Skirycz A, Radziejwoski A, Busch W, Hannah MA, Czeszejko J, Kwaśniewski M, et al. The Dof transcription factor OBP1 is involved in cell cycle regulation in Arabidopsis thaliana. Plant J. 2008;56:779–92.10.1111/j.1365-313X.2008.03641.xSearch in Google Scholar PubMed

[26] Corrales AR, Carrillo L, Lasierra P, Nebauer SG, Dominguez-Figueroa J, Renau-Morata B, et al. Multifaceted role of cycling Dof factor 3 (CDF3) in the regulation of flowering time and abiotic stress responses in Arabidopsis. Plant Cell Environ. 2017;40:748–64.10.1111/pce.12894Search in Google Scholar PubMed

[27] Yang G, Yu L, Wang Y, Wang C, Gao C. The translation initiation factor 1A (TheIF1A) from Tamarix hispida is regulated by a Dof transcription factor and increased abiotic stress tolerance. Front Plant Sci. 2017;8:513.10.3389/fpls.2017.00513Search in Google Scholar PubMed PubMed Central

[28] Kumar R, Taware R, Gaur VS, Guru S, Kumar A. Influence of nitrogen on the expression of TaDof1 transcription factor in wheat and its relationship with photosynthetic and ammonium assimilating efficiency. Mol Biol Rep. 2009;36:2209–20.10.1007/s11033-008-9436-8Search in Google Scholar PubMed

[29] Sugiyama T, Ishida T, Tabei N, Shigyo M, Konishi M, Yoneyama T, et al. Involvement of PpDof1 transcriptional repressor in the nutrient condition-dependent growth control of protonemal filaments in physcomitrella patens. J Exp Bot. 2012;63:3185–97.10.1093/jxb/ers042Search in Google Scholar PubMed PubMed Central

[30] Yanagisawa S. Dof1 and Dof2 transcription factors are associated with expression of multiple genes involved in carbon metabolism in maize. Plant J. 2000;21:281–8.10.1046/j.1365-313x.2000.00685.xSearch in Google Scholar PubMed

[31] Yanagisawa S, Akiyama A, Kisaka H, Uchimiya H, Miwa T. Metabolic engineering with Dof1 transcription factor in plants: improved nitrogen assimilation and growth under low-nitrogen conditions. Proc Natl Acad Sci U S Am. 2004;101:7833–8.10.1073/pnas.0402267101Search in Google Scholar PubMed PubMed Central

[32] Kurai T, Wakayama M, Abiko T, Yanagisawa S, Aoki N, Ohsugi R. Introduction of the ZmDof1 gene into rice enhances carbon and nitrogen assimilation under low-nitrogen conditions. Plant Biotechnol J. 2011;9:826–37.10.1111/j.1467-7652.2011.00592.xSearch in Google Scholar PubMed

[33] Rueda-López M, Crespillo R, Cánovas FM, Ávila C. Differential regulation of two glutamine synthetase genes by a single Dof transcription factor. Plant J. 2008;56:73–85.10.1111/j.1365-313X.2008.03573.xSearch in Google Scholar PubMed

[34] Rueda-López M, Cañas RA, Canales J, Cánovas FM, Ávila C. The over-expression of the pine transcription factor PpDod 5 in Arabidopsis leads to increased lignin content and affects carbon and nitrogen metabolism. Physiolog Plant. 2015;155:369–83.10.1111/ppl.12381Search in Google Scholar PubMed

[35] Santos LA, Souza SRD, Fernandes MS. OsDof25 expression alters carbon and nitrogen metabolism in Arabidopsis under high-N supply. Plant Biotechnol Rep. 2012;6:327–37.10.1007/s11816-012-0227-2Search in Google Scholar

[36] Wang Y, Fu B, Pan L, Chen L, Fu X, Li K. Overexpression of Arabidopsis DofF1, GS1 and GS2 enhanced nitrogen assimilation in transgenic tobacco grown under low-nitrogen conditions. Plant Mol Biol Report. 2013;31:886–900.10.1007/s11105-013-0561-8Search in Google Scholar

[37] Lin W, Hagen E, Fulcher A, Hren MT, Cheng ZM. Overexpressing the ZmDOF1 gene in Populus does not improve growth and nitrogen assimilation under low-nitrogen conditions. Plant Cell Tissue Organ Cult (PCTOC). 2013;113:51–61.10.1007/s11240-012-0250-6Search in Google Scholar

[38] Peipei W, Jing L, Xiaoyang G. Genome-wide screening and characterization of the Dof gene family in physic nut (Jatropha curcas L.). Int J Mol Sci. 2018;19(6):1598.10.3390/ijms19061598Search in Google Scholar PubMed PubMed Central

[39] Azam SM, Liu Y, Rahman ZU. Identification, characterization and expression profiles of Dof transcription factors in pineapple (Ananas comosus L). Tropical Plant Biol. 2018;11:49–64.10.1007/s12042-018-9200-8Search in Google Scholar

[40] Liu X, Liu Z, Hao Z. Characterization of Dof family in Pyrus bretschneideri and role of PbDof9.2 in flowering time regulation. Genomics. 2020;112(1):712–20.10.1016/j.ygeno.2019.05.005Search in Google Scholar PubMed

[41] Tokunaga S, Sanda S, Uraguchi Y. Overexpression of the Dof-Type transcription factor enhances lipid synthesis in Chlorella vulgaris. Appl Biochem Biotechnol. 2019;189:116–28.10.1007/s12010-019-02990-7Search in Google Scholar PubMed

[42] Chattha WS, Atif RM, Iqbal M. Genome-wide identification and evolution of Dof transcription factor family in cultivated and ancestral cotton species. Genomics. 2020;112(6):4155–70.10.1016/j.ygeno.2020.07.006Search in Google Scholar PubMed

[43] Lohani N, Babaei S, Singh MB, Bhalla PL. Genome-wide in silico identification and comparative analysis of Dof gene family in Brassica napus. Plants. 2021;10(4):709.10.3390/plants10040709Search in Google Scholar PubMed PubMed Central

[44] Goodstein DM, Shu S, Howson R, Neupane R, Hayes RD, Fazo J, et al. Phytozome: a comparative platform for green plant genomics. Nucleic Acids Res. 2012;40:1178–86.10.1093/nar/gkr944Search in Google Scholar PubMed PubMed Central

[45] Zhang H, Jin J, Tang L, Zhao Y, Gu X, Gao G, et al. PlantTFDB 2.0: update and improvement of the comprehensive plant transcription factor database. Nucleic Acids Res. 2011;39:D1114–7.10.1093/nar/gkq1141Search in Google Scholar PubMed PubMed Central

[46] Wilkins MR, Gasteiger E, Bairoch A, Sanchez JC, Williams KL, Appel RD, et al. Protein identification and analysis tools in the ExPASy server. Methods Mol Biol. 1999;112:531–52.10.1385/1-59259-584-7:531Search in Google Scholar

[47] Hu B, Jin J, Guo AY, Zhang H, Luo J, Gao G. GSDS 2.0: an upgraded gene feature visualization server. Bioinformatics. 2015;31:1296–7.10.1093/bioinformatics/btu817Search in Google Scholar PubMed PubMed Central

[48] Tuskan GA, Difazio S, Jansson S, Bohlmann J, Grigoriev I, Hellsten U, et al. The genome of black cottonwood, Populus trichocarpa (Torr. & Gray). Science. 2006;313:1596–604.10.1126/science.1128691Search in Google Scholar PubMed

[49] Larkin MA, Blackshields G, Brown NP, Chenna RM, Mcgettigan PA, Mcwilliam H, et al. ClustalW and ClustalX version 2.0. Bioinformatics. 2007;23:2947–8.10.1093/bioinformatics/btm404Search in Google Scholar PubMed

[50] Tamura K, Peterson D, Peterson N, Stecher G, Nei M, Kumar S. MEGA5: molecular evolutionary genetics analysis using maximum likelihood, evolutionary distance, and maximum parsimony methods. Mol Biol Evol. 2011;28:2731–9.10.1093/molbev/msr121Search in Google Scholar PubMed PubMed Central

[51] Stern A, Doron-Faigenboim A, Erez E, Martz E, Bacharach E, Pupko T. Selecton 2007: advanced models for detecting positive and purifying selection using a Bayesian inference approach. Nucleic Acids Res. 2007;35:W506–11.10.1093/nar/gkm382Search in Google Scholar PubMed PubMed Central

[52] Doronfaigenboim A, Pupko T. A combined empirical and mechanistic codon model. Mol Biol Evol. 2007;24:388–97.10.1093/molbev/msl175Search in Google Scholar

[53] Clough SJ, Bent AF. Floral dip: a simplified method for Agrobacterium-mediated transformation of Arabidopsis thaliana. Plant J. 1998;16:735–43.10.1046/j.1365-313x.1998.00343.xSearch in Google Scholar

[54] Li D, Yang C, Li X, Gan Q, Zhao X, Zhu L. Functional characterization of rice OsDof12. Planta. 2009;229:1159–69.10.1007/s00425-009-0893-7Search in Google Scholar

[55] Sasaki N, Matsumaru M, Odaira S, Nakata A, Nakata K, Nakayama I, et al. Transient expression of tobacco BBF1-related Dof proteins, BBF2 and BBF3, upregulates genes involved in virus resistance and pathogen defense. Physiolog Mol Plant Pathol. 2015;89:70–7.10.1016/j.pmpp.2014.12.005Search in Google Scholar

[56] Yanagisawa S, Izui K. Molecular cloning of two DNA-binding proteins of maize that are structurally different but interact with the same sequence motif. J Biol Chem. 1993;268:16028–36.10.1016/S0021-9258(18)82353-5Search in Google Scholar

[57] Yang X, Tuskan GA. Divergence of the Dof gene families in poplar, Arabidopsis, and rice suggests multiple modes of gene evolution after duplication. Plant Physiol. 2006;142:820–30.10.1104/pp.106.083642Search in Google Scholar PubMed PubMed Central

[58] Wang H, Zhao S, Gao Y, Yang J. Characterization of Dof transcription factors and their responses to osmotic stress in poplar (Populus trichocarpa). PLoS One. 2017;12:e0170210.10.1371/journal.pone.0170210Search in Google Scholar PubMed PubMed Central

[59] Chollet R, Vidal J, O’Leary MH. PHOSPHOENOLPYRUVATE CARBOXYLASE: A Ubiquitous, highly regulated enzyme in plants. Annurevplant Physiolplant Molbiol. 1996;47:273–98.10.1146/annurev.arplant.47.1.273Search in Google Scholar PubMed

Received: 2021-11-19
Revised: 2022-04-12
Accepted: 2022-04-15
Published Online: 2022-07-13

© 2022 Shenmeng Wang et al., published by De Gruyter

This work is licensed under the Creative Commons Attribution 4.0 International License.

Articles in the same Issue

  1. Biomedical Sciences
  2. Effects of direct oral anticoagulants dabigatran and rivaroxaban on the blood coagulation function in rabbits
  3. The mother of all battles: Viruses vs humans. Can humans avoid extinction in 50–100 years?
  4. Knockdown of G1P3 inhibits cell proliferation and enhances the cytotoxicity of dexamethasone in acute lymphoblastic leukemia
  5. LINC00665 regulates hepatocellular carcinoma by modulating mRNA via the m6A enzyme
  6. Association study of CLDN14 variations in patients with kidney stones
  7. Concanavalin A-induced autoimmune hepatitis model in mice: Mechanisms and future outlook
  8. Regulation of miR-30b in cancer development, apoptosis, and drug resistance
  9. Informatic analysis of the pulmonary microecology in non-cystic fibrosis bronchiectasis at three different stages
  10. Swimming attenuates tumor growth in CT-26 tumor-bearing mice and suppresses angiogenesis by mediating the HIF-1α/VEGFA pathway
  11. Characterization of intestinal microbiota and serum metabolites in patients with mild hepatic encephalopathy
  12. Functional conservation and divergence in plant-specific GRF gene family revealed by sequences and expression analysis
  13. Application of the FLP/LoxP-FRT recombination system to switch the eGFP expression in a model prokaryote
  14. Biomedical evaluation of antioxidant properties of lamb meat enriched with iodine and selenium
  15. Intravenous infusion of the exosomes derived from human umbilical cord mesenchymal stem cells enhance neurological recovery after traumatic brain injury via suppressing the NF-κB pathway
  16. Effect of dietary pattern on pregnant women with gestational diabetes mellitus and its clinical significance
  17. Potential regulatory mechanism of TNF-α/TNFR1/ANXA1 in glioma cells and its role in glioma cell proliferation
  18. Effect of the genetic mutant G71R in uridine diphosphate-glucuronosyltransferase 1A1 on the conjugation of bilirubin
  19. Quercetin inhibits cytotoxicity of PC12 cells induced by amyloid-beta 25–35 via stimulating estrogen receptor α, activating ERK1/2, and inhibiting apoptosis
  20. Nutrition intervention in the management of novel coronavirus pneumonia patients
  21. circ-CFH promotes the development of HCC by regulating cell proliferation, apoptosis, migration, invasion, and glycolysis through the miR-377-3p/RNF38 axis
  22. Bmi-1 directly upregulates glucose transporter 1 in human gastric adenocarcinoma
  23. Lacunar infarction aggravates the cognitive deficit in the elderly with white matter lesion
  24. Hydroxysafflor yellow A improved retinopathy via Nrf2/HO-1 pathway in rats
  25. Comparison of axon extension: PTFE versus PLA formed by a 3D printer
  26. Elevated IL-35 level and iTr35 subset increase the bacterial burden and lung lesions in Mycobacterium tuberculosis-infected mice
  27. A case report of CAT gene and HNF1β gene variations in a patient with early-onset diabetes
  28. Study on the mechanism of inhibiting patulin production by fengycin
  29. SOX4 promotes high-glucose-induced inflammation and angiogenesis of retinal endothelial cells by activating NF-κB signaling pathway
  30. Relationship between blood clots and COVID-19 vaccines: A literature review
  31. Analysis of genetic characteristics of 436 children with dysplasia and detailed analysis of rare karyotype
  32. Bioinformatics network analyses of growth differentiation factor 11
  33. NR4A1 inhibits the epithelial–mesenchymal transition of hepatic stellate cells: Involvement of TGF-β–Smad2/3/4–ZEB signaling
  34. Expression of Zeb1 in the differentiation of mouse embryonic stem cell
  35. Study on the genetic damage caused by cadmium sulfide quantum dots in human lymphocytes
  36. Association between single-nucleotide polymorphisms of NKX2.5 and congenital heart disease in Chinese population: A meta-analysis
  37. Assessment of the anesthetic effect of modified pentothal sodium solution on Sprague-Dawley rats
  38. Genetic susceptibility to high myopia in Han Chinese population
  39. Potential biomarkers and molecular mechanisms in preeclampsia progression
  40. Silencing circular RNA-friend leukemia virus integration 1 restrained malignancy of CC cells and oxaliplatin resistance by disturbing dyskeratosis congenita 1
  41. Endostar plus pembrolizumab combined with a platinum-based dual chemotherapy regime for advanced pulmonary large-cell neuroendocrine carcinoma as a first-line treatment: A case report
  42. The significance of PAK4 in signaling and clinicopathology: A review
  43. Sorafenib inhibits ovarian cancer cell proliferation and mobility and induces radiosensitivity by targeting the tumor cell epithelial–mesenchymal transition
  44. Characterization of rabbit polyclonal antibody against camel recombinant nanobodies
  45. Active legumain promotes invasion and migration of neuroblastoma by regulating epithelial-mesenchymal transition
  46. Effect of cell receptors in the pathogenesis of osteoarthritis: Current insights
  47. MT-12 inhibits the proliferation of bladder cells in vitro and in vivo by enhancing autophagy through mitochondrial dysfunction
  48. Study of hsa_circRNA_000121 and hsa_circRNA_004183 in papillary thyroid microcarcinoma
  49. BuyangHuanwu Decoction attenuates cerebral vasospasm caused by subarachnoid hemorrhage in rats via PI3K/AKT/eNOS axis
  50. Effects of the interaction of Notch and TLR4 pathways on inflammation and heart function in septic heart
  51. Monosodium iodoacetate-induced subchondral bone microstructure and inflammatory changes in an animal model of osteoarthritis
  52. A rare presentation of type II Abernethy malformation and nephrotic syndrome: Case report and review
  53. Rapid death due to pulmonary epithelioid haemangioendothelioma in several weeks: A case report
  54. Hepatoprotective role of peroxisome proliferator-activated receptor-α in non-cancerous hepatic tissues following transcatheter arterial embolization
  55. Correlation between peripheral blood lymphocyte subpopulations and primary systemic lupus erythematosus
  56. A novel SLC8A1-ALK fusion in lung adenocarcinoma confers sensitivity to alectinib: A case report
  57. β-Hydroxybutyrate upregulates FGF21 expression through inhibition of histone deacetylases in hepatocytes
  58. Identification of metabolic genes for the prediction of prognosis and tumor microenvironment infiltration in early-stage non-small cell lung cancer
  59. BTBD10 inhibits glioma tumorigenesis by downregulating cyclin D1 and p-Akt
  60. Mucormycosis co-infection in COVID-19 patients: An update
  61. Metagenomic next-generation sequencing in diagnosing Pneumocystis jirovecii pneumonia: A case report
  62. Long non-coding RNA HOXB-AS1 is a prognostic marker and promotes hepatocellular carcinoma cells’ proliferation and invasion
  63. Preparation and evaluation of LA-PEG-SPION, a targeted MRI contrast agent for liver cancer
  64. Proteomic analysis of the liver regulating lipid metabolism in Chaohu ducks using two-dimensional electrophoresis
  65. Nasopharyngeal tuberculosis: A case report
  66. Characterization and evaluation of anti-Salmonella enteritidis activity of indigenous probiotic lactobacilli in mice
  67. Aberrant pulmonary immune response of obese mice to periodontal infection
  68. Bacteriospermia – A formidable player in male subfertility
  69. In silico and in vivo analysis of TIPE1 expression in diffuse large B cell lymphoma
  70. Effects of KCa channels on biological behavior of trophoblasts
  71. Interleukin-17A influences the vulnerability rather than the size of established atherosclerotic plaques in apolipoprotein E-deficient mice
  72. Multiple organ failure and death caused by Staphylococcus aureus hip infection: A case report
  73. Prognostic signature related to the immune environment of oral squamous cell carcinoma
  74. Primary and metastatic squamous cell carcinoma of the thyroid gland: Two case reports
  75. Neuroprotective effects of crocin and crocin-loaded niosomes against the paraquat-induced oxidative brain damage in rats
  76. Role of MMP-2 and CD147 in kidney fibrosis
  77. Geometric basis of action potential of skeletal muscle cells and neurons
  78. Babesia microti-induced fulminant sepsis in an immunocompromised host: A case report and the case-specific literature review
  79. Role of cerebellar cortex in associative learning and memory in guinea pigs
  80. Application of metagenomic next-generation sequencing technique for diagnosing a specific case of necrotizing meningoencephalitis caused by human herpesvirus 2
  81. Case report: Quadruple primary malignant neoplasms including esophageal, ureteral, and lung in an elderly male
  82. Long non-coding RNA NEAT1 promotes angiogenesis in hepatoma carcinoma via the miR-125a-5p/VEGF pathway
  83. Osteogenic differentiation of periodontal membrane stem cells in inflammatory environments
  84. Knockdown of SHMT2 enhances the sensitivity of gastric cancer cells to radiotherapy through the Wnt/β-catenin pathway
  85. Continuous renal replacement therapy combined with double filtration plasmapheresis in the treatment of severe lupus complicated by serious bacterial infections in children: A case report
  86. Simultaneous triple primary malignancies, including bladder cancer, lymphoma, and lung cancer, in an elderly male: A case report
  87. Preclinical immunogenicity assessment of a cell-based inactivated whole-virion H5N1 influenza vaccine
  88. One case of iodine-125 therapy – A new minimally invasive treatment of intrahepatic cholangiocarcinoma
  89. S1P promotes corneal trigeminal neuron differentiation and corneal nerve repair via upregulating nerve growth factor expression in a mouse model
  90. Early cancer detection by a targeted methylation assay of circulating tumor DNA in plasma
  91. Calcifying nanoparticles initiate the calcification process of mesenchymal stem cells in vitro through the activation of the TGF-β1/Smad signaling pathway and promote the decay of echinococcosis
  92. Evaluation of prognostic markers in patients infected with SARS-CoV-2
  93. N6-Methyladenosine-related alternative splicing events play a role in bladder cancer
  94. Characterization of the structural, oxidative, and immunological features of testis tissue from Zucker diabetic fatty rats
  95. Effects of glucose and osmotic pressure on the proliferation and cell cycle of human chorionic trophoblast cells
  96. Investigation of genotype diversity of 7,804 norovirus sequences in humans and animals of China
  97. Characteristics and karyotype analysis of a patient with turner syndrome complicated with multiple-site tumors: A case report
  98. Aggravated renal fibrosis is positively associated with the activation of HMGB1-TLR2/4 signaling in STZ-induced diabetic mice
  99. Distribution characteristics of SARS-CoV-2 IgM/IgG in false-positive results detected by chemiluminescent immunoassay
  100. SRPX2 attenuated oxygen–glucose deprivation and reperfusion-induced injury in cardiomyocytes via alleviating endoplasmic reticulum stress-induced apoptosis through targeting PI3K/Akt/mTOR axis
  101. Aquaporin-8 overexpression is involved in vascular structure and function changes in placentas of gestational diabetes mellitus patients
  102. Relationship between CRP gene polymorphisms and ischemic stroke risk: A systematic review and meta-analysis
  103. Effects of growth hormone on lipid metabolism and sexual development in pubertal obese male rats
  104. Cloning and identification of the CTLA-4IgV gene and functional application of vaccine in Xinjiang sheep
  105. Antitumor activity of RUNX3: Upregulation of E-cadherin and downregulation of the epithelial–mesenchymal transition in clear-cell renal cell carcinoma
  106. PHF8 promotes osteogenic differentiation of BMSCs in old rat with osteoporosis by regulating Wnt/β-catenin pathway
  107. A review of the current state of the computer-aided diagnosis (CAD) systems for breast cancer diagnosis
  108. Bilateral dacryoadenitis in adult-onset Still’s disease: A case report
  109. A novel association between Bmi-1 protein expression and the SUVmax obtained by 18F-FDG PET/CT in patients with gastric adenocarcinoma
  110. The role of erythrocytes and erythroid progenitor cells in tumors
  111. Relationship between platelet activation markers and spontaneous abortion: A meta-analysis
  112. Abnormal methylation caused by folic acid deficiency in neural tube defects
  113. Silencing TLR4 using an ultrasound-targeted microbubble destruction-based shRNA system reduces ischemia-induced seizures in hyperglycemic rats
  114. Plant Sciences
  115. Seasonal succession of bacterial communities in cultured Caulerpa lentillifera detected by high-throughput sequencing
  116. Cloning and prokaryotic expression of WRKY48 from Caragana intermedia
  117. Novel Brassica hybrids with different resistance to Leptosphaeria maculans reveal unbalanced rDNA signal patterns
  118. Application of exogenous auxin and gibberellin regulates the bolting of lettuce (Lactuca sativa L.)
  119. Phytoremediation of pollutants from wastewater: A concise review
  120. Genome-wide identification and characterization of NBS-encoding genes in the sweet potato wild ancestor Ipomoea trifida (H.B.K.)
  121. Alleviative effects of magnetic Fe3O4 nanoparticles on the physiological toxicity of 3-nitrophenol to rice (Oryza sativa L.) seedlings
  122. Selection and functional identification of Dof genes expressed in response to nitrogen in Populus simonii × Populus nigra
  123. Study on pecan seed germination influenced by seed endocarp
  124. Identification of active compounds in Ophiopogonis Radix from different geographical origins by UPLC-Q/TOF-MS combined with GC-MS approaches
  125. The entire chloroplast genome sequence of Asparagus cochinchinensis and genetic comparison to Asparagus species
  126. Genome-wide identification of MAPK family genes and their response to abiotic stresses in tea plant (Camellia sinensis)
  127. Selection and validation of reference genes for RT-qPCR analysis of different organs at various development stages in Caragana intermedia
  128. Cloning and expression analysis of SERK1 gene in Diospyros lotus
  129. Integrated metabolomic and transcriptomic profiling revealed coping mechanisms of the edible and medicinal homologous plant Plantago asiatica L. cadmium resistance
  130. A missense variant in NCF1 is associated with susceptibility to unexplained recurrent spontaneous abortion
  131. Assessment of drought tolerance indices in faba bean genotypes under different irrigation regimes
  132. The entire chloroplast genome sequence of Asparagus setaceus (Kunth) Jessop: Genome structure, gene composition, and phylogenetic analysis in Asparagaceae
  133. Food Science
  134. Dietary food additive monosodium glutamate with or without high-lipid diet induces spleen anomaly: A mechanistic approach on rat model
  135. Binge eating disorder during COVID-19
  136. Potential of honey against the onset of autoimmune diabetes and its associated nephropathy, pancreatitis, and retinopathy in type 1 diabetic animal model
  137. FTO gene expression in diet-induced obesity is downregulated by Solanum fruit supplementation
  138. Physical activity enhances fecal lactobacilli in rats chronically drinking sweetened cola beverage
  139. Supercritical CO2 extraction, chemical composition, and antioxidant effects of Coreopsis tinctoria Nutt. oleoresin
  140. Functional constituents of plant-based foods boost immunity against acute and chronic disorders
  141. Effect of selenium and methods of protein extraction on the proteomic profile of Saccharomyces yeast
  142. Microbial diversity of milk ghee in southern Gansu and its effect on the formation of ghee flavor compounds
  143. Ecology and Environmental Sciences
  144. Effects of heavy metals on bacterial community surrounding Bijiashan mining area located in northwest China
  145. Microorganism community composition analysis coupling with 15N tracer experiments reveals the nitrification rate and N2O emissions in low pH soils in Southern China
  146. Genetic diversity and population structure of Cinnamomum balansae Lecomte inferred by microsatellites
  147. Preliminary screening of microplastic contamination in different marine fish species of Taif market, Saudi Arabia
  148. Plant volatile organic compounds attractive to Lygus pratensis
  149. Effects of organic materials on soil bacterial community structure in long-term continuous cropping of tomato in greenhouse
  150. Effects of soil treated fungicide fluopimomide on tomato (Solanum lycopersicum L.) disease control and plant growth
  151. Prevalence of Yersinia pestis among rodents captured in a semi-arid tropical ecosystem of south-western Zimbabwe
  152. Effects of irrigation and nitrogen fertilization on mitigating salt-induced Na+ toxicity and sustaining sea rice growth
  153. Bioengineering and Biotechnology
  154. Poly-l-lysine-caused cell adhesion induces pyroptosis in THP-1 monocytes
  155. Development of alkaline phosphatase-scFv and its use for one-step enzyme-linked immunosorbent assay for His-tagged protein detection
  156. Development and validation of a predictive model for immune-related genes in patients with tongue squamous cell carcinoma
  157. Agriculture
  158. Effects of chemical-based fertilizer replacement with biochar-based fertilizer on albic soil nutrient content and maize yield
  159. Genome-wide identification and expression analysis of CPP-like gene family in Triticum aestivum L. under different hormone and stress conditions
  160. Agronomic and economic performance of mung bean (Vigna radiata L.) varieties in response to rates of blended NPS fertilizer in Kindo Koysha district, Southern Ethiopia
  161. Influence of furrow irrigation regime on the yield and water consumption indicators of winter wheat based on a multi-level fuzzy comprehensive evaluation
  162. Discovery of exercise-related genes and pathway analysis based on comparative genomes of Mongolian originated Abaga and Wushen horse
  163. Lessons from integrated seasonal forecast-crop modelling in Africa: A systematic review
  164. Evolution trend of soil fertility in tobacco-planting area of Chenzhou, Hunan Province, China
  165. Animal Sciences
  166. Morphological and molecular characterization of Tatera indica Hardwicke 1807 (Rodentia: Muridae) from Pothwar, Pakistan
  167. Research on meat quality of Qianhua Mutton Merino sheep and Small-tail Han sheep
  168. SI: A Scientific Memoir
  169. Suggestions on leading an academic research laboratory group
  170. My scientific genealogy and the Toronto ACDC Laboratory, 1988–2022
  171. Erratum
  172. Erratum to “Changes of immune cells in patients with hepatocellular carcinoma treated by radiofrequency ablation and hepatectomy, a pilot study”
  173. Erratum to “A two-microRNA signature predicts the progression of male thyroid cancer”
  174. Retraction
  175. Retraction of “Lidocaine has antitumor effect on hepatocellular carcinoma via the circ_DYNC1H1/miR-520a-3p/USP14 axis”
Downloaded on 8.12.2025 from https://www.degruyterbrill.com/document/doi/10.1515/biol-2022-0084/html
Scroll to top button