Introduction

The development of de novo roots, or new roots (Davies et al. 2018), on above-ground plant parts like stems is referred to as adventitious rooting (AR) or adventitious root formation (Roussos 2023). This process can occur naturally, sometimes serving as a survival response to abiotic stresses like flooding or salt exposure (Roussos 2023) or it can be induced artificially (Janick 1986) as a tool for producing new plants, including those grown in the horticulture industry. New plants regenerated or reproduced by nonsexual means, derived from plant tissues or plant parts, and not involving sexual recombination, have been propagated vegetatively (Janick 1986). With few exceptions, an explant generated using vegetative or clonal propagation techniques will maintain specific, desired characteristics and remain true-to-type (Dirr and Heuser 2006). The horticulture industry has thereby employed this phenomenon to overcome sexual reproduction barriers and commercialize the production of many desirable horticultural commodities and specialties ranging from ornamental crops to fruit, nuts, and vegetable crops (Davies et al. 1994). The ease of evoking AR on horticultural commodities can vary by species (Stokes et al. 2023) and for this reason, improving AR of desirable clones of plants has become a central theme of horticultural crop development and advancement (Davies et al. 2018), as well as profitability in the horticulture industry (Konjoian 2017). Advancing technologies or techniques that improve AR can yield increase in the diversity of plant selections cultivated, which subsequently become available for application by humans (Preece 2003). In a modern horticultural context, this trend can equate to economic opportunities for producers and broader crop and plant selection availability for consumers. There are a variety of factors that play a role in the success of AR formation on plants; however, this review explores the potential for using new opportunities for chemical biology to improve the AR of plants valuable for horticultural applications.

Early history

AR of fruit trees has a long history in the evolution of human behavior because under domestication the maintenance of desired genotypes becomes practical only by vegetative propagation (Zohary and Spiegel-Roy 1975; Weiss 2015). In fact, in very ancient writings, such as the Tanakh (Tanakh 1985), there are notations about “degenerate plants” that allude to the inherent problems of seed-propagated fruit crops. Propagation by cuttings was discussed in Book 2 of Historia Plantarum (Theophrastus; Einarson 1976) as well as in Natural History (Pliny the Elder; Anthony et al. 2010). The ability of people to select and maintain unique phenotypes likely advanced very early in human history and supported the change from a nomadic lifestyle to resident agriculture because fruit tree agriculture requires site-specific long-term residence. Fruit crops that can be easily propagated vegetatively could be considered preadapted for domestication (Zohary and Spiegel-Roy 1975; Zohary et al. 2012) and formed the basis for an early understanding of innate AR ability. However, it was a rich history that early agriculturists did not rely only on spontaneous rooting and treatments that accelerated root formation (Marston 1955). For example, Weaver (1972) discussed how, for centuries, farmers in Afghanistan and gardeners in Dutch used seeds of grains, such as barley, to induce rooting on cuttings. In addition, before the use of auxins specifically was established for rooting, treatments that likely resulted in a wound-auxin response (LaRue 1941; Xuan et al. 2008; Guo et al. 2008; Canher et al. 2020) were employed. These included manipulations such as simply cutting the stem itself or root removal (Steffens and Rasmussen 2016), as well as treatments with permanganate (Curtis 1918) or carbon monoxide (Zimmerman et al. 1933). In L.H. Bailey’s classic “The Nursery-Manual” (Bailey 1920), he notes three important conditions for successful “cuttage”: a moist and uniform atmosphere, porous soil, and sometimes bottom heat. Elevated root temperature was important in the 19th and early 20th century propagation greenhouse, and it could be due to an auxin regulatory mechanism, the increase in auxin levels with elevated temperature (Gray et al. 1998). Elevated temperature effects on auxin levels are mediated by phytochrome interacting factor 4 (PIF4; Franklin et al. 2011), thus it is connected to phytochrome regulated environmental treatments such as light quality and dark exposure periods (Halliday et al. 2009; Tillmann et al. 2022) that alter plant auxin responses. The early 20th century progress on AR was reviewed by Preece (2003).

Auxins and rooting

The first reports that auxins promoted root formation in cuttings were from Thimann and Went (1934), and Thimann and Koepfli (1935). Zimmerman and Wilcoxon (1935), and later Hitchcock and Zimmerman (1939) reported soon after about their studies on a variety of different auxin-like compounds, including indole-3-butyric acid (IBA) (see Table 1 for a list of compound abbreviations and acronyms). Zimmerman and Hitchcock trademarked IBA as Hormodin, and this started the first commercial use of IBA for rooting cuttings for propagation. In the subsequent decade, many compounds were tested in numerous trials for their ability to initiate AR, mainly with woody plant cuttings (Thimann and Behnke-Rogers 1950). Such early characterizations of auxins as “root-forming hormones” established a long-standing link between auxin and auxin-like compounds and root development (Went 1929; Thimann and Went 1934). Commercial plant propagation quickly recognized and adapted the technique of applying auxin for the rooting of stem cuttings of many forest and nursery crops. These techniques were applied to cuttings collected from taxa historically considered more challenging to root and greatly expanded the number of plants that could be propagated commercially. Evaluation of the merits of various auxin application methods continued through the second half of the 20th century (Huckenpahler 1955), with IBA as a basal quick-dip and powder application methods establishing themselves as the most broadly employed techniques in horticulture (Blythe et al. 2007; http://www.rooting-hormones.com/IBAmethd.htm; http://www.getroots.net/search.html; https://npn.rngr.net/npn/propagation).

Table 1 Compounds discussed in the text, including their proper names, the abbreviations and acronyms used in the text, and alternatives that are commonly used or are trade names. Each compound is also identified by either a primary literature reference or the Chemical Abstracts Service Registry Number (https://scifinder-n.cas.org/). The compounds are listed in order of their appearance in the text

When IBA was first found to have significant activity in the rooting process, it was widely assumed that it was a synthetic auxin, and in many growth bioassays, it was shown to be a weaker auxin than indole-3-acetic acid (IAA) (Woodward and Bartel 2005; Simon et al. 2013). A decade or more after its first uses for rooting were described, there appeared an early report of its presence in plants based on paper chromatography and bioassay in potato peelings (Blommaert 1954). It was also described as an endogenous auxin based on gas chromatography of extracts from Nicotiana (Bayer 1969). lBA was identified by gas chromatography-mass spectrometry (GC-MS) in pea root nodules (Badenoch-Jones et al. 1984) and reported in pea root and epicotyl (Schneider et al. 1985). Both free IBA and IBA released by hydrolysis from ester-conjugated lBA were clearly identified by GC-MS in the kernels and leaves of Zea mays (Epstein et al. 1989; Fallik et al. 1989; Ludwig-Müller and Epstein 1991) and subsequently in at least seven other plant species (Epstein and Ludwig-Müller 1993). Arabidopsis plants, for example, can accumulate a detectable amount of IBA (Epstein and Ludwig-Müller 1993). The endogenous presence of IBA is, however, somewhat inconsistent. For example, in maize, one variety exhibited detectable IBA and another did not (Epstein et al. 1989); in Arabidopsis, IBA has been found in both greater or lesser amounts, and in one study it was not detected at all (Novak et al. 2012). What controls the levels of endogenous IBA in specific varieties and growth conditions is not fully understood, although for Arabidopsis the pH of the growth medium, light intensity, and the volume of the culture flask can all have an effect (Ludwig-Müller et al. 1993). Similarly, the processes by which IBA is synthesized in plants need further clarification (Ludwig-Müller 2007; Damodaran and Strader 2019).

The biochemical conversion of IBA to IAA has been demonstrated in a variety of plants using isotope tracer methods (Epstein and Lavee 1984; Ludwig-Müller and Epstein 1991; Nordström et al. 1991; Van der Krieken et al. 1992; Baraldi et al. 1993, 1995; Kreiser et al. 2016). The isolation and characterization of mutants that are resistant to inhibitory concentrations of IBA or other long-side chain auxins but then respond normally to IAA or synthetic auxins has allowed the isolation of mutants defective in IBA responses, peroxisomal β-oxidation and peroxisome biogenesis (Zolman et al. 2000; Woodward and Bartel 2005; Baker et al. 2006; Damodaran and Strader 2019). The in vivo conversion of IBA to IAA involves six steps (Fig. 1), all essentially analogous to the β-oxidation of fatty acids: activation (thioesterification), oxidation, hydration, dehydration, thiolysis, and hydrolysis (Adham et al. 2005; Spiess and Zolman 2013; Rinaldi et al. 2016; Jawahir and Zolman 2021). IBA, or the synthetic auxin precursor 2,4-dichlorophenoxybutryric acid (2,4-DB) as well as chlorinated and dechlorinated IAAs, are converted to IAA or their respective analogues, indicating a somewhat permissive biochemical process. Following β-oxidation, the IAA- or auxin-like product that is generated seems to be exported from the peroxisomes. The genetic screens for Arabidopsis mutants resistant to exogenous IBA (initially named ibr, IBA resistant, or ped, peroxisome defective) has played an important role in our understanding of auxin metabolism and the special role IBA plays in this process. For example, the analysis of ech2 and other ibr mutants demonstrated that IBA-derived IAA plays an important role in root cell expansion (Strader et al. 2010b) as well as root hair and cotyledon cell expansion (Strader et al. 2010a, 2011). Although IBA application appears to have specificity for induction of root growth, it is nevertheless, only active upon its conversion to IAA (Strader et al. 2010b, 2011), indicating that it is an important auxin precursor rather than a weak auxin as originally proposed. This finding was confirmed when it was shown that the four-carbon side chain of IBA renders it unable to stimulate the formation of the TRANSPORT INHIBITOR RESPONSE 1/AUXIN SIGNALING F-BOX PROTEIN-Auxin/INDOLE-3-ACETIC ACID (TIR1/AFB-Aux/IAA) co-receptor complex required for auxin responsiveness (Uzunova et al. 2016). While studies have shown that IBA behaves differently than IAA and some have proposed it acts as a plant hormone itself (Van der Krieken et al. 1992, 1993; Chhun et al. 2004; Wang et al. 1994; Barnes 2011). Over the last two decades, the evolution of chemical biology has opened up novel high-throughput screening tools to explore auxin regulation of plant development (De Rybel et al. 2009), however, the use of these for the routine manipulation of plant materials has not replaced the regular use of IBA and other simple auxins. The goal of this review is to document the potential for both research and potential applications of the knowledge provided by these emerging discoveries. A summary of these chemical effectors and their target biochemical steps are shown in Fig. 3.

Fig. 3
figure 3

Hormonal pathways leading to root organogenesis required for adventitious root formation. Shown are the steps and pools of active compounds that are impacted by the synthetic compounds/agonists, inhibitors and activators discussed in the text. Compound abbreviations are consistent with what are defined in Table 1. IAA, indole-3-acetic acid; IBA, indole-3-butyric acid; NAA, 1-naphthaleneacetic acid; 4-CPA, 4-chlorophenoxyacetic acid; MCPA, 2-methyl-4-chlorophenoxyacetic acid; 2-DP, 2-(2,4-dichlorophenoxy) propionic acid; 4-Cl-IAA, 4-chloroindole-3-acetic acid; 5,6-diCl-IAA, 5,6-dichloroindole-3-acetic acid; 4-Cl-IBA, 4-chloroindole-3-butyric acid; 5,6-diCl-IBA, 5,6-dichloroindole-3-butyric acid; ICapA, indole-3-caproic acid; naxillin, (2E)-2-({5-[3-(trifluoromethyl) phenyl]-2-furyl} methylene) hydrazinecarbothioamide; 2,4DP-glyMe, 2-(2,4-dichlorophenoxy) propanoic acid-glycine methyl ester; 4-CPA-TrpMe, 4-chlorophenoxyacetic acid-L-tryptophan-O-methyl ester; JA, jasmonic acid; DHAP, 2,6-dihydroxyacetophenone; AIEP, adenosine-5’-[2-(1H-indol-3-yl) ethyl] phosphate; kakeimide, 4-(1,3-dioxoisoindolin-2-yl)-N-(3-isopropoxyphenyl) butanamide; nalacin, N-[4-[[6-(1H-pyrazol-1-yl)-3-pyridazinyl] amino] phenyl]-3-(trifluoromethyl)benzamide; IMT, indole-3-methyltetrazole; 4-Cl-IMT, 4-chloroindole-3-methyltetrazole; sortin2, 5-[[5-(3-chlorophenyl)-2-furanyl] methylene]-4-oxo-2-thioxo-3-thiazolidineethanesulfonic acid; retinal, (2E,4E,6E,8E)-3,7-dimethyl-9-(2,6,6-trimethylcyclohex-1-en-1-yl) nona-2,4,6,8-tetraenal; AVG, L-alpha-(2-aminoethoxyvinyl) glycine; AOA, 2-aminooxyacetic acid; rhizobitoxine, 2-amino-4-(2-amino-3-hydropropoxy)-trans-but-3-enoic acid; AIB, 2-aminoisobutyric acid; NBD, 2,5-norbornadiene; TCO, trans-cyclooctene, 1-MCP, 1-methylcyclopropene; triplin, 1-(1-morpholino-1-(thiophen-2-yl) propan-2-yl)-3-(2-(trifluoromethoxy) phenyl) thiourea; SHAM, salicylhydroxamic acid; DIECA, diethyldithiocarbamic acid; jarin-1, biphenyl-4-carboxylic acid [3-(3-methoxy-propionyl)-8-oxo-1,3,4,5,6,8-hexahydro-2H-1,5-methano-pyrido[1,2-a] [1,5] diazocin-9-yl]-amide; COMO, coronatine O-methyloxime; J4, 5-[3-(trifluoromethyl) benzylidene]-1,3-thiazolidine-2,4-dione; Y11, 2-ethoxy-4-(2-nitrovinyl) phenol; Y20, 4-hydroxy-3-[(4-methylcyclohexyl) carbonyl]-2H-chromen-2-one; lyn3, 3-[2-(Pyridin-4-yl) azepane-1-carbonyl]-1,2-dihydroisoquinolin-1-one; GA, gibberellins; H-acid, (1R,4R,5S,8S)-8-(hydroxymethyl)-1,7-dimethyl-4-propan-2-ylbicyclo [3.2.1] oct-6-ene-6-carboxylic acid; Compound 67D, (2(S)-3-phenyl-(9,10-dihydro-9,10-ethanoanthracene-11,12-dicarboximido) propanoic acid; Compound 6, (2(S)-3-methyl-(9,10-dihydro-9,10-ethanoanthracene-11,12-dicarboximido) penthanoic acid; AC94377, phthalimide 1-(3-chlorophthalimido)-cyclohexanecarboxamide; A1, N-(2-aminoethyl)-naphthalene-1-sulfonamide hydrochloride; TSPC, 3-(2-thienylsulfonyl) pyrazine-2-carbonitrile; SA, salicylic acid; chlormequat, 2-chloroethyl) trimethylammonium chloride; mepiquat, 1,1-dimethylpiperidinium chloride; chlorphonium, tributyl(2,4-dichlorobenzyl) phosphonium chloride; AMO-1618, N,N,N,2-tetramethyl-5-(1-methylethyl)-4-((1-piperidinylcarbonyl)oxy)benzenaminium chloride; ancymidol, cyclopropyl-(4-methoxyphenyl)-pyrimidin-5-ylmethanol; flurprimidol, 2-methyl-1-pyrimidin-5-yl-1-[4-(trifluoromethoxy) phenyl] propan-1-ol; HOE 074 784, 1-(2,6-diethylphenyl)-imidazole-5-carboxamide; tetcyclacis, 1-(4-chlorophenyl)-3a,4,4a,6a,7,7a-hexahydro-4,7-methano-1H-(1,2) diazeto (3,4f) benzotriazole; paclobutrazol, 1-(4-chlorophenyl)-4,4-dimethyl-2-(1,2,4-triazol-1-yl) pentan-3-ol; uniconazole, (S)-E-1-(4-chlorophenyl)-4,4-dimethyl-2-(1,2,4-triazole-1-yl) penten-3-ol; inabenfide, 4’-chloro-2’-(alpha-hydroxybenzyl)-isonicotinanilide; daminozide, 4-(2,2-dimethylhydrazinyl)-4-oxobutanoic acid; prohexadione, calcium 4-(1-oxidopropylidene)-3,5-dioxocyclohexanecarboxylate; trinexapac-ethyl, ethyl 4-[cyclopropyl(hydroxy)methylidene]-3,5-dioxocyclohexane-1-carboxylate; SL, strigolactones; TIS108, 6-phenoxy-1-phenyl-2-(1H-1,2,4-triazol-1-yl) hexan-1-one; abamine, methyl 2-[[(E)-3-(3,4-dimethoxyphenyl) prop-2-enyl]-[(4-fluorophenyl) methyl] amino] acetate; tebuconazole, 1-(4-chlorophenyl)-4,4-dimethyl-3-(1,2,4-triazol-1-ylmethyl) pentan-3-ol; 2-MN, 2-methoxy-1-naphthaldehyde; 4RG, 1-[4-(4-hydroxy-but-1-ynyl)-benzyl]-4-(3-trifluoromethyl-benzyl)-piperidine-4-carboxylic acid ethyl ester; TFQ0010, (3R, 4S)-3-methyl-4-phenethyloxetan-2-one; rhodestrin, (2E,4E,6E,8E,10E,12E,14E,16E,18E)24-hydroxy-2,6,10,14,19 pentamethyltetrecosa-2,4,6,8,10,12,14,16,18nonenyl-2(hydroxymethyl)-1H-indole-3-carboxylate

Stronger auxins for rooting

Auxins used to induce rooting are active in high concentration dips, often at 2000 mg/L (10 mM IBA) or higher, so it may be a reasoned consideration to employ “strong” auxins to lower the required concentrations or induce rooting in recalcitrant genotypes. Auxin-like compounds considered “strong” auxins include the auxinic herbicides, which typically cannot be applied at concentrations near the high-inductive doses often required due to phytotoxicity (Bottoms et al. 2011). Some have observed that auxinic herbicides at lower concentrations often induce callus and not AR (Verstraeten et al. 2013), although induction of AR has been noted as an aspect of herbicide damage (Warmund et al. 2021). The discovery of the TIR1/AFB-Aux/IAA auxin receptor (see Morffy and Strader 2022) has, nevertheless, renewed interest in auxins with very high receptor binding and less herbicidal effects. These include synthetic auxins such as NAA, 2,4-D, 4-chlorophenoxyacetic acid (4-CPA), 2-methyl-4-chlorophenoxyacetic acid (MCPA), and 2-(2,4-dichlorophenoxy) propionic acid (2-DP). In addition, 4-chloroindole-3-acetic acid (4-Cl-IAA), a naturally occurring auxin (Magnus et al. 1997), as well as related Cl-IAAs (Antolic et al. 1999) including 5,6-dichloroindole-3-acetic acid (5,6-diCl-IAA), have very high auxin activity, ranging from 10X that of IAA to 20X or more in some growth assays. Tests with many chloro-substituted IAAs found these forms to be quite active auxins (Engvild 1994; Antolic et al. 1999) rather than toxic (Slovin 1997) and this phenomenon has been confirmed in several studies. In fact, many of the monochloro and di-chloro-IAA compounds were later shown to have very high TIR1/AFB-Aux/IAA receptor binding activity (Jayasinghege et al. 2019) and minimal toxicity. Few root-induction or growth studies have further tested these compounds in practice. When tested, halogenated auxins showed promise for root induction. 5,6-diCl-IAA-methyl ester treatment significantly increased root numbers on hypocotyl cuttings of mung bean at a lower concentration than IBA (Pan and Tian 1999). Most reports were tests in bioassays other than difficult-to-root systems, possibly because they are considered rare compounds. Fortunately, facile methods for their synthesis exist (Cohen et al. 2022) and some are currently commercially available. IBA derivatives like 4-Cl-IBA and 5,6-diCl-IBA have received scant attention, although an old patent was issued (Marumo et al. 1991). We have found in preliminary studies that the halogenated IBAs have significant activity in vitro, suggesting that they are processed in the peroxisome like IBA to yield the active halogenated auxin.

Other unusual substituted IAAs have also been tested for enhanced rooting ability. For example, 4-trifluoromethylindole-3-acetic acid (4-TFM-IAA) was shown to be about 50% better than IBA at root induction in black gram [Vigna mungo (L.) Hepper] cutting, but only half as effective as 4-Cl-IAA (Katayama et al. 2008). α-Alkyl IAA derivatives were described as small-molecule agonists and antagonists of TIR1 receptor function (Hayashi et al. 2008), making them excellent candidate molecules for studies of AR, but at least in the initial reports, this biological activity was not studied.

Longer sidechain indolealkanoic acids

IAA almost never works as well as IBA at inducing rooting, and this observation has perplexed plant propagators who have not yielded a satisfactory explanation of why this happens. In some biological assays of root formation, IAA and IBA act quite differently (Chhun et al. 2004). IBA is predominantly moved as conjugates (Liu et al. 2012) but also has unique uptake mechanisms and is a saturable process (Rashotte et al. 2003), suggesting IBA uptake is carrier mediated and distinct from those involved in IAA uptake (Michniewicz et al. 2014; Frick and Strader 2018). The observation that plants convert IBA to IAA seemed to provide an easy answer to why IBA was effective (Kreiser et al. 2016). However, when concentrations of applied IAA and IBA were used on apple cuttings to produce a similar increase in internal IAA levels, IBA still gave more roots than IAA (Van der Krieken et al. 1992, 1993), so this simple idea cannot be the whole story. Although longer-side chain indolealkanoic acids have shown biological activity in growth studies comparable to that of IAA or IBA (Fawcett et al. 1960), their use in root induction appears to not have been done in difficult-to-root plants, and there is not much information on their utility (Van der Krieken et al. 1997). As with IBA, the metabolic fate of indole-3-caproic acid (ICapA; C6) is not certain, although it would be expected to mimic IBA in some regards but requires two rounds of β-oxidation to reduce the side chain from six carbons to two (Song et al. 2021). Understanding longer-chain compounds could help understand the differences between IAA and IBA in terms of the developmental signaling leading to AR initials. Also, as one increases the chain length, the compounds become more lipophilic, which could change uptake properties and require more β-oxidation activity to derive IAA. Compounds synthesized with up to an 11-carbon side chain, indole-3-undecanoic and (IUndecA; C11), have been reported. The longest even-numbered side chain for which a published synthesis is available appears to be indole-3-octanoic acid (IOctA; C8) (Avramenko et al. 1970), although the decanoic acid compound (C10) should be possible by the same procedures.

Increase rates of β-oxidation

The conversion of IBA → IAA and perhaps ICapA → IBA → IAA and longer chain conversions can also be studied in the presence of the non-auxin probe naxillin that appears to increase IBA → IAA conversion by β-oxidation in the root cap (De Rybel et al. 2012). Mutant analyses suggest that naxillin requires the endogenous IBA conversion pathway and thus acts through IBA-derived IAA to promote the development of lateral roots. Its function in tissues other than the root cap has not been extensively investigated, nor has its effect in AR or on auxin metabolism beyond the targeted reaction IBA → IAA been reported. However, other than attempting to increase the rate of uptake of IBA by methylation (Avery et al. 1937; Zimmerman and Hitchcock 1939; Rayle et al. 1970; Schenck et al. 2010) and optimization of methods of application (Blythe et al. 2007), there are few alternative chemical approaches for altering the efficacy of IBA itself.

A so-far unexplored method for up-regulating the targeted reaction IBA → IAA might be to change the carbon source for explants to favor enhanced peroxisome function. Peroxisome biology is complex, but it is clear that both stress and the need for fatty acid β-oxidation results in significant changes in metabolic capacity (Pan et al. 2020) and rates of pexophagy (Reumann and Bartel 2016). Application of mild stress or growth on fatty acids enriched media could potentially change the capacity of plants in tissue culture for higher peroxisome functions (Poirier et al. 1999) and thus potentially improve IBA → IAA activity.

Auxin conjugated forms

Auxin conjugation appears to play an important and complex role in the efficacy of AR (Haissig 1974). Landmark studies by Haissig (1989) described several esters of IAA and IBA for use in cutting propagation (Boyles et al. 1983). These included the aryl ester and aryl amide forms of IAA and IBA referred to as phenyl-IAA, phenyl-IBA, phenyl thioester-IBA, and phenyl amide-IBA. Their study showed these compounds to be more effective alternatives to the use of IAA, IBA, and NAA. Such modified derivatives have not received wide attention but have been reported to improve rooting and growth in oak and maple (Struve and Arnold 1986a, b; Struve and Rhodus 1988). While the role of conjugation remains complicated with many unresolved issues regarding its role in AR regulation, progress with the development of methods to study this has been observed in recent years. In Arabidopsis, the argonaute1 (ago1) mutants rarely form AR, and AUXIN RESPONSE FACTOR 17 (ARF17), which represses Gretchen Hagen 3 (GH3) gene expression and, thus, auxin conjugation, negatively regulates AR formation in ago1 mutants (Pacurar et al. 2014). Earlier studies of the metabolic fate of applied IBA showed that indole-3-butyryl-L-aspartate (IBAsp) levels reached a maximum 1 day after IBA treatment of cuttings. The conjugates thus formed were active in inducing the rooting of cuttings, with IBAsp being superior to free IBA. It was suggested that IBAsp might serve as an important source of auxin during later steps in AR (Wiesman et al. 1989; Riov 1993). Other IBA conjugates, such as IBAla, have also shown higher activity than IBA (Epstein and Wiesman 1987; Mihaljević and Salopek-Sondi 2012), while under specific conditions, IAA conjugates have also shown activity (Zelená and Fuksová 1991). Several different “slow release” forms have been tested for rooting, including different linkages to bovine serum albumin as well as IBA-anhydride, IBA-amino acids, IBA-polyamine-IBA, and IAA-polyamine-IAA, some with significantly positive results (Van der Krieken et al. 1997). Some synthetic auxin conjugates have shown rooting activity even when their “parent compound” was less effective. For example, auxin conjugate 2-(2,4-dichlorophenoxy) propanoic acid-glycine methyl ester (2,4DP-glyMe) was effective for vegetative propagation of mature pine tree cuttings (Riov et al. 2020). A recent report using a focused chemical screen of conjugated forms of four synthetic auxins (NAA, 2-DP, MCPA, and 4-CPA) identified 4-chlorophenoxyacetic acid-L-tryptophan-O-methyl ester (4-CPA-TrpMe) as enhancing the effect of K-IBA on AR in several recalcitrant woody plants (Roth et al. 2024). 4-CPA-TrpMe was shown not to interact directly with the TIR1-Aux/IAA7 auxin-perception complex, thus the activity seems to be related to the slow release of 4-CPA. However, it should be noted that tryptophan conjugates of IAA or jasmonic acid (JA) are endogenous auxin inhibitors (Staswick 2009). Indole-3-acetyl-L-tryptophan (IATrp) inhibited root gravitropic growth in seedlings, greatly reduced root inhibition from applied IAA, and inhibited the stimulation of lateral roots by IAA (Staswick 2009). Finally, and possibly related, various conjugates of 3-phenyllactic acid and tryptophan and esters of the conjugates showed good rooting activity in an Adzuki bean bioassay (Maki et al. 2022). While in combination, these studies would suggest that the ability to form and hydrolyze conjugates is important for AR, the issue is far from resolved (Salope-Sondi et al. 2015).

Inhibiting the induction of amide conjugation

While studies with applied auxin conjugates, mutants in conjugation, and the process of conjugation during IBA application logically all seem to suggest that conjugate formation and hydrolysis are both important aspects of AR, the results with small molecule inhibitors of auxin conjugation point to a specific role for conjugate formation. The first reported inhibitor of auxin amino acid conjugation, 2,6-dihydroxyacetophenone (DHAP) (Lee and Starratt 1986), doubled the number of roots in IBA treated cuttings (Epstein et al. 1993). Better characterized inhibitors of IAA-amido synthetase activity such as adenosine-5’-[2-(1H-indol-3-yl) ethyl] phosphate (AIEP), which mimics the adenylated intermediate of the GH3 auxin-amido synthetase reaction (Böttcher et al. 2012), bring specificity to studies of the enzyme activity resulting in indole auxin amino acid conjugation. Cano et al. (2018) found that in difficult-to-root carnation stem cuttings enhanced conjugation of auxin by GH3 enzymes leads to poor AR, and rooting ability could be restored with AIEP, the inhibitor of conjugation. After the description of AIEP, kakeimide (Hayashi et al. 2021; Fukui et al. 2022) and nalacin (** functions in their signal pathways that regulate aspects of plant development including root growth and tissue expansion. Both auxin and GA signaling have several points of convergence that allow crosstalk for the regulation of developmental events (Franklin et al. 2011; Richter et al. 2013), but they are not redundant in their functions. However, GA could apparently alleviate aspects of Aux/IAA gain-of-function phenotypic expression (Frigerio et al. 2006). Such studies suggest that changes in GA levels can, in part, also mediate aspects of auxin developmental activities (Hanson 1976). Other studies suggest that GA inhibits the formation of AR in plants by disrupting endogenous hormonal processes, including hormone levels and auxin transport (Willige et al. 2011; Mauriat et al. 2014; Li et al. 2023); however, the precise regulatory basis for “the observed interactions in root formation and plasticity are still to be discovered”.

Conclusions and perspectives

The discovery of the effects of IBA on AR 90 years ago had a profound impact on applied plant propagation methodology, bringing new forests, gardens, and fruit crops into wider use. AR, however, is a complex response, as might be expected for the generation of new organs from differentiated or partially differentiated tissues. The AR response is also sensitive to the plant’s environment, including light or darkness (Monteuuis and Bon 2000; Sorin et al. 2005; Klopotek et al. 2010; Pincelli-Souza et al. 2022), temperature (Corrêa and Fett-Neto 2004), as well as stress factors such as water availability and mineral nutrition (De Almeida et al. 2017). Regardless of the advances in research that have established IBA natural occurrence, the conversion of IBA to IAA as a critical process, and an array of hormonal and environmental needs, most of the commercial plant propagation remains primarily focused on the proper application of IBA, NAA, or combinations of the two (Sharma and Thapa 2022). It should be possible to do better, and as outlined in the review there are several opportunities to improve and many avenues to explore. A quickly evolving library of chemical effectors suggests new opportunities for future investigations that will allow us not only practical avenues for pretreatments but will also enable us to identify the molecular and signal transduction mechanisms underlying adventitious rooting as a key determinant in clonal propagation efficiency. A better understanding of cellular signals and regulatory cascades in development that are involved in adventitious root formation, underpinned by advances in systems and chemical biology, will provide a more complete understanding of rooting recalcitrance. It is the hope of this review that it will encourage a 21st century effort to bring new ideas to both research and the application of AR for plant improvement.