Open access peer-reviewed chapter

Mosses: Accessible Systems for Plant Development Studies

Written By

Jordi Floriach-Clark, Han Tang and Viola Willemsen

Submitted: 16 September 2021 Reviewed: 20 September 2021 Published: 29 October 2021

DOI: 10.5772/intechopen.100535

From the Edited Volume

Model Organisms in Plant Genetics

Edited by Ibrokhim Y. Abdurakhmonov

Chapter metrics overview

497 Chapter Downloads

View Full Metrics

Abstract

Mosses are a cosmopolitan group of land plants, sister to vascular plants, with a high potential for molecular and cell biological research. The species Physcomitrium patens has helped gaining better understanding of the biological processes of the plant cell, and it has become a central system to understand water-to-land plant transition through 2D-to-3D growth transition, regulation of asymmetric cell division, shoot apical cell establishment and maintenance, phyllotaxis and regeneration. P. patens was the first fully sequenced moss in 2008, with the latest annotated release in 2018. It has been shown that many gene functions and networks are conserved in mosses when compared to angiosperms. Importantly, this model organism has a simplified and accessible body structure that facilitates close tracking in time and space with the support of live cell imaging set-ups and multiple reporter lines. This has become possible thanks to its fully established molecular toolkit, with highly efficient PEG-assisted, CRISPR/Cas9 and RNAi transformation and silencing protocols, among others. Here we provide examples on how mosses exhibit advantages over vascular plants to study several processes and their future potential to answer some other outstanding questions in plant cell biology.

Keywords

  • bryophyte
  • moss
  • model organism
  • plant development
  • regeneration
  • cell polarity
  • reprogramming
  • asymmetric division
  • stem cell
  • water-to-land
  • 2D-to-3D

1. Introduction

1.1 Mosses in context

Mosses are plants that belong to the Bryophytes, a cosmopolitan sister group of vascular plants with the last common ancestor between 400 and 500 million years ago [1, 2]. As a mostly avascular lineage, Bryophytes, that include mosses, liverworts and hornworts, thrive in mostly moist niches near the surface and stay compact (<10 cm), with some neovascularised exceptions that grow up to 65 cm [3, 4, 5]. Their cosmopolitan distribution in a variety of biotopes including moist and arid environments, can be explained by unique adaptations like drought, freezing and salinity tolerance [4, 6, 7]. Mosses’ life cycle is dominantly gametophytic (the photosynthetic and growing phase is haploid), and the size and architecture of their organs is smaller and simpler than that of vascular plants, with leaf-like structures (phyllids) and sexual organs (antheridia and archegonia) of often only one cell of thickness, stem-like structures of circa ten cells and spore-bearing containers (sporangia) of single-cell spores [4, 8, 9, 10].

This miniaturised body renders mosses accessible systems for the study and dissection of cell and molecular aspects of plant biology that require close monitoring in time and space [11]. Such studies greatly benefit of the accessibility to single cells in a multicellular context. For instance, asymmetric and directional cell divisions are key life developmental tools to build an organism, but we lack understanding on how these processes are exactly controlled and regulated [12]. Importantly, these developmental drivers are shared between mosses and vascular plants, and to some extent with animals, and associated gene functions seem to be highly conserved despite the long period of independent evolution [13].

In the last two decades, mosses have gained high interest in plant research, with Physcomitrium patens (Hedw.) Mitten becoming a central model system. P. patens organelle genomes were sequenced in 2003 (plastids) and 2007 (mitochondria) and nuclear genome in 2008, with the latest revision in 2018. In addition to this, a myriad of genetic tools has emerged that allow close study of all processes of this moss as a representative of this lineage of the Bryophytes.

Hereby, we present how mosses, thanks to their simplified body plan and genetic networks in development, and with special focus in P. patens, can become cornerstone model organisms to study several developmental processes that determine plant architecture of most land plants [11]. We selected a number of outstanding developmental processes that pose central research questions in developmental biology of plants and that started to be investigated in mosses in the last years. These processes are introduced in a bottom-up approach, with special attention to their molecular and cellular basis, going from the early stages to the final plant organisation, chronologically. The essential and most used tools available to investigate these aspects of mosses are briefly described to facilitate the initiation into this model system. Finally, we show how the moss revolution has recently started with the rise in moss genomes sequenced and increase in moss research with additional species and questions beyond P. patens.

1.2 Moss morphology and life cycle

As most land plants, mosses have alternating generations between the haploid gametophyte and diploid sporophytes. However, unlike vascular plants, mosses spend most of their life cycle in the gametophytic stage, in which most of the organism asexual development, including photosynthesis and growth, occurs. The sexual organs eventually develop at this stage to give rise to the embryo after fertilisation, that produces the sporophyte over the gametophyte. This fruiting stage is diploid until haploidisation in spore formation takes place [11, 14].

Starting from a spore, the first developmental stage of the moss is the chloronema, a chloroplast-rich and single cell-wide filamentous tissue that serves for initial colony expansion, early photosynthesis and nutrient absorption. The cells are slightly elongated (~80 μm long), and the intercellular cell walls are oriented perpendicular to the growth direction [15]. This filament eventually transitions to caulonema, a quick-growing filamentous cell type that have underdeveloped chloroplasts at early stage, with longer and narrower cells (~250–300 μm long) that grow twice as fast, and with oblique intercellular cell walls [15, 16, 17]. This tissue has exploratory purposes and is favoured in stressful, light-poor, and nutrient-poor conditions, possibly with the aim of finding more suitable conditions [18, 19]. Caulonema can transition again to (secondary) chloronema [17]. The filamentous tissues are collectively referred to as protonemata and can laterally grow and divide to branch as new filaments. The protonemata grow in a mat-like fashion that shapes the two-dimensional (2D) developmental stage of mosses where the growth is confined in a plane of few millimetres of thickness.

Sometimes, the lateral cell outgrowth (the cell initial) gives rise to a bud cell instead of a branch cell, which is the beginning of gametophore development and the transition to three-dimensional (3D) growth [20]. The identity of the cell initial can be predicted by the division plane angle, implying that identity is determined before division (Figure 1). Currently, the list of known genes involved in the path selection and division plane orientation is growing, but the early determinants of bud formation and branching remain unknown [18, 21]. The bud grows by well-defined asymmetrical and oriented divisions to form the gametophore, the leafy shoot-like plantlet of mosses that ultimately bears gamete-producing organs.

Figure 1.

Scheme of morphology and location of different developmental processes.

The bud basal cell gives rise to a new type of filamentous tissue, the rhizoids, with pigmented, caulonema-like morphology. They function as anchorage to the ground to stabilise the up-growing gametophore and contribute to nutrient and water uptake, similar to roots and root hairs of seed plants, but with the tissue complexity of root hairs [22, 23]. The bud apical cells divide in precise directions to give rise to oriented phyllid initial cells with a particular phyllotactic pattern (i.e. lateral organ organisation around the shoot; e.g. spiral) to develop the gametophore.

The apical cells eventually arrest their proliferation, or terminally give rise to sexual organs (firstly antheridia, and later archegonia) under autumnal/spring conditions: short day (8 h), low light (20 μmol/m2/s) and low temperature (15°C) [24, 25]. Despite the asynchronous development of male and female gametangia, this moss is self-fertilising, and thus tends to genetically self-isolate [26]. Flagella-driven spermatozoids (male gametes) move towards the archegonial venter in liquid water and fertilise the egg cell to give rise to a diploid zygote. The zygote will subsequently develop, via an embryonic stage with a new 2D-to-3D transition, into the sporophyte, that consists of the foot (the interface with the gametophyte) and a short stalk (seta) with a terminal capsule. In the capsule, meiosis gives rise to up to few thousands of haploid spores [8, 26].

The first documented ecotype, known as ‘Gransden’ (United Kingdom, 1962), has reduced rates of sporophyte formation, probably due to long asexual propagation in laboratories [27]. In many laboratory lineages, it has become self-sterile, rendering it unattractive for studies dependant on sexual reproduction. On the contrary, the more recently isolated ecotypes ‘Villersexel’ (France, 2003) and ‘Reute’ (Germany, 2006) have 15 times more sporophytes (77% of total gametophores), indicating a high fertility rate [26]. Despite these differences, all ecotypes can be propagated asexually in identical conditions from any tissue thanks to the high regeneration rate of this moss, through which explants will redifferentiate into chloronema and initiate the life cycle from there [28, 29].

Advertisement

2. Outstanding developmental processes

Several essential developmental processes are shared between all land plants, including angiosperms and bryophytes. Strikingly, in mosses many shared processes take place with a simplified set of genes and sometimes in single cells. Therefore, the underlying genetic regulatory networks of development are easier to study. In this section, we highlight some of these processes and their unique ease of study in the moss P. patens.

2.1 The 2D-to-3D transition

The ability to structure organs in three dimensions (3D) was an essential feature for water-to-land transition. Many aquatic plants develop in a homogeneous environment, whereas land plants faced a highly distinct environment at ground surface level (a plane) and at the air/soil axis (perpendicular to this plane). This required land plants to develop specific tissues to efficiently grow in each dimension and cope with new challenges [20].

From a physiological perspective, the transition from 2D-to-3D growth in mosses consists in the development of complex structures such as the gametophore that grows out of the surface plane where protonema thrive, exhibiting negative gravitropism and positive phototropism [30, 31, 32]. The development of the gametophore relies on the ability of the moss to define different organisation in each dimension of space. The basal units from which this spatial assembly takes place are ultimately single cell primordia, from which organs emanate [33]. To this end, a cell must spatially sense intrinsic and extrinsic signals and transduce them to subcellular structures that pave the way to division plane positioning and subsequent asymmetric and oriented cell division. However, the molecular basis of spatial sensing and its transduction to organisation, action and maintenance of the division machinery has not been fully elucidated yet [12].

Cellularly, the 2D-to-3D transition in a moss occurs during bud formation (Figure 1). When a cell initial is formed in protonemata, in 5% of the cases it has a bud initial cell identity instead of a branch initial identity. Each bud initial swells and undergoes divisions oriented in all three axes of space [34]. This transition is supported and maintained by a different genetic and molecular machinery than that of protonemal development, and is the molecular basis of 3D growth. These proteins are specifically present from the first cell of the bud and onwards (e.g. DEK1, NOG1), and remains active in subsequent proliferating organs (e.g. shoot apical meristem, phyllids, gametangia) [13, 34, 35, 36]. In general, gradients or local clusters of proteins, peptides, nucleic acids, and hormones can be signals, sensors and/or actuators in developmental processes, and what is upstream of the cascade of 2D-to-3D transition remains a mystery. Some actors that are already on the radar include oscillations of auxin and cytokinin concentrations, ROP and SOSEKI proteins and CLE/CLV peptides-receptors, and they have been pinned down to different moments of the process and at different locations [12, 37, 38]. However, how do they coordinate their activity remains elusive and an active area of study.

In vascular plants, the 2D-to-3D transition occurs only once in their lifecycle, during embryogenesis, where orthologs of the essential moss 3D machinery genes are also expressed [34, 35]. This event is confined in the endosperm during embryogenesis and its observation in seed plants requires seed and ovule microdissection, which makes in vivo monitoring difficult [39]. Furthermore, knockouts of some of these genes are lethal in this early stage. On the contrary, P. patens exhibits both growth fashions (2D and 3D) simultaneously and frequent transition events (bud formation) during all its vegetative stage in each colony (months), and deletion or functional mutants are non-lethal due to the indefinite growth character of the remaining 2D tissues [34, 35].

These features provide a privileged seat for in vivo long-term tracking and subcellular study of molecular markers, gene expression and protein localization that allows to shed light to the necessary cellular events required establish 3D growth in single cells to build a full plant.

2.2 The shoot apical meristem/cell (SAM/SAC) assembly

The protagonist of 2D-to-3D growth transition is the formation of the shoot apical meristem (SAM), a region at the apical growth side responsible for the continuous formation of the aerial organs of the plant, including leaves and reproductive organs [40, 41]. The histology of the SAM in angiosperms describes a central zone with stem cells that self-renew, divide and radially differentiate into peripheral cells that determine organ initiation at specific locations (e.g. by placement and establishment of new leaf primordia) [42]. While angiosperms present multicellular stem cell centre SAMs during their sporophytic phase, bryophytes present an unicellular structure, the shoot apical cell (SAC), both in their gametophytic and sporophytic phase (Figure 1) [43]. Despite the differences, both SAM/SAC share a common organisation, with stem cell(s) at the centre, surrounded by regularly differentiating tissue [44].

The mechanisms that establish and maintain these pluripotent stem cell(s) in the SAM/SAC is unknown. It involves spatial sensing, cell-to-cell communication, and asymmetric and oriented cell divisions, which render mosses attractive systems for their accessibility. De novo establishment of SAC is especially easy to study in mosses because it occurs once per bud formation (hundreds of events per colony) and in sporophyte development after egg cell fertilisation (on top of ~77% of gametophores). These SAC establishment events consist of one relatively exposed cell, that is easy to monitor during several division rounds for weeks [16, 26]. Angiosperms present a SAM and numerous equivalent lateral meristems, sometimes big and manageable (e.g. cauliflower meristems), but in general their study requires dissection for visualisation and it consists of complex multicellular structures that complicate characterisation.

The developmental origin of the sporophytic SAC in P. patens is either a de novo SAC establishment after egg fertilisation or a gametophytic SAC redefinition [45, 46], and in any case the genetic and signalling basis and developmental mechanisms of its establishment seem conserved between angiosperms and bryophytes [22, 47].

The genetic make-up in both taxa has shown to rely on auxin response through AUXIN RESPONSE FACTORS (ARFs), cytokinin signalling by ARABIDOPSIS RESPONSE REGULATORS (ARRs), CHASE domain-containing histidine kinases (CHKs) and CYTOKININ OXIDASE/DEHYDROGENASE (CKX), local coordination through several transcription factors families like CLAVATAs (CLVs), CUP-SHAPED COTELYDONS (CUC), LATERAL ORGAN BOUNDARY (LOB) and signalling peptides like CLAVATA3/EMBRYO SURROUNDING REGIONRELATED (CLE) and chromatin modification by Polycomb Repressive Complex 1 and 2 (PRC1,2), among others [48]. Although many key factors have conserved roles in SAM formation and maintenance in seed plants and mosses, some important factors in angiosperms, like the key regulator of stem cell maintenance WUSCHEL, are not found in P. patens. These kind of differences can be insightful in defining the basic network to maintain stemcellness, tailoring a SAM and help understanding cell identity switch and organ formation [44].

2.3 Phyllotaxis from a cell

The most noticeable outcome of the shoot apical meristem/cell (SAM/SAC) activity is the organised and oriented initiation of leaf primordia along the stem, which leads to a unique geometric pattern of leaves and shoot branches named phyllotaxis [49]. In land plants, phyllotaxis may be defined by both genetic and environmental factors (light abundance, wavelength intensity ratio, etc.). For instance, leaf organisation can be adapted by shade avoidance syndrome [50]. However, only the genetic factors are shown to play a role in organ primordium location determination. A phyllotactic pattern is quantified by a fraction in which the denominator is the number of organs of the same type until the same orientation repeats (e.g. in P. patens, every fifth phyllid lies almost exactly below or above the first) and the numerator is the number of turns it takes (e.g. two turns in P. patens). This ratio (2/5) is then the fraction of a turn (e.g. 2/5 x 360°) or angle between two consecutive organs. When the angle between organs tends to the golden angle (137.5°, with fractions of turn derived from the Fibonacci sequence: 2/5, 3/8, 5/13, etc.), a spiral pattern emanates. Different angles can be observed in different species, like e.g. the 180° angle that gives rise to a distichous (or alternate) pattern or 120° for a tristichous pattern. Both Arabidopsis and P. patens follow a spiral pattern [43, 49, 51, 52, 53].

However, the pattern arises from essentially different SAMs/SACs: in angiosperms, phyllotaxis derives from a multicellular system with well-reported oscillating auxin peaks around the SAM growth axis, whereas P. patens effectively generates a pattern from a single apical cell (SAC). During the first division rounds in bud formation, the initial cell divides asymmetrically and gives rise to an inverted tetrahedral SAC with three lateral faces (Figure 1) [24, 54]. An oriented cell division of the SAC produces a new central SAC and a peripheral derivative cell, the merophyte, which develops into the future phyllid and a portion on the stem. The change of the stem cell division plane orientation in the SAC in each round results in a spiral phyllotactic pattern of the phyllids [55], which requires some unknown round-to-round cue to achieve rotation. Surprisingly, the rotation direction or chirality of the division orientations appears to be randomly determined, showing both clockwise (S) and counter clockwise (Z) patterns, yet there is high frequency of switch from one to the other (antidromy) in branches of gametophores of other moss species [56].

The limited understanding of the origin and underlying molecular mechanisms of this rotating pattern and derived phyllotactic pattern is largely confined to the sporophyte of the evolutionarily recent group of flowering plants (angiosperms). The available transcriptomic data of bud and tip cells and gametophores (or 3D shoots) may provide more insight in the transition from uniplanar to triplanar meristematic growth in moss [48].

Aligned with the phenotypic similarities of moss and angiosperm phyllotactic patterns, several factors known from Arabidopsis have also been found in mosses, including receptor signalling genes involved in shoot meristem size and patterning, hormone biosynthesis genes, transcription factors that control cell-specific mechanism of developmental pattering, chromatin remodelling complexes and cell cycle [48]. Many of these factors are essentially executive and likely controlled by some spatial sensing machinery. Comparing them with new contributors or absent members in the minimal regulatory network of mosses may help unravelling the fundamental elements that trigger the orientation-specification machinery that greatly impacts plant architecture in all land plants, including relevant crops.

2.4 Regeneration

Most described processes in this chapter require cell identity acquisition and maintenance. In certain circumstances (e.g. wounding), differentiated plant cells can reprogram to become new stem cells, divide and redifferentiate for organ regeneration [57, 58]. P. patens is an excellent system to investigate cell reprogramming and regeneration due to its fast and broadly occuring cell pluripotency [59].

In most tissue cultures of other plant species, exogenous hormones (e.g. auxin and cytokinin) are required to induce callus formation and plant regeneration [60]. However, in moss, cells are capable of regenerating from protoplasts or excised phyllids into new protonema filaments in the absence of exogenous hormones (Figure 1) [61]. This implies that the whole regeneration toolkit is present in mosses and can be endogenously activated on demand, which makes them different from other taxa (e.g. angiosperms) [62].

When a phyllid is excised, cells neighbouring the cutting edge can reprogram from somatic cells to protonemal stem cells, which can then start a new life cycle [14]. This regeneration process is easy to study in mosses for several reasons: firstly, cell identity conversion can be easily tracked with protonema stem cell reporters [29]; secondly, aside of the simplicity of in vivo observation (see section Imaging), the unistratose (i.e. single cell-layered) phyllid simplifies single-cellular extractions (e.g. laser ablation) for single cell omics and other high precision studies [63]. Interestingly, the result of the reprogramming cascade is timely visible 48 hours after cutting, which also allows large scale collection of excised tissue for time-course tracking of gene expression evolution during regeneration activation [64].

The mechanistic studies of the cell fate acquisition can benefit from this simple cell type conversion in comparison to other model systems used in cell reprogramming investigation (e.g. regenerative callus or Arabidopsis roots that consist of multiple cell types that possess different tissue identity) for its minimality and event frequency [65, 66]. In angiosperms, regeneration is often reduced to localised stem cell pools (e.g. the base of leaves), takes longer to establish, and it is multicellular and asynchronous at the explant level [58, 67].

Previous studies have taken advantage of the abovementioned features to investigate gene expression profile during phyllid cell reprogramming, which revealed that genes involved in stress, proteolysis, and hormone signalling pathways are induced from 6 to 24 h after cutting [64]. Some genes have been demonstrated to play essential roles in moss leaf reprogramming, including Cyclin-dependent kinase A (CDKA), found to link cell cycle reactivation and other cellular responses that promote cell outgrowth as a new protonema filament [29]. Similarly, the outgrowth of reprogrammed protonema cell requires WUSCHEL-related homeobOX 13-Like (WOX13L) genes and Cold-Shock domain Protein 1 (PpCSP1), induced in the cells facing the cutting edge within 24 h [68, 69]. Finally, an AP2/ERF transcription factor STEMIN1 (STEM CELL-INDUCING FACTOR 1) was discovered to induce cell reprogramming in moss leaves without excision or wounding [70]. These studies have identified new pieces in the puzzle of cellular reprogramming, and future studies will aim to unravel mechanisms behind the cell identity conversion and reprogramming.

One interesting feature of moss regeneration is inhibition of neighbouring cells. The necessary cell–cell (apoplastic e.g. Ca2+-mediated) or cell-to-cell (plasmodesmata-mediated) communication makes regeneration an attractive developmental process to study this cell crosstalk [71, 72, 73]. The phytohormone ABA is a key responsible of the dynamic regulation of the permeability of plasmodesmata in response to changing environments, such as wounding. Control in plasmodesmata pore size can influence the signalling molecules that can pass through or can be blocked in particular cells, which can have a direct effect in development of the processes mentioned until now [74, 75].

2.5 Hormone regulation

The signalling pathways and functions of plant hormones are substantially conserved in P. patens. Given the differences in physiological structures and relative evolutionary positions between angiosperms and bryophytes, mechanistic studies of hormone regulation in mosses can bring new insights in the hormone regulatory networks of all plants that resolve current questions.

Three plant hormones—auxin, cytokinin and strigolactone—have shown to regulate shoot branching patterns (phyllotaxis) and activation in angiosperms. Auxin moves down the main shoot of angiosperms to inhibit branch development, while cytokinin promotes branching. In addition to branching, auxin is the key molecule in the control of plant growth and development, and promotes organ differentiation [76, 77]. Exogenous application of auxin or its inhibitors results in irregular cell shapes and inhibit lateral organ formation, for instance in shoot apical meristem (SAM) maintenance. The understanding of hormone regulation and signalling in angiosperms progresses slowly due to tissue and gene network complexity. In P. patens interfering with auxin transport via the auxin efflux protein PIN-FORMED (PINs) knock-outs reveals the same effects on SACs as that have been observed in Arabidopsis SAMs. Also, the interaction between core components in auxin signalling and their response to auxin in P. patens is also conserved when compared to Arabidopsis [78, 79, 80]. Furthermore, exogenous application of auxin leads to termination of gametophore and differentiation into rhizoids, as it happens with shoots and roots in Arabidopsis [81]. In protonema cells, PIN-mediated auxin transport is essential for the chloronema-to-caulonema transition. When PINs are overexpressed, tip auxin levels deplete, which results in cell fate transition inhibition, while the PIN knock-out mutants show a faster transition from chloronema to caulonema [82]. Despite of these similarities, it has been shown that mosses may not weave their architecture with PIN-based transport as angiosperms do. On the contrary, they require bi-directional auxin transport to generate the observed patterns of shoot branching, as was confirmed by modelling and empirical evidence [83]. It derives that plasmodesmata-based transport may play a key role, which renders cell-to-cell communication essential in plant architecture definition and has not been reported in angiosperms [83].

Cytokinins and strigolactones also influence plant architecture, both in angiosperms and mosses. The levels of cytokinin are high and precisely distributed in the central stem cell region of SAM in angiosperms to maintain stemcellness [84]. In the root apical meristem, auxin and cytokinin act antagonistically in meristem size control, but its levels, distribution and interaction in P. patens single apical cell environment are unknown. Despite that, both hormones are present in this moss and are likely to play a role. The application of high concentrations of cytokinin in culture causes ectopic shoot formation and inhibition of leaf formation [83, 85]. Also, in gametophore development, cytokinin inhibits rhizoid formation by opposing auxin, like in roots of Arabidopsis. As expected by this homologous functionality, the mutants that stimulate cytokinin degradation lead to a strong increase of rhizoids in both number and length [86]. The last mentioned hormone, strigolactone, is reported to inhibit shoot branching in angiosperms and its localisation is restricted to the base of shoots. The same compound is able to stimulate the pattern of shoot branching in P. patens. In filamentous tissues, strigolactone is produced to inhibit chloronema branching and to regulate the colony extension [15, 87].

As shown, many processes that define plant architecture are regulated in similar ways both in angiosperms and mosses. However, mosses offer a reduced gene network and regulation complexity that facilitates the analysis of hormone functions in the related developmental processes. Furthermore, subcellular and tissue-level transport and distribution of hormones can be best visualised in their simple plant bodies. In such plant models, new hormone functions will prove to be easier to study and translate to agronomically relevant plants.

Advertisement

3. Protocols and tools

Many valuable online resources with information on protocols, stocks, tools and genetic information have been exhaustively compiled elsewhere [11]. Hereby, we provide some additional information and summary of the essentials of research in P. patens.

3.1 Imaging

3.1.1 Accessibility

The small size and simple architecture of moss organs allows detailed microscopic visualisation easy to accomplish in almost all tissues. The strings of cells in protonema and their branching is trivial to closely visualise, and the transition to gametophores can be well tracked until the stem-like centre becomes slightly thicker than a dozen of cells and grows out of the plane. From it, the leaf-like structures (phyllids) have only one cell of thickness except in the midrib and can be tore apart for up-close visualisation. The terminal sexual organs (antheridia and archegonia) have a 3D structure that is easy to fully dissect or directly visualise due to their monolayered sack structures. The subsequently developed spore-bearing containers (sporangia) are full of single-cell spores [4].

This miniaturised body renders mosses accessible systems for the study and dissection of cell and molecular biology that require close monitoring in time and space. Such studies greatly benefit of accessibility to single cells in their context for observation of subcellular responses in vivo, e.g. protein localization, cytoskeleton rearrangement, and cell divisions along the developmental progress in a better resolution than most other multicellular plant tissues.

3.1.2 Reporter or marker lines

In P. patens, fluorescent marker lines that label different organelles (e.g. ER, chloroplasts, mitochondria, peroxisomes, Golgi apparatus, vacuoles, and nucleus) are available. In addition, given the predictable division patterns of the protonema tip cells, P. patens has been extensively used to investigate mitosis. Therefore, marker lines containing fluorescently labelled proteins such as several microtubule-associated proteins relevant to cell divisions were generated for mitosis imaging. Other published reporter lines show the concentration of the hormone auxin (DR5, GH3 and R2D2), cell identities, like protonema-specific proteins (RM09 and RM55) or mature rhizoids (RSL1,2), and developmental switches such as 2D-to-3D transition markers. In Table 1, references to all these reporters are indicated.

VisualisedFusion/TargetPurposeReference
NucleusNLS4-GFP-GUSNuclear localisation[21]
Endomembrane systemEndoplasmic reticulum[11]
α-1,2-mannosidaseGolgi apparatus imaging[88]
Targeting signal type 1 (SKL)Peroxisome imaging[88]
MitochondriaCytochrome c oxidaseMitochondria imaging[88]
CytoskeletonLifeActActin cytoskeleton[89]
Tubulin αMicrotubule cytoskeleton[90]
MAP65Antiparallel Microtubule-microtubule contacts[91]
Kinesins[90]
Plasma membraneSNAP-TM-mCherryMembrane tracking[21]
AuxinpDR5v2:GFP-GUSAux. induced fluor.[92]
GH3:GFP-GUSAux. induced fluor.
R2D2Ratiometric induction
ProtonemapRM09:NLS4-GFP-GUSProtonema identity reporting[29]
pRM55:NLS4-GFP-GUSProtonema identity reporting

Table 1.

Compilation of key molecular reporter lines published in literature to study cell and developmental processes in P. patens.

3.1.3 Microfluidics

Despite the advantageous physiological features of P. patens, observing cellular and subcellular processes with high resolution and for long periods of time is challenging. Traditionally, monitoring the intrinsic changes involved in regeneration, tip growth, bud formation, gametophore development and phyllid development, such as cytoskeleton organisation, protein distribution, organelle location, etc. has been done in glass-bottom petri dishes for as long as culture media could sustain, or in coverslip-sandwich sample preparations for up to few hours, due to lack of gas exchange [16, 37, 93, 94, 95]. However, the advent of microfluidics for bioimaging offers a new tool to overcome some limitations.

Microfluidic devices are transparent and flexible structures commonly produced using the biocompatible and air-permeable polydimethylsiloxane (PDMS) polymer. In biology, they have been used for study and imaging of cell and tissue development in vivo, including 3D development. For instance, it has been beneficial for high-throughput Arabidopsis root research [96]. Remarkably, the growth fashion of the pollen tube and the embryo development in Arabidopsis or the protonemal growth and early gametophore development of P. patens are ideal candidates for high-throughput and high-resolution imaging in microfluidic devices with light- and fluorescence-based microscopies [93, 97].

Until now, P. patens-tailored microfluidic devices have proven to be a reliable system for the monitoring of previously mentioned processes, as they offer tracking for up to weeks thanks to the air-permeability and possibility to refresh the media by circulating it from a reservoir [16]. Furthermore, it is then possible to introduce chemical agents or co-culture other organisms to image their cytological and gene expression effects on the plant over long periods of time [93]. The close tracking allows for quantitative cell measurements such as biomechanic parameters, growth rate and size of different tissues, frequency and geometry of divisions, developmental time and pace studies, etc. Microfluidic devices can capture subtle phenotypes of mutant lines for full analysis and high-quality phenotype reporting [16, 93].

3.2 Manipulation

3.2.1 Forward genetics

Some groups carried out forward genetic screening by X-ray or chemical mutagens that generated mutants with hormone resistance or abnormal tropic responses [85, 98]. However, due to the lack of genomic information, the disrupted genes that caused the phenotypes were never identified.

Recently, the completion of genome sequencing and the establishment of its genetic mapping tools removed the obstacles in the forward genetic screening of P. patens. To establish genetic mapping, two genetically divergent ecotypes of P. patens, Gransden and Villersexel were used [99].

With this genetic mapping resource, in the past few years, researchers started to perform forward genetic screening by treating protoplast with UV light. In such screenings, mutants with phenotypic defects were successfully obtained and the causal lesions were identified by outcrossing and whole genome sequencing. Notably, essential genes that are crucial for growth may not be identified due to the lethality of their knock-outs; therefore, conditional screening was performed to overcome this problem. After the UV light treatment, plants were cultured under different temperatures and in such conditions, plants showed growth defect only in high temperature were selected as a temperature-sensitive mutant [100].

Another screening aimed to discover genes that are essential for the 2D-to-3D transition is also limited by developmental defects, given that mutants in this process cannot produce gametophores necessary for sexual crossing. To overcome this problem, instead of crossing, researchers generated somatic hybrids between Villersexel mutants and Gransden wildtype, which produced diploid sporophytes [34]. Spores released from this hybrid sporophyte exhibited consistent phenotypic segregation ratio with meiosis. Mutant plants generated from these diploid spores were sequenced and genomically mapped to achieve the identification of new crucial genes for moss 2D-to-3D transition (e.g. NO GAMETOPHORES 1 and 2, or NOG1, NOG2).

In addition to UV light, tobacco Tnt1 retrotransposon was used to produce insertional mutations in genic and GC-rich regions [101]. Both PEG- or Agrobacterium-mediated transformations were applied and successfully produced mutants.

3.2.2 Gene identification

P. patens was the first moss fully sequenced. Organelle genomes were sequenced in 2003 (plastids) and 2007 (mitochondria) and nuclear genome in 2008, with the latest fully annotated revision made and genetic mapping obtained in 2018 [47]. This information and the molecular tools available allow targeted mutagenesis to dissect functions of genes of interest. Additionally, full genome structure and SNP variation between four main ecotypes (Gransden, Reute, Villersexel and Kaskaskia) was reported in 2017, completing the toolbox for reverse genetics and bioinformatics research.

3.2.3 Neutral locus integration

The integration of DNA constructs (including promoter, gene of interest and selection cassettes) necessary to produce transformants with stable expression and non-disrupting phenotypes requires targeting of neutral loci that do not intrinsically produce a phenotype when disrupted, often due to gene redundancy. In Table 2, there are several standard neutral loci indicated which reportedly showed no visible phenotypes or morphological defects when it is replaced by an entire gene expression cassette.

LocusVectorPurposeReference
PIG1 locuspGX8
pGG626
XVE inducible overexpression
XVE inducible RNAi expression
[102, 103]
Pp108 locuspUGGiConstitutive RNAi expression[104]
Pp108 locuspTH-Ubi-GateConstitutive expression by the maize ubiquitin promoter[105]
Redundant copy of the ARPC2 genepTK-Ubi-Gate
Redundant copy of the ARPC3 genepTZ-Ubi-Gate
PTA1 locuspT10GOverexpression by EF1α promoter[106]
BS213 locuspMJ1[107]

Table 2.

A compilation of vectors designed to target proven neutral loci in P. patens.

Currently there are several vector sets released to specifically target neutral loci, that contain their flanking regions homologous to parts of the locus at start and end of the vector. Cloning the gene of interest in between readily allows replacement of the targeted locus with the entire cassette via homologous recombination (see section Homologous recombination).

In P. patens, the gene loci have three commonly seen annotations in literature. The first and standardised since 2017 (with the third chromosome annotation version) is PpGcX_uuyyyVn.m, where G is the genome release version (version 3), c stands for chromosome, X stands for chromosome number (from 1 to 27), and uuyyy is an arbitrary flexible number that indicates the exact locus; V stands for version, n for annotation version and m for locus version. Previous nomenclatures and equivalences can be found elsewhere [108]. In Table 2, loci are named as the original publication for traceability.

3.2.4 Homologous recombination

P. patens possesses an extremely high capacity of homologous recombination, which allows researchers to alter moss genomic DNA in any desired endogenous locus [14]. A common workflow is gene deletion by replacement with an antibiotic cassette. Also, protein localization studies with endogenous expression level is easily achieved by fusion of the fluorescent gene reporter sequence right after the target gene. Some vector sets for knock-out and knock-in to edit moss genome have been established and can be requested from several research groups (see Table 2) [11, 109]. Due to the ancestral genome duplication events in moss evolution, there is high functional redundancy of several gene families that decrease the risk of unwanted ortholog disruption.

3.2.5 Targeted double strand break and directed repair (CRISPR/Cas9)

The game-changing CRISPR/Cas9 method has proven an efficient and effective tool in P. patens to achieve large deletions, localised knock-in and point mutations [110, 111]. Transient transformation, flexibility of selection strategy and easy cloning workflow has rendered CRISPR/Cas9 transformation an established tool for P. patens research. Two groups developed whole CRISPR/Cas9 platforms independently with high editing efficiencies.

In Nogué’s lab, a co-delivery method was developed where each element of the system (Cas9, sgRNA and selection cassettes) was present in a separate plasmid. In this method, Cas9 expression is driven by an actin promoter and ready to use as is. The selection strategy can be chosen freely due to the lack of integration, minding the presence of resistance in the background lines. This method is also suitable for multiple mutations in different genes at once, given that more than one sgRNA plasmid can be simultaneously delivered in one transformation with still sufficiently high efficiencies of transformation [110].

In Bezanilla’s lab, a whole set of gateway destination vectors for CRISPR/Cas9 system was developed. The strategy was to design a vector set to finally put all the three essential components (Cas9, sgRNA and selection cassettes) in a single expression plasmid. In both protocols, the sgRNA can be designed and optimised using the online design tool CRISPOR V1 against P. patens genome Phytozome V11 [112].

To increase the accuracy of mutations, the CRISPR/Cas9 system is applied with a homology-directed repair (HDR), which allows for seamless knock-in or point mutation in desired sites. The template DNA can be a donor plasmid that harbours homologous fragments or oligodeoxynucleotides (ODNs) [111, 113]. By co-transforming the plasmid or ODNs together with CRISPR/Cas9 and sgRNA vectors, both methods show high accuracy to generate a desired point mutation or scarless insertion with a fluorescence tag at any suitable location of the gene.

3.2.6 RNA interference

Given that moss possesses a relatively big gene family, arguably due to its double genome duplication, the employment of gene deletion strategies might be inefficient to investigate gene functions due to the high redundancy rate (e.g. there are four ROP genes with highly homologous or identical sequences) [38]. For this, RNA interference (RNAi) strategies offer an alternative to overcome this problem, and the procedure has been well-established.

To generate an RNAi construct for a gene of interest, a DNA sequence of 300 to 1000 bp is subcloned in a destination vector. After standard PEG-mediated transformation, the silencing effect can be detected after 24 h and last up to 3 weeks [104]. To avoid lethal effects when constitutively expressing interfering RNA, an oestradiol-inducible RNAi system is available [102]. Coupling RNAi silencing activation with fluorescent reporters facilitates screenings of loss of function phenotypes [104].

3.3 Transformation

Standard transformation protocols have been applicable to P. patens for a long time. In Table 3, three methods are shown, with key protocol references for their experimental application. The most essential step after transformation is phenotypic characterisation, that is often performed at the colony level (as it is derived from the microscopic phenotypes). Some phenotypes that must be compared with the reference wild type ecotype include colony size, shape, colour, texture, gametophore count and ratio of gametophore number to colony surface. In the microscopic level, protonemal parameters such as chloronema and caulonema cell length, thickness, growth rate and transition and ratio of one to the other are valuable indicators of several hormone and developmental processes. Naturally, the phenotyping should include characteristics associated to the process of study.

StrategyHighlightsReference
PEG/MannitolOne to two round selection, 10% of transformants, 3–4 weeks for stable transformants[114, 115, 116, 117]
Agrobacterium tumefaciensFour-round selection, 100% positive transformants, 12–16 weeks[118]
Particle bombardmentEasy to conduct; less used.
Transient and stable DNA integration.
[119]

Table 3.

Summary table of the classical and current transformation techniques and reference protocols for application.

Advertisement

4. Beyond P. patens

Due to their accessibility, tractability and close yet independent phylogenetic position, the interest in Bryophytes has increased dramatically in the last decade [1]. Beyond P. patens, mosses have garnered special interest for their physiology and development, involvement in carbon sequestration, abiotic stresses management and biotic interactions. Eight species have had their nuclear genome sequenced and drafted in the last few years, and at least thirteen others are currently being sequenced [120, 121, 122, 123, 124, 125, 126, 127]. Furthermore, the mitochondrial and/or plastid genomes of more than forty other moss species (not cited) has been published in the last six years, which may precede their nuclear genome study as with P. patens. This level of knowledge is an essential tool to dissect the molecular basis of processes under study, and the recent and future increase in the availability of this information is going to dramatically accelerate research in mosses, among other bryophytes [128]. In this section, a collection of mosses at the frontier of moss cell and molecular research are highlighted for their ecological relevance, distinct physiology and genetic composition, among others. We believe these will be the next generation of mosses for research that will provide new insights in plant research beyond P. patens in the coming years.

4.1 Ceratodon purpureus (Hedw.) Brid.

The fire moss C. purpureus (Dicranales, Bryopsida) is a cosmopolitan species that thrives in diverse ecosystems, including hostile post-wildfire or heavy metal-contaminated areas, and those with high radiation and freezing temperatures [129, 130]. The life cycle of this moss involves male and female haploid individuals due to the presence of sexual chromosomes. Consequently, it has become a reference for dioecious reproduction and sexual dimorphism, with some developmental differences in sexual and non-sexual features [131, 132]. In 2021, male and female nuclear annotated genomes were published, making Ceratodon the third sequenced moss genus [127]. Furthermore, gene targeting in this species has been proven effective, providing all the basic tools for cell and molecular biology research [133]. Importantly, the similarity of growth fashion between C. purpureus and P. patens will provide a new reference to study the discussed developmental processes in higher depth.

4.2 Hypnales W.R. Buck & Vitt

The Hypnales (Bryopsida) are the biggest and most diverse order of mosses with varied morphology, and mostly exhibit pleurocarpous (i.e. non-erect) growth fashion. It includes Fontinalis antipyretica Hedw., Pleurozium schreberi (Brid.) Mitt. and Calohypnum plumiforme (Wilson) Kučera & Ignatov, the genome sequences of which have been published in the last two years [123, 125, 134]. P. schreberi is attractive due to its documented symbiotic relationships with N2-fixing cyanobacteria and C. plumiforme for bryophyte-exclusive biosynthetic gene clusters research [123, 134]. As pleurocarps, all of them exhibit a non-erect plant architecture that suggests an adapted regulation of stem development.

Remarkably, the common aquatic moss F. antipyretica is a globally distributed species and serves as a reference organism for the study of land-to-water habitat reversal and its genetic basis. From the developmental processes’ perspective, this moss has a distinct interest due to its tristichous phyllotactic pattern (120° rotation from organ to organ) that can serve as a reference in the investigation of genetic regulation of asymmetric cell divisions in the shoot apical cell at the gametophore apex (see Figure 1) [24].

4.3 Sphagnum L.

The genus Sphagnum (Sphagnales, Sphagnopsida) plays an important ecological role in the climate change situation, as its species are important carbon fixators. For this reason, The Sphagnome Project was created in 2018 in the aim to sequence fifteen species across the genus [124]. At this moment, the genomes of Sphagnum fallax and Sphagnum magellanicum have been published [124]. From the developmental point of view, Sphagnum spp. are attractive due to their branching gametophores and subsequent implications in lateral shoot meristems and phyllotaxis, which is different from P. patens and C. purpureus. Furthermore, Sphagnum spp. do not show rhizoids and some species are mostly or fully aquatic, serving as models for land-to-water reversal. Remarkably, Sphagnum spp. do not develop filamentous protonemata as most mosses, but thalloid protonemata (i.e. disk-like, bidimensional), as that of liverworts [135].

4.4 Polytrichopsida Doweld

The moss class Polytrichopsida is the second biggest (~200 species) after Bryopsida (~11500 species), and its species have unique morphological characteristics that make them uniquely interesting for developmental biology [5]. Despite mosses being regarded as avascular plants, some exhibit hydroids and leptoids, a functionally analogous tissue to tracheids and sieve elements of vascular plants that slightly differs morphologically and developmentally. Polytrichopsida has several genera with such structures, in some cases underdeveloped and, in others, completely functional, like in Polytrichum and Dawsonia, with up to 65-cm tall gametophores [5, 136]. This distinct characteristic suggests that xylem-like structures evolved independently and thus the genetic and molecular machinery necessary to its development may have similar origins to that of fern, gymnosperm, and angiosperm vasculature. Another attractive feature of several genera of Polytrichaceae is the perpendicular lamellae on the unistratose phyllids, that represents an increase in leaf complexity and has proven to be an alternative evolutionary path for increased photosynthetic capacity.

Despite its potential to provide valuable insight, genetic tools have barely been developed for this taxon, but full mitochondrial and plastid genomes have recently been published for Polytrichum commune, a cosmopolitan species ~10 cm long that has a complete stem vasculature [137, 138].

4.5 Other mosses

The desert moss Syntrichia caninervis Mitt. (Pottiales, Bryopsida) is the first outstanding example of dessication-tolerant moss to have its genome sequenced [126]. This genome will provide tools to dissect the development of the unique sub-micron structures of its phyllids that stimulate water capture and what genes are involved in the asymmetric growth and divisions necessary for this structure [139].

The heavy metal-tolerant moss Scopelophila cataratae (Mitt.) Broth. (Pottiales, Bryopsida), capable of thriving in copper-rich environments, has had its genome drafted (unpublished, 2016) and CRISPR/Cas9 mutagenesis demonstrated [121].

Funaria hygrometrica Hedw. (Funariales, Bryopsida) is evolutionary close to P. patens, and they have virtually identical morphology at the gametophyte generation, but remarkably different sporophyte generation. F. hygrometrica has a longer seta, different mechanisms and regulation of spore release [140]. However, the level of difference in their transcriptome is unexpectedly high and transcripts seem to be shifted in expression time but not in sequence. The recently published genome may allow investigate how time-shifted expression of regulators impacts morphology [120].

Advertisement

5. Conclusions

We have shown how mosses are an increasingly relevant model group to plant developmental biology due to their distinct accessibility at physiological and molecular level. Their evolutionary distance with agronomically relevant plants does not diminish their potential to help to understand fundamental questions of development that remain unsolved, given that most essential regulatory networks are conserved. This has been shown in the hormonal regulation of branching and shoot and root development, regeneration, the establishment of shoot apical cells and the genetic make-up in 2D-to-3D transition. The utility of mosses has convinced the scientific community to the point of promoting the sequencing of eight new species to explore other physiological processes in the last few years. We have also shown how Physcomitrium patens is a workhorse in cell and molecular biology of plants, and provided evidence that it is likely to become a standard tool of plant developmental biology together with a number of other mosses.

Advertisement

Acknowledgments

The authors would like to thank Dr. Jeroen de Keijzer and Dr. Tijs Ketelaar for their thoughtful and detailed review of the manuscript. Also, the funding agencies Technology, Knowledge and Innovation, division Horticulture and Propagating Material (TKI T&U) and the Dutch Research Council (NWO) (reference number: TKILWV20.390) for funding JFC and the ERC grant to Prof. J. Friml (reference number: PR1023ERC02) for funding HT. The authors would like to sincerely apologise for the literature not cited that may be relevant for this chapter and is not present due to space constraints.

References

  1. 1. Cheng S, Xian W, Fu Y, Marin B, Keller J, Wu T, et al. Genomes of Subaerial Zygnematophyceae Provide Insights into Land Plant Evolution. Cell. 2019;179(5):1057-1067.e14
  2. 2. McDaniel SF. Bryophytes are not early diverging land plants. New Phytol. 2021;230(4):1300-4
  3. 3. Green TGA, Clayton-Greene KA. Studies on Dawsonia superba Grev. II. Growth rate. J Bryol. 1981;11(4):723-31
  4. 4. Medina NG, Draper I, Lara F. Biogeography of mosses and allies: does size matter? In: Fontaneto D, editor. Biogeography of Microscopic Organisms: Is Everything Small Everywhere? [Internet]. Cambridge: Cambridge University Press; 2011. p. 209-33. (Systematics Association Special Volume Series). Available from: https://www.cambridge.org/core/books/biogeography-of-microscopic-organisms/biogeography-of-mosses-and-allies-does-size-matter/11003DB9B3BC0FE5CF5F603BFB4AF4A1
  5. 5. Bell N, Kariyawasam I, Flores J, Hyvönen J. The diversity of the Polytrichopsida—a review. Bryophyt Divers Evol [Internet]. 2021 Jun 30 [cited 2021 Sep 7];043(1):98-111. Available from: https://www.biotaxa.org/dbe/article/view/bde.43.1.8
  6. 6. Ekwealor JTB, Fisher KM. Life under quartz: Hypolithic mosses in the Mojave Desert. PLoS One [Internet]. 2020 [cited 2021 Jul 28];15(7):e0235928. Available from: https://journals.plos.org/plosone/article?id=10.1371/journal.pone.0235928
  7. 7. Górski P, Gądek B, Gąbka M. Snow as a parameter of bryophyte niche partitioning in snow-beds of the Tatra Mountains (Western Carpathians). Ecol Indic. 2020 Jun 1;113:106258
  8. 8. Nakosteen PC, Hughes KW. Sexual Life Cycle of Three Species of Funariaceae in Culture. Bryologist [Internet]. 1978 Aug 31;81(2):307-14. Available from: https://www.jstor.org/stable/3242191
  9. 9. Reski R. Quantitative moss cell biology. Curr Opin Plant Biol. 2018 Dec 1;46:39-47
  10. 10. Xu B, Ohtani M, Yamaguchi M, Toyooka K, Wakazaki M, Sato M, et al. Contribution of NAC transcription factors to plant adaptation to land. Science (80-). 2014;343(6178):1505-8
  11. 11. Rensing SA, Goffinet B, Meyberg R, Wu S-ZZ, Bezanilla M. The moss physcomitrium (Physcomitrella) patens: A model organism for non-seed plants. Plant Cell [Internet]. 2020 May 1 [cited 2021 Jun 16];32(5):1361-76. Available from: www.plantcell.org/cgi/doi/10.1105/tpc.19.00828
  12. 12. de Keijzer J, Rios AF, Willemsen V. Physcomitrium patens: A single model to study oriented cell divisions in 1d to 3d patterning. Int J Mol Sci [Internet]. 2021 Mar 1 [cited 2021 Jun 16];22(5):1-16. Available from: https://doi.org/10.3390/ijms22052626
  13. 13. Perroud P-F, Meyberg R, Demko V, Quatrano RS, Olsen O-A, Rensing SA. DEK1 displays a strong subcellular polarity during Physcomitrella patens 3D growth. New Phytol [Internet]. 2020 May 1 [cited 2021 Aug 4];226(4):1029-41. Available from: https://nph-onlinelibrary-wiley-com.ezproxy.library.wur.nl/doi/full/10.1111/nph.16417
  14. 14. Prigge MJ, Bezanilla M. Evolutionary crossroads in developmental biology: Physcomitrella patens. Development. 2010 Nov 1;137(21):3535-43
  15. 15. Hoffmann B, Proust H, Belcram K, Labrune C, Boyer F-D, Rameau C, et al. Strigolactones Inhibit Caulonema Elongation and Cell Division in the Moss Physcomitrella patens. PLoS One [Internet]. 2014 Jun 9 [cited 2021 Sep 2];9(6):e99206. Available from: https://journals.plos.org/plosone/article?id=10.1371/journal.pone.0099206
  16. 16. Bascom CS, Wu S-Z, Nelson K, Oakey J, Bezanilla M. Long-Term Growth of Moss in Microfluidic Devices Enables Subcellular Studies in Development. Plant Physiol [Internet]. 2016 Sep 1 [cited 2021 Jun 17];172(1):28-37. Available from: https://academic.oup.com/plphys/article/172/1/28-37/6115611
  17. 17. Cove DJ, Knight CD, Lamparter T. Mosses as model systems. Trends Plant Sci [Internet]. 1997 Mar 1 [cited 2021 Sep 2];2(3):99-105. Available from: http://www.cell.com/article/S136013859610056X/fulltext
  18. 18. Jaeger R, Moody LA. A fundamental developmental transition in Physcomitrium patens is regulated by evolutionarily conserved mechanisms [Internet]. Vol. 23, Evolution and Development. Blackwell Publishing Inc.; 2021 [cited 2021 Jun 16]. p. 123-36. Available from: https://onlinelibrary.wiley.com/doi/full/10.1111/ede.12376
  19. 19. Thelander M, Olsson T, Ronne H. Effect of the energy supply on filamentous growth and development in Physcomitrella patens. J Exp Bot [Internet]. 2005 Feb 1 [cited 2021 Sep 2];56(412):653-62. Available from: https://academic.oup.com/jxb/article/56/412/653/580262
  20. 20. Moody LA. The 2D to 3D growth transition in the moss Physcomitrella patens. Curr Opin Plant Biol [Internet]. 2019 Feb 1 [cited 2021 Aug 5];47:88-95. Available from: https://doi.org/10.1016/j.pbi.2018.10.001
  21. 21. Tang H, Duijts K, Bezanilla M, Scheres B, Vermeer JEM, Willemsen V. Geometric cues forecast the switch from two- to three-dimensional growth in Physcomitrella patens. New Phytol. 2020;225(5):1945-55
  22. 22. Sakakibara K, Nishiyama T, Sumikawa N, Kofuji R, Murata T, Hasebe M. Involvement of auxin and a homeodomain-leucine zipper I gene in rhizoid development of the moss Physcomitrella patens. Development. 2003 Oct 15;130(20):4835-46
  23. 23. Jang G, Yi K, Pires ND, Menand B, Dolan L. RSL genes are sufficient for rhizoid system development in early diverging land plants. Development [Internet]. 2011 Jun 1 [cited 2021 Aug 31];138(11):2273-81. Available from: http://mrbayes.csit.fsu.edu/
  24. 24. Véron E, Vernoux T, Coudert Y. Phyllotaxis from a Single Apical Cell. Trends Plant Sci [Internet]. 2021 Feb 1 [cited 2021 Jun 16];26(2):124-31. Available from: https://doi.org/10.1016/j.tplants.2020.09.014
  25. 25. Hohe A, Rensing SA, Mildner M, Lang D, Reski R. Day Length and Temperature Strongly Influence Sexual Reproduction and Expression of a Novel MADS-Box Gene in the Moss Physcomitrella patens. Plant Biol [Internet]. 2002 Sep 1 [cited 2021 Aug 31];4(5):595-602. Available from: https://onlinelibrary.wiley.com/doi/full/10.1055/s-2002-35440
  26. 26. Hiss M, Meyberg R, Westermann J, Haas FB, Schneider L, Schallenberg-Rüdinger M, et al. Sexual reproduction, sporophyte development and molecular variation in the model moss Physcomitrella patens: introducing the ecotype Reute. Plant J [Internet]. 2017 May 1 [cited 2021 Jul 27];90(3):606-20. Available from: https://onlinelibrary.wiley.com/doi/full/10.1111/tpj.13501
  27. 27. Meyberg R, Perroud P-F, Haas FB, Schneider L, Heimerl T, Renzaglia KS, et al. Characterisation of evolutionarily conserved key players affecting eukaryotic flagellar motility and fertility using a moss model. New Phytol [Internet]. 2020 Jul 1 [cited 2021 Jul 28];227(2):440-54. Available from: https://nph.onlinelibrary.wiley.com/doi/full/10.1111/nph.16486
  28. 28. Cervantes-Pérez D, Ortega-García A, Medina-Andrés R, Batista-García RA, Lira-Ruan V. Exogenous Nitric Oxide Delays Plant Regeneration from Protoplast and Protonema Development in Physcomitrella patens. Plants 2020, Vol 9, Page 1380 [Internet]. 2020 Oct 16 [cited 2021 Jul 28];9(10):1380. Available from: https://www.mdpi.com/2223-7747/9/10/1380/htm
  29. 29. Ishikawa M, Murata T, Sato Y, Nishiyama T, Hiwatashi Y, Imai A, et al. Physcomitrella Cyclin-Dependent Kinase A Links Cell Cycle Reactivation to Other Cellular Changes during Reprogramming of Leaf Cells. Plant Cell [Internet]. 2011 Aug 1 [cited 2021 Jul 28];23(8):2924-38. Available from: https://academic.oup.com/plcell/article/23/8/2924/6097203
  30. 30. KNIGHT CD, COVE DJ. The polarity of gravitropism in the moss Physcomitrella patens is reversed during mitosis and after growth on a clinostat. Plant Cell Environ [Internet]. 1991 Dec 1 [cited 2021 Aug 31];14(9):995-1001. Available from: https://onlinelibrary.wiley.com/doi/full/10.1111/j.1365-3040.1991.tb00970.x
  31. 31. Jenkins GI, Cove DJ. Phototropism and polarotropism of primary chloronemata of the moss Physcomitrella patens: responses of mutant strains. Planta 1983 1595 [Internet]. 1983 Nov [cited 2021 Aug 31];159(5):432-8. Available from: https://link.springer.com/article/10.1007/BF00392079
  32. 32. Bao L, Yamamoto KT, Fujita T. Phototropism in gametophytic shoots of the moss Physcomitrella patens. Plant Signal Behav [Internet]. 2015 Apr 7 [cited 2021 Aug 31];10(3). Available from: https://www.tandfonline.com/doi/abs/10.1080/15592324.2015.1010900
  33. 33. Harrison CJ, Roeder AHK, Meyerowitz EM, Langdale JA. Local Cues and Asymmetric Cell Divisions Underpin Body Plan Transitions in the Moss Physcomitrella patens. Curr Biol. 2009 Mar 24;19(6):461-71
  34. 34. Moody LA, Kelly S, Rabbinowitsch E, Langdale JA. Genetic Regulation of the 2D to 3D Growth Transition in the Moss Physcomitrella patens. Curr Biol. 2018 Feb 5;28(3):473-478.e5
  35. 35. Perroud P-F, Demko V, Johansen W, Wilson RC, Olsen O-A, Quatrano RS. Defective Kernel 1 (DEK1) is required for three-dimensional growth in Physcomitrella patens. New Phytol [Internet]. 2014 Aug 1 [cited 2021 Jul 28];203(3):794-804. Available from: https://nph-onlinelibrary-wiley-com.ezproxy.library.wur.nl/doi/full/10.1111/nph.12844
  36. 36. Moody LA, Kelly S, Clayton R, Weeks Z, Emms DM, Langdale JA. NO GAMETOPHORES 2 Is a Novel Regulator of the 2D to 3D Growth Transition in the Moss Physcomitrella patens. Curr Biol. 2021 Feb 8;31(3):555-563.e4
  37. 37. Yi P, Goshima G. Rho of Plants GTPases and Cytoskeletal Elements Control Nuclear Positioning and Asymmetric Cell Division during Physcomitrella patens Branching. Curr Biol [Internet]. 2020;30(14):2860-2868.e3. Available from: https://doi.org/10.1016/j.cub.2020.05.022
  38. 38. Cheng X, Mwaura BW, Chang Stauffer SR, Bezanilla M. A Fully Functional ROP Fluorescent Fusion Protein Reveals Roles for This GTPase in Subcellular and Tissue-Level Patterning. Plant Cell [Internet]. 2020 Nov 2 [cited 2021 Aug 11];32(11):3436-51. Available from: https://academic.oup.com/plcell/article/32/11/3436/6099412
  39. 39. Kimata Y, Higaki T, Kawashima T, Kurihara D, Sato Y, Yamada T, et al. Cytoskeleton dynamics control the first asymmetric cell division in Arabidopsis zygote. Proc Natl Acad Sci [Internet]. 2016 Dec 6 [cited 2021 Sep 1];113(49):14157-62. Available from: https://www.pnas.org/content/113/49/14157
  40. 40. Sussex IM, Kerk NM. The evolution of plant architecture. Curr Opin Plant Biol. 2001 Feb 1;4(1):33-7
  41. 41. Shi B, Vernoux T. Patterning at the shoot apical meristem and phyllotaxis. Curr Top Dev Biol. 2019 Jan 1;131:81-107
  42. 42. Barton MK. Twenty years on: The inner workings of the shoot apical meristem, a developmental dynamo. Dev Biol. 2010 May 1;341(1):95-113
  43. 43. Harrison CJ. Development and genetics in the evolution of land plant body plans. Philos Trans R Soc B Biol Sci [Internet]. 2017 Feb 5 [cited 2021 Aug 23];372(1713). Available from: https://royalsocietypublishing.org/doi/abs/10.1098/rstb.2015.0490
  44. 44. Hata Y, Kyozuka J. Fundamental mechanisms of the stem cell regulation in land plants: lesson from shoot apical cells in bryophytes. Plant Mol Biol 2021 [Internet]. 2021 Feb 20 [cited 2021 Aug 23];1:1-13. Available from: https://link.springer.com/article/10.1007/s11103-021-01126-y
  45. 45. Haig D. Homologous Versus Antithetic Alternation of Generations and the Origin of Sporophytes. Bot Rev 2008 743 [Internet]. 2008 Jul 26 [cited 2021 Aug 23];74(3):395-418. Available from: https://link-springer-com.ezproxy.library.wur.nl/article/10.1007/s12229-008-9012-x
  46. 46. Bennici A. Origin and early evolution of land plants. http://www-tandfonline-com.ezproxy.library.wur.nl/action/authorSubmission?journalCode=kcib20&page=instructions [Internet]. 2008 Oct [cited 2021 Sep 12];1(2):212-8. Available from: https://www-tandfonline-com.ezproxy.library.wur.nl/doi/abs/10.4161/cib.1.2.6987
  47. 47. Rensing SA, Lang D, Zimmer AD, Terry A, Salamov A, Shapiro H, et al. The Physcomitrella Genome Reveals Evolutionary Insights into the Conquest of Land by Plants. Science (80- ) [Internet]. 2008 Jan 4 [cited 2021 Aug 23];319(5859):64-9. Available from: https://science-sciencemag-org.ezproxy.library.wur.nl/content/319/5859/64
  48. 48. Frank MH, Scanlon MJ. Cell-specific transcriptomic analyses of three-dimensional shoot development in the moss Physcomitrella patens. Plant J [Internet]. 2015 Aug 1 [cited 2021 Aug 23];83(4):743-51. Available from: https://onlinelibrary.wiley.com/doi/full/10.1111/tpj.12928
  49. 49. Hofmeister W. Allgemeine Morphologie der Gewächse. Leipzig; 1868. 259 p
  50. 50. Givnish TJ. Ecological constraints on the evolution of plasticity in plants. Evol Ecol 2002 163 [Internet]. 2002 [cited 2021 Sep 15];16(3):213-42. Available from: https://link.springer.com/article/10.1023/A:1019676410041
  51. 51. Bravais L, Bravais A. Essai sur la disposition des feuilles curvisériées. Annales des sciences naturelles (Botanique); 1837. 69 p
  52. 52. Gola EM, Banasiak A. Diversity of phyllotaxis in land plants in reference to the shoot apical meristem structure. Acta Soc Bot Pol [Internet]. 2016 Dec 31 [cited 2021 Aug 24];85(4). Available from: https://pbsociety.org.pl/journals/index.php/asbp/article/view/6873
  53. 53. Jean R V. Phyllotaxis: A Systemic Study in Plant Morphogenesis [Internet]. Cambridge: Cambridge University Press; 1994. 401 p. Available from: https://www.cambridge.org/core/books/phyllotaxis/272D9010BE175D26B61D5A2ED8D87A3C
  54. 54. Moody LA. Three-dimensional growth: a developmental innovation that facilitated plant terrestrialization. J Plant Res 2020 1333 [Internet]. 2020 Feb 24 [cited 2021 Aug 4];133(3):283-90. Available from: https://link-springer-com.ezproxy.library.wur.nl/article/10.1007/s10265-020-01173-4
  55. 55. Kamamoto N, Tano T, Fujimoto K, Shimamura M. Rotation angle of stem cell division plane controls spiral phyllotaxis in mosses. J Plant Res [Internet]. 2021 May 1 [cited 2021 Jun 16];134(3):457-73. Available from: https://doi.org/10.1007/s10265-021-01298-0
  56. 56. Zagórska-Marek B, Sokołowska K, Turzańska M. Chiral events in developing gametophores of Physcomitrella patens and other moss species are driven by an unknown, universal direction-sensing mechanism. Am J Bot [Internet]. 2018 Dec 1 [cited 2021 Aug 24];105(12):1986-94. Available from: https://bsapubs.onlinelibrary.wiley.com/doi/full/10.1002/ajb2.1200
  57. 57. Xu L. De novo root regeneration from leaf explants: wounding, auxin, and cell fate transition. Curr Opin Plant Biol. 2018 Feb 1;41:39-45
  58. 58. Mathew MM, Prasad K. Model systems for regeneration: Arabidopsis. Development [Internet]. 2021 Mar 15 [cited 2021 Sep 2];148(6). Available from: https://dev.biologists.org/collection/
  59. 59. Sato Y, Sugimoto N, Hirai T, Imai A, Kubo M, Hiwatashi Y, et al. Cells reprogramming to stem cells inhibit the reprogramming of adjacent cells in the moss Physcomitrella patens. Sci Reports 2017 71 [Internet]. 2017 May 15 [cited 2021 Jul 28];7(1):1-12. Available from: https://www.nature.com/articles/s41598-017-01786-1
  60. 60. Ikeuchi M, Ogawa Y, Iwase A, Sugimoto K. Plant regeneration: Cellular origins and molecular mechanisms. Dev. 2016 May 1;143(9):1442-51
  61. 61. Kofuji R, Hasebe M. Eight types of stem cells in the life cycle of the moss Physcomitrella patens. Curr Opin Plant Biol. 2014 Feb 1;17(1):13-21
  62. 62. Ikeuchi M, Favero DS, Sakamoto Y, Iwase A, Coleman D, Rymen B, et al. Molecular Mechanisms of Plant Regeneration. Annu Rev Plant Biol [Internet]. 2019 Apr 29 [cited 2021 Sep 2];70(1):377-406. Available from: https://www.annualreviews.org/doi/10.1146/annurev-arplant-050718-100434
  63. 63. Kubo M, Nishiyama T, Tamada Y, Sano R, Ishikawa M, Murata T, et al. Single-cell transcriptome analysis of Physcomitrella leaf cells during reprogramming using microcapillary manipulation. Nucleic Acids Res [Internet]. 2019 May 21 [cited 2021 Aug 24];47(9):4539-53. Available from: https://academic.oup.com/nar/article/47/9/4539/5381068
  64. 64. Nishiyama T, Miyawaki K, Ohshima M, Thompson K, Nagashima A, Hasebe M, et al. Digital Gene Expression Profiling by 5′-End Sequencing of cDNAs during Reprogramming in the Moss Physcomitrella patens. PLoS One [Internet]. 2012 May 4 [cited 2021 Aug 24];7(5):e36471. Available from: https://journals.plos.org/plosone/article?id=10.1371/journal.pone.0036471
  65. 65. Zhou W, Lozano-Torres JL, Blilou I, Zhang X, Zhai Q , Smant G, et al. A Jasmonate Signaling Network Activates Root Stem Cells and Promotes Regeneration. Cell. 2019 May 2;177(4):942-956.e14
  66. 66. Kareem A, Durgaprasad K, Sugimoto K, Du Y, Pulianmackal AJ, Trivedi ZB, et al. PLETHORA genes control regeneration by a two-step mechanism. Curr Biol [Internet]. 2015;25(8):1017-30. Available from: http://dx.doi.org/10.1016/j.cub.2015.02.022
  67. 67. Subban P, Kutsher Y, Evenor D, Belausov E, Zemach H, Faigenboim A, et al. Shoot Regeneration Is Not a Single Cell Event. Plants 2021, Vol 10, Page 58 [Internet]. 2020 Dec 29 [cited 2021 Sep 2];10(1):58. Available from: https://www.mdpi.com/2223-7747/10/1/58/htm
  68. 68. Sakakibara K, Reisewitz P, Aoyama T, Friedrich T, Ando S, Sato Y, et al. WOX13-like genes are required for reprogramming of leaf and protoplast cells into stem cells in the moss Physcomitrella patens. Development. 2014 Apr 15;141(8):1660-70
  69. 69. Li C, Sako Y, Imai A, Nishiyama T, Thompson K, Kubo M, et al. A Lin28 homologue reprograms differentiated cells to stem cells in the moss Physcomitrella patens. Nat Commun 2017 81 [Internet]. 2017 Jan 27 [cited 2021 Aug 24];8(1):1-13. Available from: https://www.nature.com/articles/ncomms14242
  70. 70. Ishikawa M, Morishita M, Higuchi Y, Ichikawa S, Ishikawa T, Nishiyama T, et al. Physcomitrella STEMIN transcription factor induces stem cell formation with epigenetic reprogramming. Nat Plants 2019 57 [Internet]. 2019 Jul 8 [cited 2021 Aug 24];5(7):681-90. Available from: https://www.nature.com/articles/s41477-019-0464-2
  71. 71. Storti M, Costa A, Golin S, Zottini M, Morosinotto T, Alboresi A. Systemic Calcium Wave Propagation in Physcomitrella patens. Plant Cell Physiol [Internet]. 2018 Jul 1 [cited 2021 Aug 24];59(7):1377-84. Available from: https://academic.oup.com/pcp/article/59/7/1377/5033790
  72. 72. Kleist TJ, Cartwright HN, Perera AM, Christianson ML, Lemaux PG, Luan S. Genetically encoded calcium indicators for fluorescence imaging in the moss Physcomitrella: GCaMP3 provides a bright new look. Plant Biotechnol J. 2017 Oct 1;15(10):1235-7
  73. 73. Nicolas WJ, Grison MS, Bayer EM. Shaping intercellular channels of plasmodesmata: the structure-to-function missing link. J Exp Bot [Internet]. 2018 Jan 1 [cited 2021 Aug 24];69(1):91-103. Available from: https://academic.oup.com/jxb/article/69/1/91/4107278
  74. 74. Kitagawa M, Fujita T. Quantitative imaging of directional transport through plasmodesmata in moss protonemata via single-cell photoconversion of Dendra2. J Plant Res 2013 1264 [Internet]. 2013 Feb 5 [cited 2021 Aug 24];126(4):577-85. Available from: https://link.springer.com/article/10.1007/s10265-013-0547-5
  75. 75. Kitagawa M, Tomoi T, Fukushima T, Sakata Y, Sato M, Toyooka K, et al. Abscisic Acid Acts as a Regulator of Molecular Trafficking through Plasmodesmata in the Moss Physcomitrella patens. Plant Cell Physiol [Internet]. 2019 Apr 1 [cited 2021 Aug 24];60(4):738-51. Available from: https://academic.oup.com/pcp/article/60/4/738/5267838
  76. 76. Reinhardt D, Mandel T, Kuhlemeier C. Auxin Regulates the Initiation and Radial Position of Plant Lateral Organs. Plant Cell. 2000 Apr;12(4):507
  77. 77. Vanneste S, Friml J. Auxin: A Trigger for Change in Plant Development. Cell. 2009 Mar 20;136(6):1005-16
  78. 78. MJ P, M L, NW A, M E. Physcomitrella patens auxin-resistant mutants affect conserved elements of an auxin-signaling pathway. Curr Biol [Internet]. 2010 Nov 9 [cited 2021 Sep 2];20(21):1907-12. Available from: https://pubmed.ncbi.nlm.nih.gov/20951049/
  79. 79. Paponov IA, Teale W, Lang D, Paponov M, Reski R, Rensing SA, et al. The evolution of nuclear auxin signalling. BMC Evol Biol 2009 91 [Internet]. 2009 Jun 3 [cited 2021 Sep 2];9(1):1-16. Available from: https://bmcecolevol.biomedcentral.com/articles/10.1186/1471-2148-9-126
  80. 80. Lavy M, Prigge MJ, Tao S, Shain S, Kuo A, Kirchsteiger K, et al. Constitutive auxin response in Physcomitrella reveals complex interactions between Aux/IAA and ARF proteins. Elife. 2016 Jun 1;5(JUN2016)
  81. 81. TA B, MM L, T A, NM B, M B, Y C, et al. Plasma membrane-targeted PIN proteins drive shoot development in a moss. Curr Biol [Internet]. 2014 Dec 1 [cited 2021 Sep 2];24(23):2776-85. Available from: https://pubmed.ncbi.nlm.nih.gov/25448003/
  82. 82. Viaene T, Landberg K, Thelander M, Medvecka E, Pederson E, Feraru E, et al. Directional Auxin Transport Mechanisms in Early Diverging Land Plants. Curr Biol. 2014 Dec 1;24(23):2786-91
  83. 83. Coudert Y, Palubicki W, Ljung K, Novak O, Leyser O, Harrison CJ. Three ancient hormonal cues co-ordinate shoot branching in a moss. Elife. 2015;4
  84. 84. Zürcher E, Tavor-Deslex D, Lituiev D, Enkerli K, Tarr PT, Müller B. A Robust and Sensitive Synthetic Sensor to Monitor the Transcriptional Output of the Cytokinin Signaling Network in Planta. Plant Physiol [Internet]. 2013 Feb 28 [cited 2021 Sep 15];161(3):1066-75. Available from: https://academic.oup.com/plphys/article/161/3/1066/6110587
  85. 85. Ashton NW, Grimsley NH, Cove DJ. Analysis of Gametophytic Development in the Moss, Physcomitrella patens, Using Auxin and Cytokinin Resistant Mutants. Planta. 1933;435(5):1-46
  86. 86. Hyoung S, Cho SH, Chung JH, So WM, Cui MH, Shin JS. Cytokinin oxidase PpCKX1 plays regulatory roles in development and enhances dehydration and salt tolerance in Physcomitrella patens. Plant Cell Reports 2019 393 [Internet]. 2019 Dec 20 [cited 2021 Sep 15];39(3):419-30. Available from: https://link.springer.com/article/10.1007/s00299-019-02500-3
  87. 87. Proust H, Hoffmann B, Xie X, Yoneyama K, Schaefer DG, Yoneyama K, et al. Strigolactones regulate protonema branching and act as a quorum sensing-like signal in the moss Physcomitrella patens. Development [Internet]. 2011 Apr 15 [cited 2021 Sep 16];138(8):1531-9. Available from: http://rsb.info.nih.gov/ij/
  88. 88. Furt F, Lemoi K, Tüzel E, Vidali L. Quantitative analysis of organelle distribution and dynamics in Physcomitrella patens protonemal cells. BMC Plant Biol 2012 121 [Internet]. 2012 May 17 [cited 2021 Aug 24];12(1):1-15. Available from: https://bmcplantbiol.biomedcentral.com/articles/10.1186/1471-2229-12-70
  89. 89. Vidali L, Gisbergen PAC van, Guérin C, Franco P, Li M, Burkart GM, et al. Rapid formin-mediated actin-filament elongation is essential for polarized plant cell growth. Proc Natl Acad Sci [Internet]. 2009 Aug 11 [cited 2021 Aug 24];106(32):13341-6. Available from: https://www.pnas.org/content/106/32/13341
  90. 90. Hiwatashi Y, Obara M, Sato Y, Fujita T, Murata T, Hasebe M. Kinesins Are Indispensable for Interdigitation of Phragmoplast Microtubules in the Moss Physcomitrella patens. Plant Cell [Internet]. 2008 Dec 31 [cited 2021 Sep 2];20(11):3094-106. Available from: https://academic.oup.com/plcell/article/20/11/3094/6092527
  91. 91. Kosetsu K, de Keijzer J, Janson ME, Goshima G. MICROTUBULE-ASSOCIATED PROTEIN65 Is Essential for Maintenance of Phragmoplast Bipolarity and Formation of the Cell Plate in Physcomitrella patens. Plant Cell [Internet]. 2013 Dec 30 [cited 2021 Aug 24];25(11):4479-92. Available from: https://academic.oup.com/plcell/article/25/11/4479/6096775
  92. 92. Thelander M, Landberg K, Sundberg E. Minimal auxin sensing levels in vegetative moss stem cells revealed by a ratiometric reporter. New Phytol. 2019;224(2):775-88
  93. 93. Kozgunova E, Goshima G. A versatile microfluidic device for highly inclined thin illumination microscopy in the moss Physcomitrella patens. Sci Rep [Internet]. 2019 Dec 1 [cited 2021 Jun 17];9(1):1-8. Available from: https://doi.org/10.1038/s41598-019-51624-9
  94. 94. Leong SY, Edzuka T, Goshima G, Yamada M. Kinesin-13 and Kinesin-8 Function during Cell Growth and Division in the Moss Physcomitrella patens. Plant Cell [Internet]. 2020 Mar 2 [cited 2021 Aug 3];32(3):683-702. Available from: https://academic-oup-com.ezproxy.library.wur.nl/plcell/article/32/3/683/6099160
  95. 95. Sakai K, Charlot F, Saux T Le, Bonhomme S, Nogué F, Palauqui JC, et al. Design of a comprehensive microfluidic and microscopic toolbox for the ultra-wide spatio-temporal study of plant protoplasts development and physiology. Plant Methods [Internet]. 2019 Dec 24 [cited 2021 Jun 17];15(1):1-12. Available from: https://plantmethods.biomedcentral.com/articles/10.1186/s13007-019-0459-z
  96. 96. Busch W, Moore BT, Martsberger B, Mace DL, Twigg RW, Jung J, et al. A microfluidic device and computational platform for high-throughput live imaging of gene expression. Nat Methods 2012 911 [Internet]. 2012 Sep 30 [cited 2021 Aug 4];9(11):1101-6. Available from: https://www.nature.com/articles/nmeth.2185
  97. 97. Horowitz LF, Rodriguez AD, Ray T, Folch A. Microfluidics for interrogating live intact tissues [Internet]. Vol. 6, Microsystems and Nanoengineering. Springer Nature; 2020 [cited 2021 Jun 17]. p. 1-27. Available from: www.nature.com/micronano
  98. 98. Cove DJ, Schild A, Ashton NW, Hartmann E. GENETIC AND PHYSIOLOGICAL STUDIES OF THE EFFECT OF LIGHT ON THE DEVELOPMENT OF THE MOSS, PHYSCOMITRELLA PATENS. Photochem Photobiol [Internet]. 1978 Feb 1 [cited 2021 Aug 24];27(2):249-54. Available from: https://onlinelibrary.wiley.com/doi/full/10.1111/j.1751-1097.1978.tb07596.x
  99. 99. Kamisugi Y, Stackelberg M Von, Lang D, Care M, Reski R, Rensing SA, et al. A sequence-anchored genetic linkage map for the moss, Physcomitrella patens. Plant J [Internet]. 2008 Dec 1 [cited 2021 Sep 1];56(5):855-66. Available from: https://onlinelibrary.wiley.com/doi/full/10.1111/j.1365-313X.2008.03637.x
  100. 100. Ding X, Pervere LM, Jr. CB, Bibeau JP, Khurana S, Butt AM, et al. Conditional genetic screen in Physcomitrella patens reveals a novel microtubule depolymerizing-end-tracking protein. PLOS Genet [Internet]. 2018 May 1 [cited 2021 Aug 24];14(5):e1007221. Available from: https://journals.plos.org/plosgenetics/article?id=10.1371/journal.pgen.1007221
  101. 101. Mohanasundaram B, Rajmane VB, Jogdand S V., Bhide AJ, Banerjee AK. Agrobacterium-mediated Tnt1 mutagenesis of moss protonemal filaments and generation of stable mutants with impaired gametophyte. Mol Genet Genomics 2019 2943 [Internet]. 2019 Jan 28 [cited 2021 Aug 24];294(3):583-96. Available from: https://link.springer.com/article/10.1007/s00438-019-01532-4
  102. 102. Nakaoka Y, Miki T, Fujioka R, Uehara R, Tomioka A, Obuse C, et al. An Inducible RNA Interference System in Physcomitrella patens Reveals a Dominant Role of Augmin in Phragmoplast Microtubule Generation. Plant Cell [Internet]. 2012 Aug 10 [cited 2021 Aug 24];24(4):1478-93. Available from: https://academic.oup.com/plcell/article/24/4/1478/6102364
  103. 103. Kubo M, Imai A, Nishiyama T, Ishikawa M, Sato Y, Kurata T, et al. System for Stable β-Estradiol-Inducible Gene Expression in the Moss Physcomitrella patens. PLoS One [Internet]. 2013 Sep 27 [cited 2021 Aug 24];8(9):e77356. Available from: https://journals.plos.org/plosone/article?id=10.1371/journal.pone.0077356
  104. 104. Bezanilla M, Perroud P-F, Pan A, Klueh P, Quatrano RS. An RNAi System in Physcomitrella patens with an Internal Marker for Silencing Allows for Rapid Identification of Loss of Function Phenotypes. Plant Biol [Internet]. 2005 Apr 15 [cited 2021 Aug 24];7(03):251-7. Available from: http://www.thieme-connect.com/products/ejournals/html/10.1055/s-2005-837597
  105. 105. Bezanilla M. The Bezanilla Lab Moss Methods [Internet]. 2012 [cited 2021 Sep 2]. Available from: https://sites.dartmouth.edu/bezanillalab/moss-methods/
  106. 106. Aoyama T, Hiwatashi Y, Shigyo M, Kofuji R, Kubo M, Ito M, et al. AP2-type transcription factors determine stem cell identity in the moss Physcomitrella patens. Development [Internet]. 2012 Sep 1 [cited 2021 Aug 20];139(17):3120-9. Available from: http://genome.jgi-psf.org/Phypa11/
  107. 107. Ulfstedt M, Hu G-Z, Johansson M, Ronne H. Testing of Auxotrophic Selection Markers for Use in the Moss Physcomitrella Provides New Insights into the Mechanisms of Targeted Recombination. Front Plant Sci. 2017 Nov 3;0:1850
  108. 108. Lang D, Ullrich KK, Murat F, Fuchs J, Jenkins J, Haas FB, et al. The Physcomitrella patens chromosome-scale assembly reveals moss genome structure and evolution. Plant J [Internet]. 2018 Feb 1 [cited 2021 Aug 20];93(3):515-33. Available from: https://onlinelibrary.wiley.com/doi/full/10.1111/tpj.13801
  109. 109. de Keijzer J, Kieft H, Ketelaar T, Goshima G, Janson ME. Shortening of Microtubule Overlap Regions Defines Membrane Delivery Sites during Plant Cytokinesis. Curr Biol. 2017 Feb 20;27(4):514-20
  110. 110. Lopez-Obando M, Hoffmann B, Géry C, Guyon-Debast A, Téoulé E, Rameau C, et al. Simple and Efficient Targeting of Multiple Genes Through CRISPR-Cas9 in Physcomitrella patens. G3 Genes|Genomes|Genetics [Internet]. 2016 Nov 1 [cited 2021 Aug 24];6(11):3647-53. Available from: https://academic.oup.com/g3journal/article/6/11/3647/6031123
  111. 111. Mallett DR, Chang M, Cheng X, Bezanilla M. Efficient and modular CRISPR-Cas9 vector system for Physcomitrella patens. Plant Direct [Internet]. 2019 Sep 1 [cited 2021 Aug 24];3(9):e00168. Available from: https://onlinelibrary.wiley.com/doi/full/10.1002/pld3.168
  112. 112. Concordet J-P, Haeussler M. CRISPOR: intuitive guide selection for CRISPR/Cas9 genome editing experiments and screens. Nucleic Acids Res [Internet]. 2018 Jul 2 [cited 2021 Sep 2];46(W1):W242-5. Available from: https://academic.oup.com/nar/article/46/W1/W242/4995687
  113. 113. Yi P, Goshima G. Transient cotransformation of CRISPR/Cas9 and oligonucleotide templates enables efficient editing of target loci in Physcomitrella patens. Plant Biotechnol J [Internet]. 2020 Mar 1 [cited 2021 Aug 24];18(3):599-601. Available from: https://onlinelibrary.wiley.com/doi/full/10.1111/pbi.13238
  114. 114. Vives C, Charlot F, Mhiri C, Contreras B, Daniel J, Epert A, et al. Highly efficient gene tagging in the bryophyte Physcomitrella patens using the tobacco (Nicotiana tabacum) Tnt1 retrotransposon. New Phytol [Internet]. 2016 Nov 1 [cited 2021 Aug 24];212(3):759-69. Available from: https://nph.onlinelibrary.wiley.com/doi/full/10.1111/nph.14152
  115. 115. Roberts AW, Dimos CS, Budziszek MJ, Goss CA, Lai V. Knocking Out the Wall: Protocols for Gene Targeting in Physcomitrella patens BT - The Plant Cell Wall: Methods and Protocols. In: Popper ZA, editor. Totowa, NJ: Humana Press; 2011. p. 273-90. Available from: https://doi.org/10.1007/978-1-61779-008-9_19
  116. 116. Schaefer DG. Gene targeting in Physcomitrella patens. Curr Opin Plant Biol. 2001 Apr 1;4(2):143-50
  117. 117. Hohe A, Egener T, Lucht JM, Holtorf H, Reinhard C, Schween G, et al. An improved and highly standardised transformation procedure allows efficient production of single and multiple targeted gene-knockouts in a moss, Physcomitrella patens. Curr Genet 2003 446 [Internet]. 2003 Oct 29 [cited 2021 Aug 24];44(6):339-47. Available from: https://link.springer.com/article/10.1007/s00294-003-0458-4
  118. 118. Li LH, Yang J, Qiu HL, Liu YY. Genetic transformation of Physcomitrella patens mediated by Agrobacterium tumefaciens. African J Biotechnol. 2010;9(25):3719-25
  119. 119. Cho S-H, Chung Y-S, Cho S-K, Rim Y-W, Shin and J-S. Particle Bombardment Mediated Transformation and GFP Expression in the Moss Physcomitrella patens. Mol Cells [Internet]. 1999 [cited 2021 Aug 24];9(1):14-9. Available from: http://www.molcells.org/journal/view.html?spage=14&volume=9&number=1
  120. 120. Prihatna C, Chen R, Barbetti MJ, Barker SJ, Tuset SI, Demirer GS, et al. Phyllotaxis: A Matthew Effect in Auxin Action Dolf. Adv Phytonanotechnology [Internet]. 2019 Mar 1 [cited 2020 Jan 30];1(1):1-12. Available from: http://dx.doi.org/10.1038/s42003-020-0917-1
  121. 121. Nomura T, Sakurai T, Osakabe Y, Osakabe K, Sakakibara H. Efficient and Heritable Targeted Mutagenesis in Mosses Using the CRISPR/Cas9 System. Plant Cell Physiol [Internet]. 2016 Dec 1 [cited 2021 Sep 7];57(12):2600-10. Available from: https://academic.oup.com/pcp/article/57/12/2600/2629317
  122. 122. Kirbis A, Waller M, Ricca M, Bont Z, Neubauer A, Goffinet B, et al. Transcriptional Landscapes of Divergent Sporophyte Development in Two Mosses, Physcomitrium (Physcomitrella) patens and Funaria hygrometrica. Front Plant Sci. 2020 Jun 10;0:747
  123. 123. Mao L, Kawaide H, Higuchi T, Chen M, Miyamoto K, Hirata Y, et al. Genomic evidence for convergent evolution of gene clusters for momilactone biosynthesis in land plants. Proc Natl Acad Sci [Internet]. 2020 Jun 2 [cited 2021 Sep 7];117(22):12472-80. Available from: https://www-pnas-org.ezproxy.library.wur.nl/content/117/22/12472
  124. 124. Weston DJ, Turetsky MR, Johnson MG, Granath G, Lindo Z, Belyea LR, et al. The Sphagnome Project: enabling ecological and evolutionary insights through a genus-level sequencing project. New Phytol [Internet]. 2018 Jan 1 [cited 2021 Sep 7];217(1):16-25. Available from: https://nph.onlinelibrary.wiley.com/doi/full/10.1111/nph.14860
  125. 125. Yu J, Li L, Wang S, Dong S, Chen Z, Patel N, et al. Draft genome of the aquatic moss Fontinalis antipyretica (Fontinalaceae, Bryophyta). Gigabyte [Internet]. 2020 Nov 16 [cited 2021 Sep 7];2020:1-9. Available from: https://gigabytejournal.com/articles/8
  126. 126. Silva AT, Gao B, Fisher KM, Mishler BD, Ekwealor JTB, Stark LR, et al. To dry perchance to live: Insights from the genome of the desiccation-tolerant biocrust moss Syntrichia caninervis. Plant J [Internet]. 2021 Mar 1 [cited 2021 Sep 7];105(5):1339-56. Available from: https://onlinelibrary-wiley-com.ezproxy.library.wur.nl/doi/full/10.1111/tpj.15116
  127. 127. Carey SB, Jenkins J, Lovell JT, Maumus F, Sreedasyam A, Payton AC, et al. Gene-rich UV sex chromosomes harbor conserved regulators of sexual development. Sci Adv. 2021 Jun 1;7(27)
  128. 128. Horn A, Pascal A, Lončarević I, Marques RV, Lu Y, Miguel S, et al. Natural Products from Bryophytes: From Basic Biology to Biotechnological Applications. https://doi-org.ezproxy.library.wur.nl/101080/0735268920211911034 [Internet]. 2021 [cited 2021 Sep 7];40(3):191-217. Available from: https://www-tandfonline-com.ezproxy.library.wur.nl/doi/abs/10.1080/07352689.2021.1911034
  129. 129. Campos ML, Prado GS, dos Santos VO, Nascimento LC, Dohms SM, da Cunha NB, et al. Mosses: Versatile plants for biotechnological applications. Vol. 41, Biotechnology Advances. Elsevier Inc.; 2020. p. 107533
  130. 130. Biersma EM, Convey P, Wyber R, Robinson SA, Dowton M, van de Vijver B, et al. Latitudinal Biogeographic Structuring in the Globally Distributed Moss Ceratodon purpureus. Front Plant Sci. 2020 Aug 28;0:1332
  131. 131. Kollar LM, Kiel S, James AJ, Carnley CT, Scola DN, Clark TN, et al. The genetic architecture of sexual dimorphism in the moss Ceratodon purpureus. Proc R Soc B [Internet]. 2021 Mar 10 [cited 2021 Sep 8];288(1946). Available from: https://royalsocietypublishing-org.ezproxy.library.wur.nl/doi/abs/10.1098/rspb.2020.2908
  132. 132. Slate ML, Rosenstiel TN, Eppley SM. Sex-specific morphological and physiological differences in the moss Ceratodon purpureus (Dicranales). Ann Bot [Internet]. 2017 Nov 10 [cited 2021 Sep 8];120(5):845-54. Available from: https://academic.oup.com/aob/article/120/5/845/3947929
  133. 133. Trouiller B, Charlot F, Choinard S, Schaefer DG, Nogué F. Comparison of gene targeting efficiencies in two mosses suggests that it is a conserved feature of Bryophyte transformation. Biotechnol Lett 2007 2910 [Internet]. 2007 Jun 13 [cited 2021 Sep 8];29(10):1591-8. Available from: https://link.springer.com/article/10.1007/s10529-007-9423-5
  134. 134. Pederson ERA, Warshan D, Rasmussen U. Genome Sequencing of Pleurozium schreberi: The Assembled and Annotated Draft Genome of a Pleurocarpous Feather Moss. G3 Genes, Genomes, Genet [Internet]. 2019 Sep 1 [cited 2021 Sep 7];9(9):2791-7. Available from: https://www.g3journal.org/content/9/9/2791
  135. 135. Heck MA, Lüth VM, Gessel N van, Krebs M, Kohl M, Prager A, et al. Axenic in vitro cultivation of 19 peat moss (Sphagnum L.) species as a resource for basic biology, biotechnology, and paludiculture. New Phytol [Internet]. 2021 Jan 1 [cited 2021 Sep 8];229(2):861-76. Available from: https://nph-onlinelibrary-wiley-com.ezproxy.library.wur.nl/doi/full/10.1111/nph.16922
  136. 136. Kariyawasam IU, Price MJ, Bell NE, Long DG, Mill RR, Hyvönen J. Unearthing a lectotype for Polytrichum commune Hedw. (Bryophyta, Polytrichaceae). Taxon [Internet]. 2021 Jun 1 [cited 2021 Sep 7];70(3):653-9. Available from: https://onlinelibrary.wiley.com/doi/full/10.1002/tax.12444
  137. 137. Goryunov D V., Sotnikova EA, Goryunova S V., Kuznetsova OI, Logacheva MD, Milyutina IA, et al. The Mitochondrial Genome of Nematodontous Moss Polytrichum commune and Analysis of Intergenic Repeats Distribution Among Bryophyta. Divers 2021, Vol 13, Page 54 [Internet]. 2021 Feb 1 [cited 2021 Sep 6];13(2):54. Available from: https://www.mdpi.com/1424-2818/13/2/54/htm
  138. 138. Jin X-J, Zhu R-L. The complete plastome of Polytrichum commune Hedw. (Polytrichaceae, Bryophyta). http://www-tandfonline-com.ezproxy.library.wur.nl/action/authorSubmission?journalCode=tmdn20&page=instructions [Internet]. 2021 [cited 2021 Sep 7];6(5):1645-7. Available from: https://www-tandfonline-com.ezproxy.library.wur.nl/doi/abs/10.1080/23802359.2021.1927223
  139. 139. Pan Z, Pitt WG, Zhang Y, Wu N, Tao Y, Truscott TT. The upside-down water collection system of Syntrichia caninervis. Nat Plants 2016 27 [Internet]. 2016 Jun 6 [cited 2021 Sep 8];2(7):1-5. Available from: https://www.nature.com/articles/nplants201676
  140. 140. Rahmatpour N, Perera N V., Singh V, Wegrzyn JL, Goffinet B. High gene space divergence contrasts with frozen vegetative architecture in the moss family Funariaceae. Mol Phylogenet Evol. 2021 Jan 1;154:106965

Written By

Jordi Floriach-Clark, Han Tang and Viola Willemsen

Submitted: 16 September 2021 Reviewed: 20 September 2021 Published: 29 October 2021