15
ORIGINAL PAPER Deciphering the ontogeny of a sympodial tree Evelyne Costes Yann Gue ´don Received: 26 April 2011 / Revised: 18 August 2011 / Accepted: 23 November 2011 / Published online: 13 December 2011 Ó Springer-Verlag 2011 Abstract This paper addresses the identification and characterization of developmental patterns in the whole structure of a sympodial species, the apple tree. Dedicated stochastic models (hidden variable-order Markov chains) were used to (i) categorise growth units (GUs) on the basis of their morphological characteristics (number of nodes and presence/absence of flowering) and position along axes, (ii) analyse dependencies between successive GUs and (iii) identify repeated patterns in GU sequences. Two successive phases, referred to as ‘‘adolescent’’ and ‘‘adult’’, were identified in two apple tree cultivars. In the adolescent phase, ‘‘very’’ long monocyclic GUs were followed by long polycyclic GUs, whereas in the adult phase medium GUs were preferentially followed by short GUs. Flowering GUs constituted a preferential pathway between vegetative GUs of decreasing vigour (long, medium and short) and generated patterns that were interpreted with respect to fruiting regularity. The proposed modelling gave a global and quantitative picture of the two-scale structuring of apple tree ontogeny: a coarse scale corresponding to the succession of the previously mentioned phases and a fine scale corresponding to the alternation between flowering and vegetative GUs. This led us to propose a synthetic scheme of apple tree ontogeny that combines growth phases, polycyclism and flowering, and which could be transposed to other sympodial trees. Keywords Floral differentiation Growth phase Hidden Markov model Tree architecture Variable-order Markov chain Introduction Plant development results from meristem activity occurring through a sequence of developmental phases, abstracted by the term ‘‘ontogeny’’. During ontogeny, the morphological characteristics of plant entities such as growth units (GUs) or annual shoots change over time (Nozeran 1984). Such changes have been investigated in a number of forest and fruit tree species (e.g. Sabatier and Barthe ´le ´my 1999; Suzuki 2002; Costes et al. 2003; Solar and Stampar 2006) on the basis of sampling of shoot categories. The floral differentiation of a meristem is a key developmental stage in plant ontogeny. Contrary to monocarpic plants where floral differentiation occurs in all aerial meristems and ends the plant’s life cycle, floral differentiation in perennial polycarpic species occurs recurrently throughout ontogeny but in a proportion of meristems only (Bangerth 2009). The occurrence of the first flowering ends up the juvenile phase, the plant then entering its mature phase (Hackett 1985). During the mature phase, the floral differentiation of shoot meristems is not ‘‘automatic’’, but has often been reported as irregular, in particular in fruit trees; see Wilkie et al. (2008) and references therein. After floral differentiation of a shoot apical meristem, growth resumes usually by sym- podial branching, i.e. through activation of one or several lateral meristems (Halle ´ et al. 1978). The development of a Communicated by A. Braeuning. E. Costes INRA, UMR AGAP, Architecture and Functioning of Fruit Species Team, 34398 Montpellier, France e-mail: [email protected] Y. Gue ´don (&) CIRAD, UMR AGAP and INRIA, Virtual Plants, 34398 Montpellier, France e-mail: [email protected] 123 Trees (2012) 26:865–879 DOI 10.1007/s00468-011-0661-8

Deciphering the Ontogeny of a Sympodial Tree

Embed Size (px)

DESCRIPTION

8

Citation preview

Page 1: Deciphering the Ontogeny of a Sympodial Tree

ORIGINAL PAPER

Deciphering the ontogeny of a sympodial tree

Evelyne Costes • Yann Guedon

Received: 26 April 2011 / Revised: 18 August 2011 / Accepted: 23 November 2011 / Published online: 13 December 2011

� Springer-Verlag 2011

Abstract This paper addresses the identification and

characterization of developmental patterns in the whole

structure of a sympodial species, the apple tree. Dedicated

stochastic models (hidden variable-order Markov chains)

were used to (i) categorise growth units (GUs) on the basis

of their morphological characteristics (number of nodes

and presence/absence of flowering) and position along

axes, (ii) analyse dependencies between successive GUs

and (iii) identify repeated patterns in GU sequences. Two

successive phases, referred to as ‘‘adolescent’’ and ‘‘adult’’,

were identified in two apple tree cultivars. In the adolescent

phase, ‘‘very’’ long monocyclic GUs were followed by

long polycyclic GUs, whereas in the adult phase medium

GUs were preferentially followed by short GUs. Flowering

GUs constituted a preferential pathway between vegetative

GUs of decreasing vigour (long, medium and short) and

generated patterns that were interpreted with respect to

fruiting regularity. The proposed modelling gave a global

and quantitative picture of the two-scale structuring of

apple tree ontogeny: a coarse scale corresponding to the

succession of the previously mentioned phases and a fine

scale corresponding to the alternation between flowering

and vegetative GUs. This led us to propose a synthetic

scheme of apple tree ontogeny that combines growth

phases, polycyclism and flowering, and which could be

transposed to other sympodial trees.

Keywords Floral differentiation � Growth phase � Hidden

Markov model � Tree architecture � Variable-order Markov

chain

Introduction

Plant development results from meristem activity occurring

through a sequence of developmental phases, abstracted by

the term ‘‘ontogeny’’. During ontogeny, the morphological

characteristics of plant entities such as growth units (GUs)

or annual shoots change over time (Nozeran 1984). Such

changes have been investigated in a number of forest and

fruit tree species (e.g. Sabatier and Barthelemy 1999;

Suzuki 2002; Costes et al. 2003; Solar and Stampar 2006)

on the basis of sampling of shoot categories. The floral

differentiation of a meristem is a key developmental stage

in plant ontogeny. Contrary to monocarpic plants where

floral differentiation occurs in all aerial meristems and ends

the plant’s life cycle, floral differentiation in perennial

polycarpic species occurs recurrently throughout ontogeny

but in a proportion of meristems only (Bangerth 2009). The

occurrence of the first flowering ends up the juvenile phase,

the plant then entering its mature phase (Hackett 1985).

During the mature phase, the floral differentiation of shoot

meristems is not ‘‘automatic’’, but has often been reported

as irregular, in particular in fruit trees; see Wilkie et al.

(2008) and references therein. After floral differentiation of

a shoot apical meristem, growth resumes usually by sym-

podial branching, i.e. through activation of one or several

lateral meristems (Halle et al. 1978). The development of a

Communicated by A. Braeuning.

E. Costes

INRA, UMR AGAP, Architecture and Functioning of Fruit

Species Team, 34398 Montpellier, France

e-mail: [email protected]

Y. Guedon (&)

CIRAD, UMR AGAP and INRIA, Virtual Plants,

34398 Montpellier, France

e-mail: [email protected]

123

Trees (2012) 26:865–879

DOI 10.1007/s00468-011-0661-8

Page 2: Deciphering the Ontogeny of a Sympodial Tree

polycarpic sympodial plant is thus characterized, at a fine

scale, by the occurrence of repeated patterns (also called

words or motifs) resulting from the alternation between

flowering and vegetative growth and, at a coarser scale, by

two successive developmental phases, the juvenile and

mature phases. This two-scale structuring can be consid-

ered as a general property of polycarpic sympodial plants

even though the growth characteristics at each scale, such

as the length of a phase or the number of nodes of GUs, are

modulated by the environmental conditions. The apple tree

is an interesting example of a polycarpic sympodial tree

since it combines terminal flowering followed by sympo-

dial branching (Crabbe and Escobedo-Alvarez 1991) with

the capability to develop polycyclic shoots (i.e. shoots that

include several GUs within a year, Halle et al. 1978). Its

architectural development has been largely studied, but

despite several studies dedicated to the apple tree estab-

lishment phase (Costes et al. 2003) or to the maintenance

of equilibrium between fruit and leaf production in the

mature phase (Lauri et al. 1995), there is no global and

quantitative picture that integrates these two scales of

structuring of apple tree ontogeny. This is mainly due to

the irregularity of flowering combined with the complexity

of perennial plant structures. In the present study, we aimed

at proposing a new modelling approach to capture simul-

taneously the two scales at which polycarpic sympodial

trees organise during ontogeny, taking the apple tree as a

case study. In particular, we addressed two main questions:

(i) how do the patterns of alternation between flowering

and vegetative GUs change between phases during tree

ontogeny? (ii) Are there useful markers of transition

between phases?

We propose to apply models combining structure with

probabilities for analysing complex developmental patterns

often found in perennial plants. Probabilities enable to

represent fluctuations due to unobservable biological

functions and their interactions acting at more microscopic

scales than the scale of measurement. Generally, the

objective of this type of analysis is to identify and char-

acterise complex patterns which cannot be grasped directly

from the data rather than to quantify well-known behav-

iours. In this respect, the analysis of complex perennial

plant patterns fits into the general framework of pattern

theory (Grenander and Miller 2006) and the models of

interest are often stochastic models. Different stochastic

models have previously been proposed to analyse tree

structure development at the GU or annual shoot scale

(Durand et al. 2005; Guedon et al. 2007; Chaubert-Pereira

et al. 2009). In these studies, it was assumed that the

functioning of a meristem during a growing period influ-

ences its functioning during subsequent periods. As in

these previous studies, we assumed dependencies between

successive GUs built by the same meristem or deriving

from one another by sympodial branching. In Durand et al.

(2005), it was assumed that the differentiation state of

meristems can be deduced from the morphological char-

acteristics of the GUs, in particular the number of nodes,

the length and the presence/absence of flowering. A sto-

chastic model, called the hidden Markov tree model, was

applied to infer the differentiation state sequences in two

tree species (apple tree and bush willow). In apple tree,

four GU types were identified: three vegetative (long,

medium and short) and one flowering, and these were

shown to have specific locations in trees. However, the

hidden Markov tree model did allow neither to model the

dependency of a GU beyond its parent GU nor to highlight

successive phases. Here, we again chose to apply a ‘‘hid-

den’’ model (since GU types will be deduced from GU

morphological characteristics) to analyse frequent patterns

that include more than two successive GUs, for instance

three GUs, e.g. [flowering GU, vegetative (either long

medium or short) GU, flowering GU]. Most of the methods

for analysing local dependencies in sequences rely on high-

order Markov chains. However, the number of free

parameters of a Markov chain increases exponentially with

its order, i.e. with the memory length taken into account.

For instance, in the case of four states (corresponding to

four GU types), the number of free parameters is 3 for a

zero-order, 12 for a first-order, 48 for a second-order

Markov chain, etc. Since there are no models ‘‘in

between’’, this very discontinuous increase in the number

of free parameters causes the estimated high-order Markov

chains to be generally overparameterized. This drawback

can be overcome by defining sub-classes of parsimonious

high-order Markov chains such as variable-order Markov

chains (Ron et al. 1996; Buhlmann and Wyner 1999) where

the order is variable and depends on the ‘‘context’’ within

the sequence, instead of being fixed. We thus chose to

model the sequence of meristem differentiation states

during apple tree ontogeny using a hidden variable-order

Markov chain, the objective being to draw a global picture

of the ontogeny of a perennial sympodial species. Our goal

was to characterise jointly successive growth phases,

changes in GU morphological characteristics and devel-

opmental events such as terminal flowering.

Materials and methods

Plant material and tree description

The studied trees belong to two apple scion-cultivars,

‘Braeburn’ and ‘Fuji’ that were grafted on Lancep Pajam�1

(type M9). Trees were planted with 1.8 9 6 m spacing in

the winter of 1994–1995 at the Melgueil INRA experi-

mental station (South East France). In the first year of

866 Trees (2012) 26:865–879

123

Page 3: Deciphering the Ontogeny of a Sympodial Tree

growth the trees were trained in a vertical axis by selecting

long axillary shoots along the main axis which was main-

tained without pruning. In the spring of the third year, the

trunks and long shoots were bent. In June of years 4–6,

young fruits were thinned manually to one fruit per inflo-

rescence and all axillary fruits were removed on 1-year-old

wood.

Two 6-year-old apple trees for each of the Braeburn and

Fuji cultivars were fully described; see Costes et al. (2003).

Each tree was described using three scales of organisation

corresponding to the axes, GUs and metamers (as defined

by White 1979). Flowering GUs for which the apical

meristem had differentiated into an inflorescence were

distinguished from vegetative GUs (Fig. 1). The number of

nodes was counted on the extension GUs only (for non-

extended vegetative GUs referred to as short GUs in the

following, the number of nodes was set at the default value

of 1). Bud scars and metamers whose axillary buds were

not visible to the naked eye were ignored.

GU sequence modelling

From the initial database, sequences of GUs were extracted

for all axes, including the trunks (Fig. 1). Each sequence

thus corresponded to a new lateral branching (except the

main axes of the trees). After a flowering occurrence, the

new axis arising from sympodial branching was considered

as the continuation in the sequence. In apple tree, this new

axis is called ‘‘bourse shoot’’. When two axes arose from

the same flowering GU, the most distal one was chosen as

the continuation. When no continuation was observed, the

sequence was completed by a final ‘‘dead’’ GU, according

to the phenomenon called ‘‘extinction’’ by Lauri et al.

(1995). Each sequence included three variables: year of

growth, number of nodes and non-flowering/flowering

character of the GU. In all, 1,194 sequences of cumulated

length 5,523 were extracted from the two Braeburn trees

and 2,034 sequences of cumulated length 6,072 were

extracted from the two Fuji trees. For a given cultivar, the

cumulated length of the sequences is the total number of

GUs of the two trees.

In the following, we first introduce high-order Markov

chains. Then, variable-order Markov chains and hidden

Markov models based on variable-order Markov chains,

which are the stochastic models used in this study, are

defined. In the case of an rth-order Markov chain {St;

t = 0, 1, …}, the conditional distribution of St given

S0, …, St-1 depends only on the values of St-r, …, St-1

but not further on S0, …, St-r-1,

P St ¼ stjSt�1 ¼ st�1; . . .; S0 ¼ s0ð Þ¼ P St ¼ stjSt�1 ¼ st�1; . . .; St�r ¼ st�rð Þ:

In our context, the random variable St represents GU

type and can take the four possible values L, M, S and F, for

long, medium, short and flowering, respectively. These

four possible values correspond to the Markov chain states.

A supplementary ‘‘end’’ state is added to formalise axis

death (denoted by D in the following). A J-state rth-order

Markov chain has Jr(J - 1) independent transition

probabilities. Therefore, the number of free parameters of

a Markov chain increases exponentially with the order. Let

the transition probabilities of a second-order Markov chain

be given by

phij ¼ P St ¼ jjSt�1 ¼ i; St�2 ¼ hð Þ withX

j

phij ¼ 1:

These transition probabilities can be arranged as a J2 9 J

matrix where the row (phi0,…,phiJ-1) corresponds to the

transition distribution attached to the [state h, state i]

memory. If for a given state i and for all pairs of states

(h, h0) with h 6¼ h0, phij ¼ ph0ij, i.e. once St-1 is known, St-2

conveys no further information about St, the J memories of

length 2 [state h, state i] with h = 0, …, J - 1 can be

grouped together and replaced by the single [state i]

memory of length 1 with associated transition distribution

(pi0, …, piJ-1). This illustrates the principle for building a

variable-order Markov chain. In a variable-order Markov

chain, the order (or memory length) is variable and depends

on the ‘‘context’’ within the sequence. The memories of a

Markov chain can be arranged as a memory tree such that

each vertex (i.e. element of a tree graph) is either a terminal

vertex or has exactly J ‘‘offspring’’ vertices. In practice, the

memories corresponding to unobserved contexts are not

included in the memory tree (this is the case for the memory

FF that was not observed in our conditions). The memories

associated with the J vertices (memories of length r ? 1)

deriving from a given vertex (memory of length r) are

Fig. 1 Schematic representation of a 4-year-old apple tree branch at

the growth unit (GU) scale. GUs of each axis are labelled according to

their floral (F) or vegetative nature, and, in this latter case, according

to their number of nodes (L for long, M for medium and S for short).

The GU sequence along the branch is LLFLLFLFL. Four examples of

GU sequences are provided for lateral shoots: SSFS, LFMFM, SFSand MFM

Trees (2012) 26:865–879 867

123

Page 4: Deciphering the Ontogeny of a Sympodial Tree

obtained by prefixing the parent memory with each possible

state. For instance, in Figs. 2a and 3a, the 3 second-order

memories LF, MF and SF derive from the first-order

memory F. A transition distribution is associated with each

terminal vertex of this memory tree.

A specific problem with variable-order Markov chains

concerns the definition of initial probabilities for deter-

mining the first state in the sequence. We adopted the

approach proposed by Ron et al. (1996). They showed that

the model (consisting of the ‘‘permanent’’ memories cor-

responding to the terminal vertices of the memory tree) can

be augmented with the memories corresponding to the

interior vertices to model the beginning of the sequences.

These supplementary memories are called ‘‘transient’’.

They can only be visited once at the beginning of the

sequence while permanent memories can be visited more

than once in a sequence. With these supplementary tran-

sient memories, the initial probabilities pj = P(S0 = j) are

defined as usual for first-order memories. For instance, in

the case of the model shown in Fig. 2a, a sequence can be

initialized in the initial transient memory denoted by F0

(and not in the permanent memories LF, MF and SF

deriving from F). Since each sequence corresponds to a

new lateral branching (except the main axes of the trees),

the initial probabilities represent the branching probabili-

ties estimated globally for the entire trees. The building of

a variable-order Markov chain consists in selecting the

memories in an optimal way (Csiszar and Talata 2006) and

estimating the corresponding initial and transition proba-

bilities; see ‘‘Appendix A’’ for the principle of the selection

of the memories.

In this study, instead of first classifying GUs in cat-

egories and then characterising the succession of GU

types along the axes, we rather chose to merge these two

steps and build a single integrative statistical model that

can be optimally estimated. In this statistical model,

which is a hidden Markov model (Ephraim and Merhav

2002) based on a variable-order Markov chain, the non-

observable variable-order Markov chain represents the

succession of GU types along the axes. The GU types

are not observable directly but only indirectly through

the two observed variables, namely the number of nodes

and the non-flowering/flowering character of the GU. A

hidden variable-order Markov chain can be viewed as a

two-level stochastic process, i.e. a pair of stochastic

processes {St, Xt} where the ‘‘output’’ process {Xt} is

related to the ‘‘state’’ process {St}, which is a finite-state

variable-order Markov chain, by the observation

probabilities

bj yð Þ ¼ P Xt ¼ yjSt ¼ jð Þ withX

y

bj yð Þ ¼ 1:

The definition of the observation probabilities expresses

the assumption that the output process at time t depends

only on the non-observable Markov chain at time

t. Extension to the multivariate case is straightforward

since, in this latter case, the observed variables at time t are

assumed to be conditionally independent given the state

St = st. The probability of observing a vector is thus

simply the product of the probabilities of observing each

variable. There is indeed no observation distribution

attached to the death ‘‘end’’ state.

In the case of a small set of possible outputs (as in the

case of a binary observed variable such as non-flowering/

flowering), the observation probabilities are estimated

directly and can be arranged as a J 9 N matrix (N is the

number of possible outputs) with all rows summing to one.

In the case of a larger set of possible outputs (assumed to

be generated by a count variable such as the number of

nodes of a GU), discrete parametric observation distribu-

tions are estimated. In this study we chose to use as pos-

sible discrete parametric observation distributions,

binomial distributions, Poisson distributions and negative

binomial distributions with an additional shift parameter;

Fig. 2 Variable-order Markov chain for Braeburn cultivar. L, M, S, F,

stand for long, medium, short and flowering, respectively. a Memory

tree: the memory tree must be read from the left to the right, and each

‘‘column’’ corresponds to memories of a given length; b Transition

graph: each vertex represents a possible memory (of length 1 or 2).

The vertex edged by a dashed line corresponds to the initial transient

memory (visited once at the beginning of a sequence) while the

vertices edged by a continuous line correspond to permanent

memories (visited more than once in a sequence). Transitions with

associated probability [0.04 are represented by arcs. The associated

probabilities are noted nearby. Transitions towards the death end state

are not shown

868 Trees (2012) 26:865–879

123

Page 5: Deciphering the Ontogeny of a Sympodial Tree

see ‘‘Appendix B’’ for formal definitions of these

distributions.

The maximum likelihood estimation of the parameters

of a hidden variable-order Markov chain requires an iter-

ative optimisation technique, which is an application of the

Expectation–Maximization (EM) algorithm (Ephraim and

Merhav 2002). Once a hidden variable-order Markov chain

has been estimated, the most probable state sequence can

be computed for each observed sequence using the so-

called Viterbi algorithm. This state sequence, which is

called the restored state sequence, can be interpreted as the

optimal labelling of the observed sequence where a type

chosen from among long, medium, short and flowering is

affected to each successive GU.

Different methods have been proposed to assess the

modelling of repeated patterns by (hidden) variable-order

Markov chains (Guedon et al. 2001). Here, we will focus

on recurrence time distributions which appear to be par-

ticularly appropriate for analysing flowering regularity

over time. The recurrence time in a given state is defined as

the number of transitions between two occurrences of this

state. Recurrence time distributions can in particular help

to highlight pseudo-periodicities in the successive occur-

rences of a given state along sequences.

All the statistical analyses were made using the statis-

tical package of VPlants software integrated in the

OpenAlea platform and available at http://openalea.gforge.

inria.fr/wiki/doku.php?id=openalea.

Results

Selection of the memories of the variable-order Markov

chains

In a first step, a four-state hidden first-order Markov chain

was estimated for each cultivar on the basis of the sequences

of GUs where the two observed variables were the number of

nodes and the non-flowering/flowering character. The four

states of the underlying first-order Markov chain corre-

sponded to long, medium, short and flowering GUs; see

below for the characterization of these states in the case of

hidden variable-order Markov chains. This is a direct

transposition to sequences of the modelling approach pro-

posed by Durand et al. (2005) in the case of tree-structured

data. The restored state sequences were then computed using

the estimated hidden first-order Markov chain and the

memories of a variable-order Markov chain were selected on

the basis of these restored state sequences; see ‘‘Appendix

A’’ for the principle of the selection of these memories. This

approach is justified by the fact that (i) the labelling of the

observed sequences was unambiguous for sequences which

Fig. 3 Variable-order Markov

chain for Fuji cultivar. L, M, S,

F, stand for long, medium, short

and flowering, respectively.

a Memory tree: the memory tree

must be read from the left to the

right, and each ‘‘column’’

corresponds to memories of a

given length; b Transition

graph: each vertex represents a

possible memory (of length 1 or

2). The vertices edged by a

dashed line correspond to the

initial transient memories

(visited once at the beginning of

a sequence) while the vertices

edged by a continuous linecorrespond to permanent

memories (visited more than

once in a sequence). Transitions

with associated probability

[0.04 are represented by arcs.

The associated probabilities are

noted nearby. Transitions

towards the death end state are

not shown. The dotted arcscorrespond to infrequent

transitions and the vertex edged

by a dotted line corresponds to

an infrequent permanent

memory

Trees (2012) 26:865–879 869

123

Page 6: Deciphering the Ontogeny of a Sympodial Tree

do not contain extension GUs, and (ii) the posterior proba-

bilities of the restored state sequences (i.e. weight of the

restored state sequence among all the possible state

sequences that can explain a given observed sequence) were

most often high for the observed sequences which contain

extension GUs; see below further comments in the case of

hidden variable-order Markov chains. For both cultivars, the

variable-order Markov chains were mixed first-/second-

order Markov chains. For the Braeburn cultivar, the memo-

ries of the variable-order Markov chain (i.e. terminal vertices

of the memory tree) were {L, M, S, LF, MF, SF} while for the

Fuji cultivar, the memories were {L, M, LS, MS, SS, FS, LF,

MF,SF}; see the memory tree for Braeburn in Fig. 2a and for

Fuji in Fig. 3a. Thus, for both cultivars, the type of the GU

following a long or medium ‘‘parent’’ GU did not depend on

the type of the ‘‘grand-parent’’ GU (first-order memories L

and M) while the type of the GU following a flowering

‘‘parent’’ GU depended on the type of the ‘‘grand-parent’’

GU (second-order memories LF, MF, SF deriving from F).

The difference between the two cultivars concerned the type

of the GU following a short ‘‘parent’’ GU: this depended on

the type of the ‘‘grand-parent’’ GU for Fuji (second-order

memories LS, MS, SS, FS deriving from S), but not for

Braeburn (first-order memory S).

Estimation of hidden variable-order Markov chains

In a second step, a hidden variable-order Markov chain was

estimated for each cultivar where the underlying variable-

order Markov chain had the memories previously selected,

i.e. {L, M, S, LF, MF, SF} for Braeburn and {L, M, LS, MS,

SS, FS, LF, MF, SF} for Fuji. The iterative maximum

likelihood estimation procedure (EM algorithm) was ini-

tialized for the underlying variable-order Markov chain

such that all the transitions were possible. Specific transi-

tion probabilities were estimated for the initial transient

memories (F0 for both cultivars and S0 for Fuji) because of

the large number of short sequences in these data samples;

see the corresponding counts in Table 1. More generally,

the high values for most of the transition counts give

empirical evidence of the accuracy of the estimated tran-

sition probabilities. The estimated observation probability

matrix for the non-flowering/flowering variable was

degenerate (i.e. the estimated observation probabilities

Table 1 Transition probabilities of hidden variable-order Markov

chains estimated for the Braeburn and Fuji cultivars (memories in

rows, states in columns); e.g. the probability in the cell (long

flowering, medium) is pLFM ¼ P St ¼ MjSt�1 ¼ F; St�2 ¼ Lð Þ where L,

M, S, F stand for long, medium, short and flowering, respectively

Transition probability (next memory)

Long Medium Short Flowering Death Count

Braeburn memory

Long 0.3 (L) 0.03 (M) 0.04 (S) 0.61 (LF) 0.02 (D) 495

Medium 0 0.03 (M) 0 0.96 (MF) 0.01 (D) 574

Short 0.01 (L) 0.03 (M) 0.04 (S) 0.9 (SF) 0.02 (D) 1,238

Initial flowering 0.15 (L) 0.11 (M) 0.74 (S) 0 0 789

Long flowering 0.49 (L) 0.5 (M) 0 0 0.01 (D) 255

Medium flowering 0 0.27 (M) 0.72 (S) 0 0.01 (D) 348

Short flowering 0.04 (L) 0.12 (M) 0.83 (S) 0 0.01 (D) 683

Fuji memory

Long 0.31 (L) 0 0.01 (LS) 0.68 (LF) 0 444

Medium 0.05 (L) 0.06 (M) 0.05 (MS) 0.8 (MF) 0.04 (D) 494

Initial short 0.03 (L) 0.02 (M) 0.29 (SS) 0.66 (SF) 0 975

Long short 0.31 (L) 0 0 0.69 (SF) 0 2

Medium short 0 0 0.13 (SS) 0.74 (SF) 0.13 (D) 17

Short short 0.01 (L) 0.01 (M) 0.18 (SS) 0.6 (SF) 0.2 (D) 262

Flowering short 0.04 (L) 0.06 (M) 0.41 (SS) 0.38 (SF) 0.11 (D) 413

Initial flowering 0.05 (L) 0.14 (M) 0.81 (FS) 0 0 676

Long flowering 0.55 (L) 0.33 (M) 0.03 (FS) 0 0.09 (D) 195

Medium flowering 0.01 (L) 0.58 (M) 0.38 (FS) 0 0.03 (D) 158

Short flowering 0.03 (L) 0.13 (M) 0.79 (FS) 0 0.05 (D) 571

The memory reached by a transition is indicated between brackets after the corresponding probability. The transition counts (last column) were

extracted from the state sequences restored using the hidden variable-order Markov chains estimated for the Braeburn and Fuji cultivars. The

initial transient memories, which correspond to first states in a sequence, are indicated in italics

870 Trees (2012) 26:865–879

123

Page 7: Deciphering the Ontogeny of a Sympodial Tree

were either 0 or 1), with three vegetative states and one

flowering state (row: state L, M, S, F; column: non-flow-

ering/flowering):

B ¼

1 0

1 0

1 0

0 1

0BB@

1CCA:

This resulted from the iterative estimation procedure

which was initialized with a model where the observation

distributions for the non-flowering/flowering variable were

not degenerate.

The free parameters of the estimated hidden variable-

order Markov chains decompose in:

• Braeburn cultivar: 3 independent initial probabilities,

19 independent transition probabilities (Table 1),

(instead of 14 independent transitions probabilities in

the case of an underlying first-order Markov chain),

• Fuji cultivar: 3 independent initial probabilities, 31

independent transition probabilities (Table 1), (instead

of 13 independent transitions probabilities in the case of

an underlying first-order Markov chain);

• 11 free parameters for the observation distributions

estimated for the number of nodes variable (and no free

parameters for the observation distributions estimated

for the non-flowering/flowering variable).

The number of free parameters seems reasonable in view

of the sample sizes (cumulated length of the sequences:

5,523 for Braeburn and 6,072 for Fuji). It should be noted

that the death ‘‘end’’ state only adds one column in the

transition probability matrix and no observation distribu-

tions are defined for this state. The Bayesian information

criterion (BIC) favours the hidden variable-order chain

both for Braeburn (BIC12 = -20,075 instead of BIC1 =

-20,560 for a hidden first-order chain) and Fuji

(BIC12 = -22,783 instead of BIC1 = -23,470 for a hid-

den first-order chain); see ‘‘Appendix A’’ for a detailed

presentation of BIC in the case of variable-order Markov

chains. The rules of thumb of Jeffreys (see Kass and Raftery

1995) suggest that a difference of BIC of at least 2log

100 = 9.2 is needed to deem the model with the higher BIC

substantially better.

The graph of the possible transitions between memories

of the underlying estimated variable-order Markov chain is

shown in Fig. 2b for Braeburn and Fig. 3b for Fuji. In the

Braeburn model, a repetitive structure is apparent with

three groups of two memories: {L, LF}, {M, MF}, {S, SF}.

If we consider a vegetative state V which is L for the first

group, M for the second group and S for the third group, the

within-group transition probabilities pVF and pVFV are

always high ([0.49 except pMFM = 0.27) (Table 1). There

is also a preferential order of succession among the three

groups with transitions from the first towards the second

group and from the second towards the third group having

high probabilities: pLFM = 0.5 and pMFS = 0.72, respec-

tively. But the order of succession is not strict since it is

nevertheless possible to reach the second group from the

third (with pSFM = 0.12). The Fuji model (Fig. 3b) is

similar to the Braeburn model in terms of transition prob-

abilities except for the third group corresponding to state S.

The two-memory group {S, SF} of the Braeburn model is

replaced by a three-memory group {SS, FS, SF} in the

Fuji model. As seen previously, the within-group transition

probabilities pFSF, pSFS, pFSS and pSSF are all high

(Table 1). The difference between the cultivars is sup-

ported by counts for the second-order memories LS, MS,

SS and FS which are 2, 17, 262 and 413 for Fuji (see

Table 1) and 11, 0, 11 and 1,043 for Braeburn. Hence,

while for Fuji both the SS and FS memories are highly

represented in the data sample, only the FS memory is

highly represented for Braeburn, and the second-order FS

memory is therefore roughly equivalent to the first-order S

memory. Differences between the transition distributions

pLF0; . . .; pLFJ�1ð Þ; pMF0; . . .; pMFJ�1ð Þ and pSF0; . . .; pSFJ�1ð Þfor the second-order memories LF, MF and SF deriving

from F (and for the second-order memories MS, SS and FS

deriving from S in the case of Fuji) can be noted; see the

corresponding rows in Table 1. This is an a posteriori

justification of the selection of these second-order memo-

ries. For both cultivars, the distinction of the flowering

GUs on the basis of their vegetative ‘‘context’’ (with the

three memories LF, MF and SF) makes apparent the stages

in the succession of GUs. With a simple first-order Markov

chain where the three second-order memories LF, MF and

SF are collapsed onto a first-order memory F, these stages

cannot be identified since, whatever the type of the previ-

ous vegetative GU, vegetative GU of any type can followed

a flowering GU. As an illustration, among the patterns

VFV, only the patterns LFL, LFM, MFM, MFS and SFS can

occur with high probabilities in the variable-order Markov

chain case while all the patterns VFV including LFS, SFL

and SFM can occur with high probabilities in the first-order

Markov chain case.

It is noteworthy that the estimated hidden variable-order

Markov chains were only partially hidden since the vege-

tative and flowering GUs were differentiated unambigu-

ously by the non-flowering/flowering variable. Similarly,

short GUs were unambiguously defined by the number of

nodes variable set at the default value of 1. Hence, the

hidden character of the models only concerned the long and

medium GUs which were characterized by both their

number of nodes (see the corresponding observation dis-

tributions for states L and M in Fig. 4 and their character-

istics in Table 2) and position along the sequences (see

below). Consequently, the labelling of the observed

Trees (2012) 26:865–879 871

123

Page 8: Deciphering the Ontogeny of a Sympodial Tree

sequences was unambiguous for sequences which do not

contain extension GUs (609 out of 1,194 sequences for

Braeburn and 1,342 out of 2,034 for Fuji). For sequences

containing extension GUs, the posterior probabilities of the

restored state sequences were most often high: 48% above

0.8 and 91% above 0.5 for Braeburn on the basis of 585

sequences; 60% above 0.8 and 94% above 0.5 for Fuji on

the basis of 692 sequences. This justifies the use of

empirical distributions or characteristics deduced from the

restored state sequences for interpreting the output of the

estimated hidden variable-order Markov chains. It should

be noted that among the long GUs, there is a small

proportion of ‘‘very long’’ GUs which correspond to GUs

established during the first and second years of growth (18

for Braeburn and 13 for Fuji with more than 40 nodes

compared with a total of 533 and 512 long GUs, respec-

tively); see the tails of the corresponding frequency dis-

tributions in Fig. 4. Due to their small number, these very

long GUs could not be modelled as a supplementary state

in the hidden variable-order Markov chains. We checked

that the differences in mean number of nodes were small

and often statistically non-significant between sub-samples

of long GUs for years 3–6 and between sub-samples of

medium GUs for the different years (results not shown).

This is an a posteriori validation of the assumption of a

hidden Markov model based on a time-homogeneous

Markov chain. We also checked that the empirical number

of nodes distribution for extension GUs was well fitted by

the mixture of long and medium state observation distri-

butions (see Fig. 4a for Braeburn and Fig. 4b for Fuji)

using in particular P–P plots (plots not shown).

The accuracy of the estimated hidden variable-order

Markov chains for modelling patterns in the succession of

GUs can be illustrated by the improved fit of the recurrence

time distributions in the case of a hidden variable-order

Markov chain compared with a simple hidden first-order

Markov chain; see Fig. 5 for Braeburn and Fig. 6 for Fuji.

In these Figures, the empirical distributions were extracted

from the restored state sequences computed using the

estimated hidden variable-order Markov chain or hidden

first-order Markov chain (these empirical distributions only

differ for long and medium GUs between the two hidden

Markov models) while the theoretical distributions were

computed using the two compared hidden Markov chains;

see Guedon et al. (2001) for inclusion of the bias due to

short sequence length in the computation of the recurrence

time distributions.

Analysing model outputs: ontogenetic stages

For both cultivars, the transition probabilities show that a

vegetative GU (either long, medium or short) was prefer-

entially followed by a flowering GU (see the flowering

column in Table 1) while a flowering GU was systemati-

cally followed by a vegetative GU (this entails that the

estimated hidden variable-order Markov chains do not

include the ‘‘unobserved’’ FF memory). Flowering and

vegetative GUs, therefore, alternated along the sequences.

This alternation is superimposed upon the trend corre-

sponding to GU decrease in vigour along the sequences

(long ? medium ? short). This trend is highlighted by

the high transition probabilities pVFV where V is a vegeta-

tive state chosen from among L, M and S. Since for both

cultivars the transition probabilities between distinct veg-

etative states taken in order of decreasing vigour (mainly

Fig. 4 Fit of the empirical number of nodes distribution for extension

GUs by the mixture of long and medium state observation distribu-

tions: a Braeburn cultivar; b Fuji cultivar

Table 2 Means and standard deviations (indicated between brackets)

of the estimated observation distributions for the number of nodes of

extension GUs for the Braeburn and Fuji cultivars

Cultivar State

Long Medium

Braeburn 15.44 (8.85) 9.65 (3.46)

Fuji 17.57 (8.44) 8.09 (2.9)

Nodes corresponding to bud scars and whose axillary bud was not

visible were not counted

872 Trees (2012) 26:865–879

123

Page 9: Deciphering the Ontogeny of a Sympodial Tree

pLM and pMS) were nil or very low, the decrease in vigour of

the vegetative GUs was not direct but preferentially

required an intermediate flowering stage. This is also

clearly illustrated in the graph of possible transitions; see

the comment above and Fig. 2b for Braeburn and Fig. 3b

for Fuji).

Fig. 5 Fit of recurrence time

distributions for different

growth unit (GU) types for the

Braeburn cultivar. a and

b Distributions for long GUs

computed from estimated

variable- and first-order Markov

chains, respectively; c and

d distributions for medium GUs

computed from estimated

variable- and first-order Markov

chains, respectively; e and

f distributions for short and

flowering GUs, respectively

Trees (2012) 26:865–879 873

123

Page 10: Deciphering the Ontogeny of a Sympodial Tree

Fig. 6 Fit of recurrence time

distributions for different

growth unit (GU) types for the

Fuji cultivar. a and

b Distributions for long GUs

computed from estimated

variable- and first-order Markov

chains, respectively; c and

d distributions for medium GUs

computed from estimated

variable- and first-order Markov

chains, respectively; e and

f distributions for short and

flowering GUs, respectively

874 Trees (2012) 26:865–879

123

Page 11: Deciphering the Ontogeny of a Sympodial Tree

Analysing model outputs: between- and within-year

transitions and frequent GU successions

As a consequence of spring flowering followed by sym-

podial vegetative branching, transitions from a flowering

GU towards a vegetative GU (chosen from among long,

medium and short) corresponded almost exclusively to

within-year transitions while transitions from a vegetative

GU towards a flowering GU corresponded exclusively to

between-year transitions (Table 3). Transitions from a

vegetative GU towards another vegetative GU (often of the

same type) corresponded most often to between-year

transitions except in the case of two successive long GUs.

As a consequence of the integrative statistical modelling,

vegetative polycyclism corresponding to within-year tran-

sitions is a distinctive property of long GUs (Table 3). The

phenomenon occurred fairly rarely for Fuji (10% of long

GUs, i.e. 45 out of 444 GUs, and negligible for medium

and short GUs), while it was more frequent for Braeburn

(23% for long GUs, i.e. 110 out of 484, and negligible for

medium and short GUs).

As shown above on the basis of the memory trees (see

Fig. 2a for Braeburn and Fig. 3a for Fuji) and the restored

state sequences, one of the main differences between the

Braeburn and Fuji cultivars concerned short GUs. In par-

ticular, the transitions from a short GU towards another

short GU (mostly corresponding to between-year transi-

tions) were far more frequent for Fuji than Braeburn

(Table 3). As a consequence, the FSSF pattern, which

corresponds to biennial bearing, was far more frequent for

Fuji than for Braeburn (Table 4). Conversely, the FSF and

FSFSF patterns, which correspond to regular spring

flowering, were far more frequent for Braeburn than for

Fuji. These results, together with the low frequency of the

within-year SS pattern for Fuji cultivar (22 within-year SS

in Table 3, row ‘‘total short’’, column ‘‘short’’, compared

with 69 FSSF patterns in Table 4) illustrate the biennial

bearing behaviour of the Fuji cultivar. The maintenance of

Table 3 Between- and within-year transition counts extracted from the state sequences restored using the hidden variable-order Markov chains

estimated for Braeburn and Fuji cultivars

Between- and within-year transition counts

Long Medium Short Flowering Total

Braeburn memory

Long 78 100 2 4 17 4 277 2 374 110

Medium 0 0 7 5 0 0 559 0 566 5

Short 10 4 39 4 37 6 1,107 3 1,193 17

Initial flowering 0 93 0 108 6 582 0 0 6 783

Long flowering 0 116 1 136 0 0 0 0 1 252

Medium flowering 0 0 0 106 1 239 0 0 1 345

Short flowering 0 20 1 88 6 561 0 0 7 669

Fuji memory

Long 114 42 0 1 3 2 282 0 399 45

Medium 11 9 21 8 18 6 404 0 454 23

Initial short 30 0 18 1 277 1 648 0 973 2

Long short 1 0 0 0 0 0 1 0 2 0

Medium short 0 0 0 0 2 0 11 2 13 2

Short short 1 0 2 2 39 9 157 0 199 11

Flowering short 12 1 26 1 155 12 159 0 352 14

Total short 44 1 46 4 473 22 976 2 1,539 29

Initial flowering 1 23 5 101 19 527 0 0 25 651

Long flowering 1 100 0 68 0 6 0 0 1 174

Medium flowering 0 0 0 99 2 53 0 0 2 152

Short flowering 1 13 5 77 21 427 0 0 27 517

For each GU type, the first column corresponds to between-year transitions and the second column to within-year transitions. Transitions towards

the death state are not considered. The initial transient memories, which correspond to the first states in the sequences, are indicated in italics. The

row labelled ‘‘Total short’’ cumulates the counts for the initial transient ‘‘initial short’’ memory and the permanent second-order ‘‘long short’’,

‘‘medium short’’, ‘‘short short’’ and ‘‘flowering short’’ memories. This row is inserted for the Fuji cultivar to help the comparison with the first-

order ‘‘short’’ memory of the Braeburn cultivar

Trees (2012) 26:865–879 875

123

Page 12: Deciphering the Ontogeny of a Sympodial Tree

regular bearing over three successive years can be illus-

trated by patterns of length 5 starting and ending in the F

state (e.g. FSFSF); see Table 4. All these regular patterns

were more frequent for Braeburn than for Fuji.

Discussion

In this study we proposed an integrative statistical model

that provides a global and quantitative picture of the two-

scale structuring observed during apple tree ontogeny: a

coarse scale corresponding to the succession of two

developmental phases and a fine scale corresponding to the

alternation between flowering and vegetative GUs (Fig. 7).

The first phase, which is almost transient (a phase is said to

be transient if when leaving it, it is impossible to return to

it), was called the ‘‘adolescent’’ phase; see below for dis-

cussion of the chosen terminology. This corresponds to the

alternation between long and flowering GUs. The second

phase, hereafter referred to as the ‘‘adult’’ phase, includes

patterns of alternation between medium and flowering GUs

and between short and flowering GUs. This structuring is a

consequence of both the inclusion of second-order mem-

ories that specialize flowering GUs as a function of the

preceding vegetative GU, and the one-step integrative

statistical modelling. In contrast, with a two-step modelling

where long and medium GUs were defined on the basis of a

threshold on the number of nodes before modelling the GU

succession, the first transient phase did not emerge; the

results were similar with a model based on the GU length

instead of the number of nodes (results not shown). The

chosen modelling approach also led to an unexpected

characterization of the extension GUs in the adolescent and

adult phases since differentiation between the long and

medium GUs combined the number of nodes with struc-

tural properties:

• GUs with\15 nodes could be either labelled as long or

medium (but the proportion of long GUs increases with

the number of nodes; see the corresponding mixture of

observation distributions in Fig. 4) while GUs with

more than 15 nodes were systematically labelled as

long.

• Vegetative polycyclism was frequent for long GUs

whereas it was rare for medium GUs,

Table 4 Counts for frequent patterns* starting and ending in F extracted from the state sequences restored using the hidden variable-order

Markov chains estimated for the Braeburn and Fuji cultivars

3 GUs Braeburn Fuji 4 GUs Braeburn Fuji 5 GUs Braeburn Fuji

FSF 935 159 FSSF 5 69 FSFSF 281 16

FMF 397 155 FLLF 48 30 FMFMF 65 18

FLF 122 73 FSMF 19 13 FLFMF 61 14

FMLF 0 13 FSFMF 60 10

FMFSF 40 3

FLFLF 29 13

Total 1 1,454 387 90 147 553 93

Total 2 3,188 2,173 2,127 1,064 1,330 503

The cumulated counts for patterns of a given length starting and ending in F (Total 1) and for patterns of similar length (Total 2) are given. L, M,

S, F stands for long, medium, short and flowering, respectively

* (i.e. more than ten occurrences for at least one of the two cultivars)

Fig. 7 Schematic representation of apple tree ontogenetic gradients

for Braeburn (left part) and Fuji (right part). For both cultivars, the

gradual decrease in shoot growth with tree ageing results from a

reduction in neoformation which involves growth cessation during

annual growth cycle and generates polycyclic shoots during the

adolescent phase (in grey). Later, the reduction leads to transitions

towards medium and short GUs. The adult phase (in white)

corresponds to repetitions of short and medium GUs separated by

flowering GUs with specific repeated patterns depending on the

regular or alternate behaviour of the cultivar

876 Trees (2012) 26:865–879

123

Page 13: Deciphering the Ontogeny of a Sympodial Tree

• Transitions from a long GU (belonging to the adoles-

cent phase) to a short GU (belonging to the adult phase)

with an intermediate flowering GU were rare while

transition from a medium GU to a short GU (both

belonging to the adult phase) with an intermediate

flowering GU were frequent.

Both studied cultivars exhibited a similar macroscopic

structure with two successive phases. This demonstrates the

similarity of architectural organisation at the species level,

even though within-species differences also existed at the

growth unit scale (regular spring flowering for Braeburn vs.

biennial bearing for Fuji). The almost transient character of

the first phase is an expected biological result since it

includes the vegetative period that precedes first flowering

occurrences. However, first flowering represented by the LF

memory was also included in this phase. The increase in

annual shoot length observed during the first years of

development of forest trees (Guedon et al. 2007) and other

morphological changes that often occur during the first

phases of tree ontogeny (Nozeran 1984) were not observed in

our apple tree dataset. This results from the grafting of

mature material on rootstocks that is commonly employed in

fruit tree cultivation to promote early flowering (Hackett

1985). As a consequence, we considered that no sensu stricto

‘‘juvenile phase’’ was observed in the present study and that

the first identified phase should rather be qualified as ‘‘ado-

lescent’’, with reference to the terminology introduced by

Day et al. (1997). Moreover, axis bending was performed in

orchards to accelerate the occurrence of flowering (Lauri

2002). This agronomic manipulation may have enhanced the

occurrence of the first flowering and reduced the adolescent

phase. We expect that different conditions will mainly affect

the time spent in each phase as given by the transition dis-

tributions attached to the memories LF and MF (i.e. different

balances between pLFL and pLFM, and between pMFM and pMFS)

but not the main structure of the models.

One of the main outputs of the proposed modelling

approach is a characterisation of the two successive

developmental phases in terms of GU types and suc-

cession. The adolescent phase is characterized by the

occurrence of long GUs and includes GUs belonging to

polycyclic vegetative shoots. In the estimated model,

polycyclism distinguishes long from medium GUs. It can

therefore be used as a new criterion for classifying GUs

and could contribute to overcome the ambiguity between

GU categories on the basis of their sole number of

nodes. Polycyclism also appears to be an intermediate

developmental stage in the transition towards the adult

phase. This particular stage relies on the capability of the

shoot apical meristem, specific to perennial species, to

stop temporarily its organogenetic activity without

entering into floral differentiation. From an evolutionary

point of view, a propensity for polycyclism has been

interpreted as a strategy to maximize light interception

(Verdu and Climent 2007), while, from a physiological

point of view, it has been interpreted as resulting from

nutritional competition between organs (Barnola et al.

1990). Our results show that the propensity of a species

to develop polycyclic shoots can vary intra-specifically

since it differed here between the two studied cultivars.

While the repetitive nature of plant growth has been

described for a long time at the metamer scale (White

1979), the present study highlights the existence of repe-

ated patterns at more macroscopic scales that result from

the repetitive occurrence of flowering GUs in a perennial

plant. In the proposed model, flowering occurrences mark

transitions not only between the adolescent and adult

phases, but also between stages within the adult phase (i.e.

between stages characterized by medium and short vege-

tative GUs). Transitions due to flowering have in the past

been considered as abrupt, especially when ending the

juvenile phase (Hackett 1985). Our results show that, in a

perennial plant, abrupt changes following flowering occur

throughout tree life. Flowering thus appears to be a marker

of the transition between successive developmental stages

that are represented by discrete states in the proposed

model. The modelling here is consistent with the high

structuring property of GU types compared with the small

differences between years in the number of nodes they bear

(except for very long GUs that develop in the first 2 years

of growth). It is noteworthy that decrease in growth with

tree ageing is particularly rapid in the apple tree mainly

because of manipulation effects (rootstock and bending).

This contrasts with forest trees where long periods of sta-

bility have been observed over decades separated by abrupt

changes (Guedon et al. 2007). In addition, a minimal and

recurrent stage, represented by the SF memory, was

reached, corresponding to a ‘‘minimal architectural unit’’

(Barthelemy and Caraglio 2007). In our dataset, flowering

occurrence after a minimum number of nodes can also be

interpreted with respect to alternate flowering which is a

specificity of perennial plants resulting from the absence of

flowering induction in particular years. The particular sta-

tus of flowering in tree ontogeny may arise from the high

carbon cost of flower and fruit production (Bustan and

Goldschmidt 1998) and competition with current vegeta-

tive growth (Cannell 1985; Berman and DeJong 1997). The

terminal position of flowering and subsequent sympodial

branching may also be involved in the reduction in vege-

tative growth since they induce a rupture in vascular con-

nections and probably in hydraulic conductance, and this is

assumed to be responsible for age-related decline in trees

(Martinez-Vilalta et al. 2007).

From a methodological point of view, the present

study paves the way to further investigations. First-order

Trees (2012) 26:865–879 877

123

Page 14: Deciphering the Ontogeny of a Sympodial Tree

Markov chains were applied in two other studies to

analyse successions of plant entities based on data col-

lected during follow-up that lasted several years (Mail-

lette 1990; Sterck et al. 2003). For instance, in

Maillette’s work on mountain white birches, the states

are long-shoot production, short-shoot production, sum-

mer dormancy (ecodormancy) and death. The states

identified in our context can be viewed as a retrospective

analogue of the meristem states defined by Maillette in a

prospective framework except for summer dormancy

which was not included in our modelling. In our

approach, some states (long and medium) were deduced

from an optimization based both on measurements

(number of nodes) and the properties of the succession

of GUs along the sequences. The focus was also dif-

ferent since our main objective was to identify the rules

governing the succession of differentiation states along

GU sequences while in Maillette (1990) and Sterck et al.

(2003), the aim was rather to characterise the demogra-

phy of plant entities during development. Hence, an

interesting avenue for future investigations would be to

combine the two approaches and study the demography

of plant entities based on sophisticated stochastic models.

In the present analysis, the trees were explored

exclusively through terminal transitions along either

monopodial or sympodial shoots. In contrast, transitions

resulting from lateral branching were modelled only

globally by the estimated initial probabilities of the

hidden variable-order Markov chains. But, the branching

structures resulting from lateral transitions along long

and medium GUs have been widely explored in previous

studies using mainly hidden semi-Markov chains; see

Guedon et al. (2001), Renton et al. (2006) and references

therein. A simulation system based on a multiscale sto-

chastic model combining a first-order Markov chain for

modelling GU succession and hidden semi-Markov

chains for modelling GU branching structure was pro-

posed by Costes et al. (2008); see also Lopez et al.

(2008) for a similar approach in the peach tree case. A

direct output of the present study would be to replace the

first-order Markov chain by a variable-order Markov

chain to better model growth phases and repeated pat-

terns in GU succession. Such simulation systems com-

bining sophisticated stochastic models for modelling tree

topology with different types of mechanistic models (e.g.

mechanical model for branch bending or model for

carbohydrate partitioning) open new perspectives for in

silico investigations of agronomical scenarios. Similar

approaches could be applied to other perennial species,

including monopodial and sympodial species.

Acknowledgments This research was funded by both the INRA

Genetic and Breeding Department and the CIRAD Bios Department.

We thank Michael Renton for his contribution in the first steps of GU

analyses and Pierre-Eric Lauri for helpful comments.

Appendix A: Selection of the memories of a variable-

order Markov chain

The order of a Markov chain can be estimated using the

Bayesian information criterion (BIC). For each possible

order r, the following quantity is computed

BIC rð Þ ¼ 2 log Lr � Jr J � 1ð Þ log n; ð1Þ

where Lr is the likelihood of the rth-order estimated Mar-

kov chain for the observed sequences, Jr(J - 1) is the

number of independent transition probabilities of a J-state

rth-order Markov chain and n is the cumulated length of

the observed sequences. The principle of this penalized

likelihood criterion consists in making a trade-off between

an adequate fitting of the model to the data [given by the

first term in (1)] and a reasonable number of parameters to

be estimated (control by the second term, the penalty term).

In practice, it is infeasible to compute a BIC value for each

possible variable-order Markov chain of maximum order

r B R since the number of hypothetical memory trees is

very large. An initial maximal memory tree is thus built

combining criteria relative to the maximum order and to

the minimum count of memory occurrences in the observed

sequences. This memory tree is then pruned, using a two-

pass algorithm which is an adaptation of the Context-tree

maximizing algorithm (Csiszar and Talata 2006): a first

dynamic programming pass, starting from the terminal

vertices and progressing towards the root vertex for com-

puting the maximum BIC value attached to each sub-tree

rooted in a given vertex, is followed by a second tracking

pass starting from the root vertex and progressing towards

the terminal vertices for building the memory tree.

Appendix B: Definition of parametric observation

distributions for hidden variable-order Markov chains

In the case of a count variable such as the number of nodes

of a GU, the observation distributions are parametric dis-

crete distributions chosen from among binomial distribu-

tions, Poisson distributions and negative binomial

distributions with an additional shift parameter d.

The binomial distribution with parameters d, n and

p (q = 1 - p), B(d, n, p) where 0 B p B 1, is defined by

bj yð Þ ¼ n� dy� d

� �py�dqn�y; y ¼ d; d þ 1; . . .; n:

The Poisson distribution with parameters d and k, P(d,

k), where k is a real number (k[ 0), is defined by

878 Trees (2012) 26:865–879

123

Page 15: Deciphering the Ontogeny of a Sympodial Tree

bj yð Þ ¼ e�kky�d

y� dð Þ! ; y ¼ d; d þ 1; . . .

The negative binomial distribution with parameters d, r

and p, NB(d, r, p), where r is a real number (r [ 0) and

0 \ p B 1, is defined by

bj yð Þ ¼ y� d þ r � 1

r � 1

� �prqy�d; y ¼ d; d þ 1; . . .

References

Bangerth KF (2009) Floral induction in mature, perennial angiosperm

fruit trees: Similarities and discrepancies with annual/biennial

plants and the involvement of plant hormones. Sci Hortic

122:153–163

Barnola P, Alatou D, Lacointe A, Lavarenne S (1990) Biological and

biochemical study of factors determining rhythmic growth of

oak (Quercus robur). Effects of leaf removal. Ann Sci Forest

47:619–631

Barthelemy D, Caraglio Y (2007) Plant architecture: a dynamic,

multilevel and comprehensive approach to plant form, structure

and ontogeny. Ann Bot 99:375–407

Berman ME, DeJong TM (1997) Crop load and water stress effects on

daily stem growth in peach (Prunus persica). Tree Physiol

17:467–472

Buhlmann P, Wyner AJ (1999) Variable length Markov chains. Ann

Stat 27:480–513

Bustan A, Goldschmidt EE (1998) Estimating the cost of flowering in

a grapefruit tree. Plant Cell Environ 21:217–224

Cannell MGR (1985) Dry matter partitioning in tree crops. In:

Cannell MGR, Jackson JE (eds) Attributes of trees as crop

plants. Institute of Terrestrial Ecology, UK, pp 160–193

Chaubert-Pereira F, Caraglio Y, Lavergne C, Guedon Y (2009)

Identifying ontogenetic, environmental and individual compo-

nents of forest tree growth. Ann Bot 104:883–896

Costes E, Sinoquet H, Kelner JJ, Godin C (2003) Exploring within-

tree architectural development of two apple tree cultivars over

6 years. Ann Bot 91:91–104

Costes E, Smith C, Renton M, Guedon Y, Prusinkiewivz P, Godin C

(2008) Mapple T: simulation of apple tree development using

mixed statistical and biomechanical models. Funct Plant Biol

35:936–950

Crabbe J, Escobedo-Alvarez JA (1991) Activites meristematiques et

cadre temporel assurant la transformation florale des bourgeons

chez le Pommier (Malus x Domestica Borkh., cv. Golden

Delicious). In: L’Arbre: Biologie et developpement. NaturaliaMonspeliensia, pp 369–379

Csiszar I, Talata Z (2006) Context tree estimation for not necessarily

finite memory processes, via BIC and MDL. IEEE Trans Inf

Theory 52:1007–1016

Day JS, Gould KS, Jameson PE (1997) Vegetative architecture of

Eleocarpus hookerianus. Transition from juvenile to adult. Ann

Bot 79:617–624

Durand JB, Guedon Y, Caraglio Y, Costes E (2005) Analysis of the

plant architecture via tree-structured statistical models: the

hidden Markov trees. New Phytol 166:813–825

Ephraim Y, Merhav N (2002) Hidden Markov processes. IEEE Trans

Inf Theory 48:1518–1569

Grenander U, Miller MI (2006) Pattern Theory, from representation to

inference. Oxford University Press, Oxford

Guedon Y, Barthelemy D, Caraglio Y, Costes E (2001) Pattern

analysis in branching and axillary flowering sequences. J Theor

Biol 212:481–520

Guedon Y, Caraglio Y, Heuret P, Lebarbier E, Meredieu C (2007)

Analyzing growth components in trees. J Theor Biol 248:418–

447

Hackett WP (1985) Juvenility, maturation and rejuvenation in woody

plants. In: Janick J (ed) Horticultural Reviews 7. pp 109–115

Halle F, Oldeman RA, Tomlinson PB (1978) Tropical trees and

forests. Springer, Berlin

Kass RE, Raftery AE (1995) Bayes factors. J Am Stat Assoc 90:773–

795

Lauri PE (2002) From tree architecture to tree training––An overview

of recent concepts developed in apple in France. J Korean Soc

Hortic Sci 43(6):782–788

Lauri PE, Terouanne E, Lespinasse JM, Regnard J-L, Kelner JJ

(1995) Genotypic differences in the axillary bud growth and

fruiting pattern of apple fruiting branches over several years—an

approach to regulation of fruit bearing. Sci Hortic 64:265–281

Lopez G, Favreau R, Smith C, Costes E, Prusinkiewicz P, DeJong TM

(2008) Integrating simulation of architectural development and

source-sink behaviour of peach trees by incorporating Markov

models and physiological organ functions sub-models into

L-Peach. Funct Plant Biol 35:761–771

Maillette T (1990) The value of meristems states as estimated by a

discrete Markov-chain. Oıkos 59:235–240

Martinez-Vilalta J, Vanderklein D, Mencuccini M (2007) Tree height

and age-related decline in growth in Scots pine (Pinus sylvestrisL.). Ecologia 150:529–544

Nozeran R (1984) Integration of organismal development. In: Barlow

PW, Carr DJ (eds) Positional controls in Plant Development.

Cambridge University Press, Cambridge, pp 375–401

Renton M, Guedon Y, Godin C, Costes E (2006) Similarities and

gradients in growth-unit branching patterns during ontogeny in

‘Fuji’ apple trees: a stochastic approach. J Exp Bot 57:3131–

3143

Ron D, Singer Y, Tishby N (1996) The power of amnesia: Learning

probabilistic automata with variable memory length. Mach Learn

25:117–149

Sabatier S, Barthelemy D (1999) Growth dynamics and morphology

of annual shoots, according to their architectural position, in

young Cedrus atlantica (Endl.) Manetti ex Carriere (Pinaceae).

Ann Bot 84:387–392

Solar A, Stampar F (2006) Stability of the annual shoot diameter in

Persian walnut: a case study of different morphotypes and years.

Trees 20:449–459

Sterck FJ, Bongers F, During HJ, Martinez-Ramos M, De Kroon H

(2003) Module responses in a tropical forest tree analyzed with a

matrix model. Ecology 84:2751–2761

Suzuki A (2002) Influence of shoot architectural position on shoot

growth and branching patterns in Cleyera japonica. Tree Physiol

22:885–890

Verdu M, Climent J (2007) Evolutionary correlations of polycyclic

shoot growth in Acer (Sapindaceae). Am J Bot 94:1316–1320

White J (1979) The plant as a metapopulation. Annu Rev Ecol Syst

10:109–145

Wilkie JD, Margaret S, Olesen T (2008) Regulation of floral initiation

in horticultural trees. J Exp Bot 59:3215–3228

Trees (2012) 26:865–879 879

123