You are on page 1of 12

Neuroscience and Biobehavioral Reviews 81 (2017) 213–224

Contents lists available at ScienceDirect

Neuroscience and Biobehavioral Reviews


journal homepage: www.elsevier.com/locate/neubiorev

Review article

A revival of Homo loquens as a builder of labeled structures:


Neurocognitive considerations
T. Goucha a,b,∗ , E. Zaccarella a , A.D. Friederici a,b
a
Max Planck Institute for Human Cognitive and Brain Sciences, Department of Neuropsychology, Leipzig, Germany
b
Berlin School of Mind and Brain, Humboldt University, Germany

a r t i c l e i n f o a b s t r a c t

Article history: The core capacity of human language is described as the faculty to combine words into hierarchical struc-
Received 2 March 2016 tures. This review aims to isolate the fundamental computation behind the language faculty together
Received in revised form 24 January 2017 with its neural implementation. First, we present our central hypothesis by confronting recent linguis-
Accepted 26 January 2017
tic theory with evolutionary arguments: linguistic humaniqueness is reflected in the labeling of word
Available online 17 March 2017
combinations forming asymmetric hierarchical structures. Second, we review the neurolinguistic litera-
ture, especially focusing on dual-stream connectivity models. We put forward that the dorsal pathway,
Keywords:
especially the arcuate fascicle, is responsible for the rule-based combinatorial system, implementing
Labeling
Language evolution labeling and giving rise to hierarchical structures. Conversely, the ventral stream is rather responsible for
Syntax semantic associative operations. We further present evolutionary neuroanatomical evidence grounding
Hierarchical processing our hypothesis. We conclude by suggesting further avenues of research as well as open questions to be
Arcuate fascicle addressed. With the aim to expand our knowledge on the neurobiology of language, we hope to provide
a testable hypothesis for the origin of language syntax bringing together evidence from different fields.
© 2017 Elsevier Ltd. All rights reserved.

Contents

1. Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 213
2. The Merge-only hypothesis and why we need labels . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 214
3. Neural implementation of the language faculty . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 215
3.1. From hierarchical processing in the prefrontal cortex to the language network . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 215
3.2. The neural implementation of labeling in the language network . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 217
3.3. Integrating the hypothesis with evolutionary evidence in neuroanatomy . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 219
3.4. Implications for future research: Which predictions can we make? . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 220
4. Conclusions: Is there really something special about language? . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 222
References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 222

1. Introduction
support the emergence of language as a species-specific faculty.
Language is as a powerful communicative tool which enables While this quest has stimulated psycholinguistic and neurolin-
humans to exchange complex states of affairs with each other. The guistic studies (Friederici and Singer, 2015), comparative research
language system constitutes a rather recent evolutionary innova- has also provided important insights about the commonalities
tion, and appears to be unique to the human species. Given this and differences between humans and nonhuman primates (NHPs)
uniqueness, a central aim for current cognitive research is to iso- (Berwick et al., 2013) In parallel, linguistic theory has accord-
late some specific trait in humans, which would most plausibly ingly addressed the new challenges arising from the findings of
other fields (Bolhuis et al., 2014; Everaert et al., 2015). How-
ever, these attempts to isolate the language faculty are yet to
∗ Corresponding author at: Max Planck Institute for Human Cognitive and Brain achieve success. This apparent insufficiency might arise from the
Sciences, Stephanstraße 1A, 04103, Leipzig, Germany. difficulty to integrate insights from these very disparate disci-
E-mail address: goucha@cbs.mpg.de (T. Goucha). plines. Here, we try to overcome these shortcomings and provide

http://dx.doi.org/10.1016/j.neubiorev.2017.01.036
0149-7634/© 2017 Elsevier Ltd. All rights reserved.
214 T. Goucha et al. / Neuroscience and Biobehavioral Reviews 81 (2017) 213–224

a testable framework based on data collected by different disci- the combinatorial mechanism that brings two elements together
plines, always bearing in mind a coherent big picture. The present to form an unordered set (Chomsky, 2013), as in:
framework will especially draw on evidence from cognitive neu-
␣ˇ → {˛ˇ} (1)
roscience, which will accompany the argumentation along this
paper. The expression above can be read as “take two elements ␣
A multidisciplinary approach to the question of language evo- and ␤, and string them together to form a new set containing
lution and its brain implementation is the most adequate strategy both”. Merge can create a new set including words from any lex-
to provide us with helpful empirically testable frameworks that ical categories (e.g., nouns, verbs, determiners, which correspond
can contribute to the progress of the field of biolinguistics (Boeckx, to terminal nodes), or sets that had already been formed by Merge
2013; Bickerton, 2014b). A clear advantage of such a multidisci- itself. This possibility of a computation to take its own output as an
plinary perspective is the possibility to combine the constraints of input again is what defines recursion, a cornerstone in the discus-
the different fields to reduce the scope of hypotheses in considera- sion about the language faculty (Hauser et al., 2002). For the output
tion (Fitch, 2014). If we consider the language faculty to be a recent of this computation to be correctly interpreted at the interfaces
adaptation of our species, there are several aspects we must take (either the sensorimotor interface or the conceptual intentional
into account. First of all, the fundamental language computation in interface), the issuing sets need to be labeled. The label of the new
focus should not only be in agreement with linguistic theory, but set is attributed according to the labels of its elements and rela-
also with comparative cognitive research. This evolutionary differ- tionship established between their categories. Labeling determines
ence should thus constitute a cognitive novelty in our species, in a local asymmetry between the two items in the set, since one of
particular the cognitive process in question should not be present the two elements assigns its label (let it be ␣) to the new merged
in our closest relatives. Otherwise, it would be difficult to explain object, in this case the label ␥␣ (2):
how language developed exclusively in humans. Furthermore, we
␣␤ → ␥␣ {␣␤} (2)
should carefully consider the neural implementation of this cogni-
tive difference that originated language. This means that we must In linguistic theory, the element assigning the label to the
be able to provide a plausible neuroanatomical substrate for this merged object is said to project and it is called the selector or head
computation that shows some novelty in humans in comparison of the structure formed by the two merged elements. Narita (2014)
with NHPs. addresses how the labeling of the set resulting from Merge is closely
A striking fact about language is the possibility to combine related to the categories of the elements of the set. In order to illus-
words in a way that new relationships are established between trate how labeling arises, let us consider the example of one of
them, and novel meanings arise, being conveyed by the way they the most essential asymmetries between categories in language –
are combined. Because of this, a fine-grained description of the the asymmetry between nouns and verbs. In this case, the verb (V,
algorithm holding this combinatorial process takes on a crucial “eat”) is usually the selector, taking the noun (N, “apples”) as its
role. One of the dominating views in the field is the hypothe- object. The verb hence imposes its label onto the other component
sis of merge being the crucial computation, put forward by the with which it was merged, e.g. a verb phrase (VP, “eat apples”).
linguist Noam Chomsky (Chomsky, 2007; Bolhuis et al., 2014; In more abstract terms, this shows different elements in language
Everaert et al., 2015). In this framework, the language faculty play different roles when combines. Some elements in language
consists of a single computation—Merge—that combines two ele- are rather functional and dynamic, and usually refer to events.
ments and gives rise to an unordered symmetrical set. Merge is These establish the relationships between elements in a phrase and
therefore considered the basic hierarchy building computation that between phrases in a sentence, thus determining the hierarchical
enabled humans to master language within the Minimalist Pro- structure. In turn, other elements are mainly content-bearing and
gramme (Berwick and Chomsky, 2011; Bolhuis et al., 2014; but static, and rather refer to sorts. The constituents involved in (2)
see also Boeckx and Benítez-Burraco, 2014). While we support operate accordingly. The head ␣ stands in an asymmetrical rela-
the idea of a simple and parsimonious combinatorial computa- tionship with its complement ␤, based on the internal grammatical
tion at the basis of language, we will nonetheless try to elucidate relationship between the two elements. The same analysis applies
why Merge might need expansion in order to provide explana- to the phrase “which apples”. Formally, the concrete bare syntac-
tory adequacy in both cognitive and evolutionary domains. Here, tic set {which, apple} resulting from the Merge of “which” and
we put forward that the labeling of the outcome of the opera- “apple” is labeled Determiner Phrase (DP) because it is a property
tion Merge is a necessary cognitive prerequisite for a complete of the determiner (D) to require a noun (N) as a complement (3).
account of the uniqueness of human language processing. It is Therefore, the label received by the newly formed set expresses
through labeling that asymmetric hierarchical structures can orig- the asymmetric hierarchical nature of this same set. This happens
inate, thus distinguishing language from other communication because one of its elements, the determiner (D), projects its label
systems (Murphy, 2015a, 2015b). In the following sections we to the newly formed phrase, a determiner phrase (DP):
will start by explaining our proposal from a language theoretical
which, apple → DP {whichapple} (3)
perspective. We will then integrate the hypothesis with current
neuroscientific findings and discuss its concrete neural implemen- In the Minimalist Program, for the sake of theoretical ele-
tation. Subsequently we will provide comparative data in support gance, Merge is kept rather undefined. Consequently, the labeling
of our view. In the final part, we will suggest possible avenues for algorithm is excluded from the basic language computations and
future research. labels are only established for interface requirements (Berwick
and Chomsky, 2016). The definition of Merge is therefore left open
regarding both the properties of its inputs (e.g., its category) and
2. The Merge-only hypothesis and why we need labels of the structure that constitutes its output (e.g., which label the
merged set takes). This, however, appears to conflict with the
We start by discussing the assumption that the evolutionary desired explanatory adequacy of the language faculty from a biolin-
step that allowed humans to acquire language is a novel combina- guistic perspective. In particular, it does not seem to be able to
torial faculty that generates hierarchical structures unique to the directly account for the evolutionary mechanisms of the emer-
human communication system. At the center of the Minimalist Pro- gence of language. When considering evolution, we must equate
gram (Chomsky, 1995) stands the computation Merge, defined as which could have been the difference to our ancestors, which could
T. Goucha et al. / Neuroscience and Biobehavioral Reviews 81 (2017) 213–224 215

eventually have led to the fixation of that specific trait. The human Only this way is it possible to go back and forth between a hier-
language has the advantage of allowing individuals to exchange archical mental structure and a linear string produced in vocal
information about complex affairs and mental states. Humans are communication. According to Berwick et al. (2013) the interchange
able transmit combinations of meaningful units that establish rela- between the mental structure and the linear string is supported by
tions of dominance between themselves, i.e. form hierarchical the sensory-motor interface. Within the Minimalist Program, label-
structures. Yet, in most cases of animal communication, the chan- ing plays a secondary role because it constitutes an additional step
nel only allows information to be transmitted in the form of linear in combining two lexical items. In this account, labeling is only
strings (i.e., sequences where one element follows the previous one needed in language externalization, playing an interpretive role at
without any apparent hierarchical structure). Without any further the interfaces and providing the basis for the linearization (Everaert
information about a generative rule, a communication channel only et al., 2015; but see Hornstein 2009 for a different account). How-
allowing linear strings makes it difficult to transmit information ever, with the aforementioned performative advantage, we put
organized in a hierarchical fashion. However, in humans, hierar- labeling forward as plausibly contributing for the emergence of
chical structures can be conveyed due to successful linearization. language along with Merge.
This is achieved by flattening hierarchical structures during produc- Although we pinpoint labeling as a fundamental linguistic com-
tion or by building hierarchical structures from a string of words putation, labeling is indeed also necessary for more superficial
during perception. This process constitutes one of the most cru- linguistic phenomena, like word order. Most languages exhibit a
cial issues in linguistic theory, at least since the seminal work of preferred word order between two elements in a phrase, called
Kayne (1994).1 In general, the presence of asymmetric structures headedness (Comrie, 1989), which determines whether the head or
in language has long been debated from a biolinguistic perspec- the complement come first in a linear order. We do agree that lin-
tive (Sciullo and Jenkins, 2016). Kayne’s view was later extended to earization is a mere requirement of the material constraints of the
include the generation of structures exhibiting points of symme- sensorimotor interface (i.e., the necessity to convey one element at
try in the course of derivation, for which a linearization ordering a time, hence producing a string). But note that the same labeling
cannot be determined (Dynamic Antisymmetry; Moro, 2000).2 can give rise to different linear orders and the same linear orders can
Moving from these theoretical considerations, it is not straight- also correspond to a different labeling. An example of the former
forward to understand how Merge can account for the generation of would be languages with different headedness, which we would
a hierarchical structure in and of itself. Crucially, from an unordered see in the French and German translations of “blue car”, “voiture
set without labels it is not possible to retrieve any relation of dom- bleue” and “blaues Auto”, respectively with the noun correspond-
ination between the elements within the set, which would in turn ing to car in the first or last position, but without any consequences
be a necessary condition for hierarchical structure. Labeling consti- for interpretation. The latter can be illustrated by the two possibili-
tutes a crucial performative advantage for dealing with hierarchies, ties of interpretation of “green tea cup” as either {green, {tea, cup}}
especially for sharing them when hidden in linearized sequences. In or {{green, tea}, cup} (Fujita, 2014). Statistical learning could point
language use, participants share conventions concerning the label- a more likely interpretation, but only a fully hierarchical labeling
ing of the elements present in linear sequences. In particular, these procedure would allow the speaker to retrieve both interpretations.
conventions allow them to recognize which labels the single ele- This example illustrates how labeling is not only responsible for
ments carry and which combinatorial rules apply between them. surface phenomena, but is in fact a fundamental computation in
Specifically, they know which of the merged elements selects the language processing.
other, thus being the head of the new merged set at each com-
bination. Eventually, this allows the language user to retrieve the
hierarchical relationships between the elements in the sequence. 3. Neural implementation of the language faculty

3.1. From hierarchical processing in the prefrontal cortex to the


language network
1
According to Kayne’s Linear Correspondence Axiom (LCA), there is a direct map-
ping between hierarchical syntactic structure and the linear order in a sequence, Given the fundamental hierarchical nature of human language,
such that asymmetric relations between hierarchical nodes map onto precedence neurobiological investigations into the neural implementation of
relations. Formally, a lexical entry ␣ will linearly precede a lexical entry ␤ if the
corresponding non-terminal node of ␣ will asymmetrically c-command the corre-
the language faculty should be able to describe which brain areas
sponding non-terminal node of ␤. C-command, in turn, is defined as a structural and which neural circuits are consistently involved in the pro-
relation between two nodes, where one node is said to c-command another one in cessing of hierarchy. We know that the prefrontal cortex (PFC)
case the two nodes do not dominate each other, but the first node up in the tree that in general is involved in such processes regardless of the cogni-
branches dominates both of them (Reinhart, 1976). When both nodes c-command
tive domain (e.g. Koechlin et al., 1999). For the dorsolateral PFC,
each other, there is symmetric c-command. When one node c-commands the other,
but the latter does not c-command the former an asymmetric c-command config- there is an established rostrocaudal gradient regarding the degree
uration is established and linear ordering follows. The sentence John likes apples, of abstraction of mental representations (Badre, 2008), with more
for example, can be structurally represented with brackets’ notation (with reduced anterior areas being responsible for more abstract processing. This
formalism for simplicity) as [[John][likes [apples]]], where the node containing John gradient has been interpreted as reflecting a certain hierarchy of
asymmetrically c-commands the nodes containing likes and apples, while the node
containing likes asymmetrically c-commands the node containing apples. From this,
processing complexity between the subcomponents of the dorso-
it follows that John must precede likes, which in turn must precede apples. lateral PFC along this gradient (Badre and D’Esposito, 2009) as well
2
Mirror structures, like copular sentences in Italian for example (“Questa foto sul as the corresponding degree of automaticity of these processes in
muro è la causa della rivolta”, i.e., “This picture on the wall is the cause of the riot”) the opposite direction (Jeon and Friederici, 2013). Importantly, in
have been indeed suggested to represent such a case where two Noun Phrases (NPs;
the specific case of language processing, the more abstract complex
“the cause of the riot”, “this picture on the wall”) are first both included in the same
node, [è [questa foto sul muro, la causa della rivolta]]. According to Dynamic Anti- hierarchies seem to rather engage posterior areas of the inferior
symmetry, instability due to points of symmetry within a structure is then rescued frontal gyrus (IFG; Jeon and Friederici, 2015). In this respect, the
by the movement of one of the two NPs to a preverbal position. As such, if neither involvement of the IFG within the language system seems to be
NP is able to project a label, one of the two elements moves, forcing the other to governed by other principles of organization, more relevant for
project and establish the asymmetry (Moro, 2014). Similarly, movement – a perva-
sive phenomenon of language – is taken to be the computational source preserving
the language dimensions, like phonology, syntax or semantics (e.g.,
phrase structure asymmetries when symmetric structures occur and linearization Uddén and Bahlmann, 2012). That’s why we will henceforth focus
cannot be determined. our attention in the ventral part of the PFC and its subcomponents.
216 T. Goucha et al. / Neuroscience and Biobehavioral Reviews 81 (2017) 213–224

Fig. 1. Dual stream model of language processing. This model integrates data from functional and anatomical neuroimaging into two main dorsal pathways involved in
syntactic processing and articulation respectively and two main ventral pathways involved in local combinations and semantic processing in general.

In the IFG, we find both Broca’s area and the ventrally adjacent Whereas humans can process both grammars effortlessly, NHPs,
frontal operculum (FOP), and the anterior insula (Fig. 1). Broca’s in particular cotton-top tamarins, can only capture the rule of the
area is the part of the left IFG spanning between the precentral simpler FSG (Fitch and Hauser, 2004). Furthermore, the two differ-
sulcus and the horizontal branch of the sylvian fissure. Within ent kinds of grammars have differential patterns of brain activation
this area, there is a well-established gradient between more ante- in humans (Friederici et al., 2006). While violations of the FSG only
rior and more posterior areas (Hagoort, 2005), regarding function activates the evolutionarily older FOP, with a main ventral white
(Dapretto and Bookheimer, 1999), cytoarchitecture (Amunts et al., matter connection, the PSG additionally engages the lateral BA 44,
1999) and anatomical connectivity (Friederici, 2012). In terms of with a main dorsal white matter connection through the arcuate
brain function, the more posterior pars opercularis (roughly corre- fascicle. Not only violation, but in general the presence of the hier-
sponding to Brodmann Area (BA) 44) is involved syntax processing, archical rule engaged BA 44 (Bahlmann et al., 2008), even in the
as shown in meta-analytical studies (Hagoort and Indefrey, 2014; visual domain (Bahlmann et al., 2009). Other grammars based on
Rodd et al., 2015), literature reviews (Friederici, 2011), and patient several combinatorial rules, e.g. Reber grammars (Reber, 1967),
studies (Wilson et al., 2010). The more anterior pars triangularis were also investigated in both neuroimaging and brain stimula-
(BA 45) (often extending to pars orbitalis (BA 47)), in turn, is tion studies (for a review see Petersson and Hagoort, 2012), whose
rather engaged by sentence-level semantic processing (Friederici, main neural correlate was the activation of the left IFG. This was
2011; Zaccarella et al., 2015) and is systematically more activated true for studies involving sequence processing with an underly-
by semantic than syntactic tasks (Hagoort and Indefrey, 2014). ing generative rule, in particular for Reber grammars (Forkstam
Cytoarchitectonically, we can also distinguish two regions within et al., 2006), also showing a good overlap with natural language
Broca’s area with an approximate correspondence to the anatom- processing in this brain area (Petersson et al., 2012). Studies using
ically defined pars opercularis and pars triangularis. BA 44 shows brain stimulation also highlighted the importance of the IFG for the
a significantly less developed layer IV compared to the anteriorly processing of these artificial grammars, for instance in the detec-
located BA 45, being the formed classified as dysgranular and the tion of violations (Uddén et al., 2008), but also in underlining its
latter being properly granular. BA 44 also borders posteriorly with importance in the actual acquisition of the rules rather than the
the premotor BA 6, which is in turn agranular. (Brodmann, 1909; superficial familiarity (De Vries et al., 2010), also showing corre-
Amunts et al., 2010). lates in structural connectivity in the adjacent white matter (Flöel
Additionally, it is important to distinguish the dysgranular BA et al., 2009). In this sense, it seems that evolutionarily more recent
44 in the lateral cortical convexity from the adjacent FOP and the areas (i.e. the IFG) are involved in complex structure building, both
anterior dorsal insula (Sanides, 1962), which are agranular and in artificial grammars (Friederici et al., 2006) and natural language
evolutionarily older, lying ventrally and medially to Broca’s area. (Makuuchi et al., 2009), whereas areas shared with other mam-
These areas are not only cytoarchitectonically different, but also mals are rather involved in concatenation processes, which precede
play different roles in hierarchical processing. Functional stud- structure building (Zaccarella and Friederici, 2015a). In general, BA
ies on artificial grammar processing showed that BA 44 is rather 44 seems to be an area that is able to combine different levels of
engaged by more complex grammars with a hierarchical structure information for hierarchical building (Wang et al., 2015), instead
(e.g. Friederici et al., 2006; Opitz and Friederici, 2007), whereas of merely being responsible for the blind chunking of elements
more simple linear grammars only recruited the more ventral FOP regardless of their internal structure (Clerget et al., 2009).
(e.g. Friederici et al., 2006). This difference between two kinds of Even though it is possible to establish a certain functional seg-
grammars was explored in the literature especially through the regation in terms of cortical areas, the specificity of the processes
distinction between so-called phrase structure grammars (PSG) performed in a certain brain area strongly depend on its con-
and finite state grammars (FSG). PSG allow for center embedded nectivity. We must therefore distinguish general computations
structures with long distance dependencies, FSG are based on local performed by certain area from the specific function it performs
transitions. By distilling down the combinatorial rule to a very sim- within a certain network. Long range white matter fiber bundles
ple pairing of items of two classes A and B, PSG issues structures connect Broca’s area with temporal regions in the left hemisphere.
of the generalized form An Bn , whilst FSG generate sequences of First, we must distinguish the dorsal fiber tracts that connect the
the type (AB)n . Studies with these very simplified grammars were posterior part of the IFG – pars opercularis (BA 44) and the ven-
very useful to set the frame in terms of evolutionary differences tral precentral gyrus (BA 6v) – from the ventral fiber tracts that
in sequence learning and also regarding its brain implementation. connect both the more anterior IFG – pars triangularis (BA 45) and
T. Goucha et al. / Neuroscience and Biobehavioral Reviews 81 (2017) 213–224 217

orbitalis (BA 47) – and also the more ventral FOP (Friederici et al., information, there is a coordination of the frontal and temporal cor-
2006; Anwander et al., 2006; Friederici, 2012). Recent models (e.g. tex for categorization (Freedman et al., 2003), although the storage
Friederici and Gierhan, 2013) propose the existence of two dorsal aspect is a function of the temporal cortex (Szczepanski and Knight,
and two ventral fiber bundles displayed schematically on Fig. 1. 2014). In the linguistic domain, the storage of lexical items seems
Dorsally, part of the superior longitudinal fascicle (SLF) connects to be rather of function of the temporal lobe, which are then inte-
the superior (STG) and middle temporal gyri (MTG) to the premo- grated along the ventral stream (DeWitt and Rauschecker, 2012).
tor cortex via the parietal cortex (Catani and Jones, 2005; Makris Specific manipulations involving the retrieval of lexical items from
et al., 2005), associated with auditory-motor integration (Frey et al., the mental lexicon engage the posterior ventral temporal cortex
2008; Saur et al., 2008). Also belonging to the dorsal pathway, the (Lau et al., 2008). Besides, the posterior middle and inferior tem-
arcuate fascicle (AF) directly links BA 44 to the posterior STG/MTG. poral cortices are significantly more recruited by lexical semantics
This structure was singled out to play a central role in process- than syntax processing in a meta-analysis (Hagoort and Indefrey,
ing syntactically complex sentences (Friederici, 2009; Wilson et al., 2014). When linguistic information is confined to word semantics,
2011; Brauer et al., 2011), but can also be more generally seen as a the brain activations only engage the ventral stream (Graves et al.,
crucial structure for predicting the incoming input and its under- 2007). Indeed, disorders affecting temporal regions anterior to the
lying structure (Friederici, 2012), and to play an important role in primary auditory cortex, i.e. along the ventral stream, spare the pro-
the integration of syntactic and semantic information (den Ouden duction of abstract sentence structure, whereas the word retrieval
et al., 2012), with the posterior STG/MTG being a pivotal area for the and also the use of words in context are severely altered (Mesulam
integration of different sources of information (Friederici, 2011). et al., 2015). In turn, if we are tapping into merely structural pro-
Ventrally, the extreme fiber capsule system (EFCS) includes the cessing, independently of content, the brain activations are rather
inferior fronto-occipital fascicle (IFOF) connecting BA 45/47 to the circumscribed to BA 44, which is the case of studies using syntac-
temporal cortex, involved in sound-meaning mapping (Saur et al., tic manipulations controlling for working memory like scrambling
2008; Rolheiser et al., 2011; Griffiths et al., 2013). Additionally, (Meyer et al., 2012) or embedding (Makuuchi et al., 2009), or stud-
the uncinate fascicle (UF) connects the most ventral part of the ies using pseudowords (Röder et al., 2002; Goucha and Friederici,
prefrontal cortex, including the FOP, with the anterior temporal 2015). Its neighboring BA 45, mostly involved in sentence level
lobe (ATL), and has been seen as important for local concatena- semantic processing, rather comes into play when the syntactic
tions (Zaccarella and Friederici, 2015a) and adjacent dependencies manipulation has additional semantic demands, e.g. in the case
(Friederici et al., 2006). Also for artificial grammar learning, the of anaphora in movement (Santi and Grodzinsky, 2010). Regard-
functional segregation discussed earlier corresponded to a differ- ing interactions between syntax and semantics, which is especially
ence in anatomical connectivity: The activation peak of grammars relevant in thematic role assignment, the posterior superior tem-
with local transitions – FSG – had a dominant ventral white matter poral cortex comes into play (Friederici et al., 2003; Bornkessel
connection whereas the activation peak of hierarchical grammars – et al., 2005), highlighting its important role in the integration of
PSG – had a clear dorsal connection. Crucially, the functions of these the two processing streams. In terms of white matter connectivity,
white matter pathways show an overall correspondence with the we find a dominance of the dorsal stream in the processing of com-
functional segregation of the brain areas they connect, especially plex hierarchical structures, which is reflected in brain anatomy
in the IFG. (Skeide et al., 2015) and in effective connectivity (den Ouden et al.,
2012), whereas the ventral connections rather account for seman-
3.2. The neural implementation of labeling in the language tic aspects (Rolheiser et al., 2011). In fact, the fundamental role
network of the arcuate fascicle has already been stressed in the context
of language evolution (Friederici, 2012; Berwick et al., 2013), lan-
As already mentioned, an attempt to fully understand cognition guage acquisition (Perani et al., 2011; Skeide et al., 2015) and its
from a neural perspective requires us to depart from localizationist anatomical uniqueness in humans (Rilling et al., 2008). However,
approaches. We must instead appreciate how brain areas are inte- here we extend the interpretation of the functional role of the arcu-
grated into circuits, to which we can assign more circumscribed ate fascicle to the crucial connection between two brain areas with
computations, by combining the specificities of the areas involved. two essentially different kinds of computation, whose crosstalk is
In the particular case of labeling, its implementation presupposes fundamental for building language structures.
that the operation of hierarchical structure building can access the In the same way that the neural implementation of language
categories of the elements that are being combined. For this pur- can be seen as dichotomy between two processing streams, we can
pose, the cross-talk between hierarchical structure building and also discern two sides in the language’s building blocks − words.
categorization is mandatory. In the language network, the PFC, in On the one hand, words possess a categorical dimension that is
particular the posterior IFG, is largely involved in hierarchical pro- related to the syntactic rules used in hierarchal structure building
cessing of linguistic structures (Fitch and Martins, 2014; Zaccarella and, on the other hand, words contain a conceptual dimension that
et al., 2015), whereas the temporal lobe is engaged in categoriza- is related to their content. The rules that guide the combination
tion processes ranging from phonetics (e.g. Overath et al., 2015) of words and allow us to label the resulting hierarchical structure
to semantics (Clarke and Tyler, 2015) and long-term conceptual generally need two different sources of input. First, we need to
storage (Ullman, 2001; Price, 2012). retrieve the word category, which determines the general selection
In general, the cross-talk between frontal and temporal cor- rules (e.g., headedness, constituent structure), and seems to be pro-
tices in syntactic and semantic processing is central in language cessed in the inferior frontal cortex (Friederici et al., 2000a). Second,
processing. On the one hand, we can observe a certain functional the brain also has to access the grammatical relations established
segregation of the aforementioned processing streams accompa- by each specific lexical item, which bring together syntactic and
nied by a division of labor between different brain areas. On the semantic information. This is, for example, the case of the argu-
other hand, the information conveyed by these two processing ment structure of a verb, where a certain hierarchical structure
streams has to be integrated, which especially takes place at level with specific constituents marked by case, such as nominative or
of the posterior temporal cortex (Friederici, 2011). The interaction accusative, correspond to the different thematic roles of the verb,
between the frontal cortex and ventral brain areas both in the lat- such as agent or patient. This information, in turn, seems to rather
eral and in the medial temporal cortex plays an important role be stored in the in the posterior temporal lobe (Bornkessel et al.,
in categorization processes. Already in primates for non-linguistic 2005). The dorsal stream, in particular the arcuate fascicle, would
218 T. Goucha et al. / Neuroscience and Biobehavioral Reviews 81 (2017) 213–224

Fig. 2. Types of operations performed in the language network. Our framework aims to disentangle different kinds of computations performed within the perisylvian
cortex and identify which circuits are responsible for their neural implementation. We also distinguish artificial grammar processing from natural language processing.
In language comprehension and production, semantic combinatorial processing mainly takes place along the ventral stream, whereas syntactic hierarchical processing is
located in the dorsal stream.

then enable the establishment of these syntactic-semantic relations claim that they should be narrowed down to a single computa-
that require the interaction of these two sources of information. tion, thus making any distinction between simple, linear structures
This way, labeling would be anatomically be enabled by the arcu- and complex, hierarchical structures rather futile. Although it is
ate fascicle, which allows the connection with the interfaces. The clear from a linguistic theoretical point of view that the complex
dorsal pathway, especially the arcuate fascicle, is hence crucial part syntactic structures result from the reapplication of the same ele-
of the rule-based system engaged in the processing of hierarchi- mentary combinatorial algorithm (cf. 2. The Merge-only hypothesis
cal structures. The uniquely semantic dimension of words, in turn, and why we need labels), we put forward that not all computations
seems to be processed along the ventral pathway, with associative performed in the language network are necessarily hierarchical
principles of word combination. We must thus distinguish between (Frank et al., 2012; Christiansen and Chater, 2015), in particular the
the rule-based mechanism of the dorsal stream connecting BA 44 combinatorial processes that happen in the semantic domain. Here,
and the posterior temporal cortex, where sequences are assigned the distinction between competence and performance (Chomsky,
a hierarchical structure, and the combinatorial mechanisms of the 1965) might prove very helpful. Although our language competence
ventral stream connecting BA 45 and the temporal cortex. The kind implies the aforementioned rule-based mechanism of processing
of concatenation that we find in the ventral stream and in the evo- in the dorsal stream, where sequences are assigned a hierarchi-
lutionarily older regions of the frontal cortex is essentially different cal structure, the performance that is implemented in the ventral
from the hierarchical labeled structures produced by the dorsal stream can be based in more simple heuristics of template match-
stream (Friederici et al., 2006; Zaccarella and Friederici, 2015b). The ing and resort to associative thinking, producing linear sequences
combinations performed along the ventral stream are mainly based instead. Such associative mechanisms do not rely on hierarchical
on associative operations, possibly resorting to template match- structures, thus not being able to fully explain syntactic phe-
ing, which can give rise to linear sequences. The ventral stream nomena. These strategies for language comprehension have lower
was shown to be fundamental in concept combination (DeWitt processing demands, since they are based on a superficial under-
and Rauschecker, 2012), in particular in the anterior temporal lobe standing that does not require the full projection of the syntactic
(Bemis and Pylkkänen, 2013a), an area that has also been associated structure. Participants have been shown to follow such strategies
with retrieval of word meaning (Mesulam et al., 2015). The com- when not specifically instructed to pay attention to the precise
binatorial processes in the ventral stream thus seem to depend on structure (Ferreira and Patson, 2007). In general, this kind of
local transition rules, as we can also see in artificial grammar for associative strategies was recently highlighted as an important
adjacent non-hierarchical dependencies studies (Friederici et al., component of complex cognition (Heyes, 2012), despite the fact
2006). Language processing along the ventral stream appears to that they are more error-prone, based on fast and frugal heuristics
be dependent on declarative memory, operating on the associative (Gigerenzer and Goldstein, 1996).
strengths between the elements of a sequence. These mechanisms Cognitive models presenting a dissociation of rule-based learn-
of concept association have been put forward as spreading activa- ing and associative learning are often criticized (Fitch, 2014) since
tion (Collins and Loftus, 1975), where representations engage other both types of learning are still based on common mechanisms, in
meanings that neighbor their conceptual network (Pulvermüller, particular through predictive coding (Clark, 2013). Human cog-
2013), with a crucial role played by the temporal lobe. The con- nition, including language, mainly depends on predictive models
ceptual networks established by association rather give rise to formed in higher cognitive areas about the incoming stimulus and
mental scenes, without any hierarchical projection of its elements. its continuous adaptation as a function prediction errors coming
In Fig. 2 we schematically summarize the possible different kinds of from primary cortices (Giraud and Poeppel, 2012). Here, we have
computation taking place in the language network and their neu- to take into account that the predictions made in online process-
ral implementation, spanning from artificial grammar learning to ing are based upon mental models. This is precisely where the
combinatorial semantics and hierarchical structure building. difference between the two kinds of processing in the dorsal and
The possibility of different combinatorial mechanisms in lan- ventral stream resides. In the dorsal stream, we find a predictive
guage was criticized by Bornkessel-Schlesewsky et al. (2015), who model based on an internal meta-representational level of gram-
T. Goucha et al. / Neuroscience and Biobehavioral Reviews 81 (2017) 213–224 219

mar, allowing for non-adjacent dependencies (Friederici, 2012). of syntax. In fact, Goucha and Friederici (2015) and Zaccarella and
Such a model generates predictions that will be radically different Friederici (2015c) show how individuals are able to successfully
from those generated by another model exclusively based on local retrieve the combinatory rules that operate on labels, even though
transition probabilities, for example depending on the frequency of the access to the actual lexical items was blocked by the use of
item co-occurrence, generally implemented in the ventral stream. pseudowords in both experiments. The two studies were further-
In fact, it was possible to see in recent fMRI studies that differ- more able to trigger the same pattern of activation seen in syntactic
ent cues used to establish predictions about the incoming input manipulations with real words.
engaged segregated processing streams. If cues about content were By distinguishing two combinatorial systems in human cogni-
used, i.e. supported by combinatorial processes based on rela- tion with two segregated implementations, we put into question
tional semantics, the areas connected by the ventral stream were the thesis that the multimodal combination of concepts allowed
engaged, that is the anterior temporal and anterior frontal cortices by Merge is the actual novelty of language (Boeckx and Benítez-
(Tuennerhoff and Noppeney, 2016; Scharinger et al., 2015) and also Burraco, 2014). If applied to the linguistic domain, there is a corpus
the hippocampus (Piai et al., 2016). In turn, if the cues used for sen- of evidence that conceptual/semantic combination takes place in
tence processing and to make online predictions were limited to the anterior temporal lobe, a central semantic hub (Bemis and
syntactic cues, i.e., cues about sentence structure, we rather see Pylkkänen, 2013b). Multimodal association of other kinds of men-
the involvement of areas that are connected by the dorsal stream tal representations, however, does not involve the same neural
(Goucha and Friederici, 2015). Furthermore, brain oscillations at mechanisms as conceptual combination in the semantic domain
different timescales were shown to reflect combinatorial process- let alone syntactic combination. This kind of combination, gener-
ing in language in phrases and sentences in a way that could not ally coordinated by the visual and auditory ventral streams in the
be totally accounted for by mere local transitional probabilities temporal lobe, is not necessarily specific of the language system
(Ding et al., 2016). In fact, even in the case of adjacent dependen- nor it is of humans. The medial temporal lobe, part of the declar-
cies, we do not find exclusively the involvement of the ventral ative system (Ullman, 2001), is highly conserved in mammals and
stream as a function of the violation of local structure building its mechanisms have similar functioning principles that in gen-
(Herrmann et al., 2009). When this violation is a product of hier- eral allow for the association and combination of mental objects
archical rules, including agreement, there is a clear involvement of in memory. The connections of the hippocampus with the cere-
the IFG (Friederici et al., 2000b), and in particular BA 44 (Herrmann bral neocortex allow for this multimodal integration of mental
et al., 2012; Hanna et al., 2014). These studies make a point for objects in the process of memory formation and can be seen even
the neural implementation of a predictive model based on a gen- at very basic domains of cognition present in other species (Suzuki,
erative structural combinatorial algorithm. From a neural point 2007; Langston and Wood, 2010). Indeed, it can account for the
of view, we should distinguish two main combinatorial systems representation of time-oriented sequences (McKenzie et al., 2014;
in language: a dorsal and a ventral one. Both systems necessarily Howard et al., 2014), even in artificial grammars, but only when
encompass bidirectional connections with top-down predictions item-based and not for rule-based learning (Opitz and Friederici,
being sent downstream and prediction errors being sent upstream 2004).3 Further, this combinatory mechanism has no constraints
in online language comprehension. The main difference neverthe- on the representations it combines, but it clearly is not enough to
less is which predictive models are used by these two processing bring language about in other species. That’s why the mechanis-
streams. Whereas the dorsal stream acts according to a hierarchical tic difference between the combinatorial faculty of language and
model based on categorical knowledge, permitting long-distance other domains proposed here should account for the emergence of
dependencies, the ventral stream is rather dependent on associa- language in humans.
tive learning based on transition probabilities in local dependencies In the investigation of the language faculty, some particu-
and usually at fixed distances, allowing for the semantic combina- lar theoretical proposals received especial attention, in particular
tion of the sequence elements. recursion (Hauser et al., 2002). The empirical investigation of
However, the engagement of BA 44, the IFG and the dorsal recursion showed that the cognitive resources used for the rep-
stream in general in hierarchical processing is still often attributed resentation of self-similarity in non-linguistic domains does not
to mere linearization processes, i.e. the IFG is rather responsible depend on language mechanisms (Martins et al., 2015), and also
to establish the serial order of the constituents (Bornkessel- possesses a very distinct neural representation in the visual domain
Schlesewsky et al., 2015). According to this view, this region should (Martins, Fischmeister et al., 2014), with no actual overlap with
be rather engaged in converting a hierarchical structure still with- linguistic recursion, as seen in Makuuchi et al. (2009). This is an
out a fixed order to a linear sequence that can be sent to the interesting observation and suggests that recursion of asymmet-
interfaces (especially sensorimotor), which is the point at which ric structures may be specific. Within language, however, minimal
labeling would intervene. However, the capacity of a brain structure phrasal structures in language (Zaccarella et al., 2015) largely share
to linearize hierarchical structures entails the capacity of build- the brain activation patterns of more complex embedded recursive
ing hierarchical structures itself. If a certain brain area hosts the structures (Friederici et al., 2009). This pinpoints how the reap-
computations that allow it to infer the labeling of a hierarchi- plication of the same computation of labeling not only accounts
cal structure in order to produce a linear sequence, this area is for minimal structures but also for complex structures, keeping a
necessarily capable of building that same hierarchical structure. similar neural representation.
Furthermore, as discussed in the previous section, the labeling of
a hierarchical structure influences its word order, but they are 3.3. Integrating the hypothesis with evolutionary evidence in
still two separate dimensions. Boeckx (2014) also voiced criticism neuroanatomy
regarding the lexicocentricity of theoretical accounts of label-
ing portrayed in recent work of Hornstein (2009) and Bickerton In order to consolidate our anatomical hypothesis regarding the
(2014a), more recently discussed by Murphy (2015a). We would arcuate fascicle, we should integrate it within the recent devel-
like to stress that labeling can still take place in the absence of con-
crete lexical items and is rather based on an abstract dimension,
which in turn allows labels to be marked by inflectional morphol- 3
Rule-based learning of an artificial grammar due to grammatical categories
ogy. These labels are rather bound to the categories that make the allowing labeling and hiearchy building showed activation in Broca’s area once the
interface between the mental lexicon and the rule-based system hierarchical rules were learned (Opitz and Friederici, 2004).
220 T. Goucha et al. / Neuroscience and Biobehavioral Reviews 81 (2017) 213–224

opments in comparative genetic studies and more specifically in cognition through the hierarchical embedding of oscillations are
the genetics of language. Geschwind and Rakic (2013) highlight transversal across animals.4
how the big innovation in the human central nervous system In summary, our hypothesis about the neural implementation
have to do with a reshaping of the cortex, especially the pre- of the faculty of language concerns the plasticity in the brain net-
frontal and parietal cortices. More specifically, most of the genes works in both developmental and evolutionary scales. In particular,
associated with vocal behavior and language play a major role we consider the possibility of wiring of brain areas with different
in neural plasticity (Szalontai and Csiszar, 2013). Several stud- contents and functioning modes, instead of assuming a closed hard-
ies show that these genes are related to microscopic connectivity wired circuit at birth. This is the kind of non-disruptive evolution
(Charrier et al., 2012; Enard, 2011; French et al., 2012; Groszer with a plausible genetic background that can still generate great
et al., 2008). Crucially, this microstructural plasticity could be novelty. This point of view allows for coexistence of continuity in
correlated with behavioral changes, leading to a selective improve- the brain mechanisms, which can be seen in brain oscillations com-
ment in rule-based procedural memory (Schreiweis et al., 2014). mon to humans and NHPs, and discontinuity in behavior brought
Overall, the recent phylogenetic changes in the network of genes about by novelty in brain structure.
related to neural plasticity should also influence in the macro-
scopic long-range wiring of brain regions (Geschwind and Rakic,
3.4. Implications for future research: Which predictions can we
2013).
make?
This novel brain rewiring in humans could be the origin of the
connectivity differences between the prefrontal cortex and the
If we intend to investigate the language faculty and its combi-
temporal cortex that we observe in the case of the arcuate fasci-
natorial algorithm isolating its core computation, the easiest way
cle (Rilling, 2014). The cortical terminations of the arcuate fascicle
to approach it without having the strong confound of lexicality is
also seem to be particularly different in comparison with NHPs.
to test it through an artificial grammar. In order to address the role
Especially, BA 44 suffered a big expansion and acquired a left-
of the IFG and the arcuate fascicle in the neural implementation
right asymmetry (Schenker et al., 2010; Spocter et al., 2010), which
of labeling and ultimately in language emerge, it is important to
is accompanied by the strong asymmetry of the arcuate fascicle
design experimental paradigms that directly compare the acquisi-
itself (Nucifora et al., 2005; Vernooij et al., 2007; Takaya et al.,
tion of grammars based on labeled structures with other possible
2015). Rilling et al. (2008, 2011) show a striking difference between
strategies. We should therefore compare rule-based learning, to
the ventral and dorsal anatomical connectivity between NHPs and
grammars that are exclusively based on item similarity, controlling
humans. Even in our closest relatives, the chimpanzees, there is
for other variables like dependency length, that we do not think to
only a faint connection of Broca’s homologues to the posterior
be central to the computational aspect.
planum temporale, whereas humans display the robust termina-
Although most studies in AGL did not include labeling rules in
tions of the arcuate fascicle in the posterior superior and middle
their paradigms, some studies did implement the necessarily labels
temporal gyri (Fig. 3, upper panel). The ventral stream, however, is
for language-like phrase structure rules, in an attempt to attain an
already well developed in NHPs, as shown anatomically with diffu-
abstract structure that would be more comparable to language than
sion tractography and autoradiography, especially from the ventral
other artificial grammars in use. Rohrmeier et al. (2012) showed
PFC to the anterior and middle temporal gyrus (Schmahmann et al.,
behaviorally that participants acquired knowledge of the grammat-
2007), also corroborated by functional connectivity (Neubert et al.,
ical classes and their relations, i.e. the selection rules in use, which
2014; Petkov et al., 2015). In general, these crucial white matter
also applied for long distance dependencies, and were not based on
differences we have been discussing might be related to more gen-
n-gram knowledge. Studies using BROCANTO, an artificial language
eral principles of brain organization distinguishing humans from
also including language-like labels without content, could provide
NHPs (Buckner and Krienen, 2013). Already prenatally, there is
interesting insights in the processing of such structures, more akin
an accentuated slowing down in brain volume growth present in
to language hierarchies. In particular, these studies could show
chimpanzees not observed in humans (Sakai et al., 2012), to be
dissociations between more rule-based learning, based on the arti-
followed by a faster post-birth prefrontal growth in humans in com-
ficial hierarchical structure, and more associative learning, based
parison to both chimpanzees and Rhesus macaques (Sakai et al.,
on sequence similarity and local transitions (Opitz and Friederici,
2011), especially in white matter compared to gray matter (Sakai
2004). The hippocampus was initially more engaged in the acqui-
et al., 2013). Interestingly, these developmental differences across
sition of the grammar, being rather responsive to similarity-based
species find a correspondent in the particular ontogeny of the dor-
learning, but its activation dropped whereas the posterior IFG was
sal pathway in human infants (Dubois et al., 2015). In particular,
progressively more engaged (Opitz and Friederici, 2003). Crucially,
whereas the part of the dorsal stream responsible for articulation
the posterior IFG, especially BA 44, was primarily engaged in rule-
is already fully developed in newborns, the arcuate fascicle has yet
learning, in comparison with the ventral premotor cortex or the
to undergo myelination (Perani et al., 2011; Fig. 3, lower panel).
hippocampus, rather related to the local character of rule change
These insights from developmental and comparative studies sup-
(Opitz and Friederici, 2007). Furthermore, the use of analogous
port our hypothesis regarding the arcuate fascicle and its crucial
rules in cotton-top tamarins and infants showed again a differ-
role in the emergence of language.
ential between humans and NHPs (Saffran et al., 2008). Whereas
Alternative mechanisms based on brain oscillations have been
both infants and cotton-top tamarins were able to detect regular-
proposed as a crucial element for the emergence of language
ities in simple combinatorial patterns, only infants were able to
(Murphy, 2015b). The oscillatory processing of speech and the
acquire complex grammatical structures with a phrase structure
general issuing computational mechanisms are in fact crucial to
by the predictive patterns established by a grammar with selection
understand how the brain encodes different linguistic aspects
rules.
(Giraud and Poeppel, 2012). However, those mechanisms seem to
already be in place in other species. For example, despite the crucial
brain expansion that took place in primates and especially humans
4
compared to other mammals, the rhythmical hierarchy of oscilla- However, there is first evidence that brain oscillations in the theta band related
to verbal working memory can be segregated both functional and topographi-
tions is mainly kept unchanged (Buzsáki et al., 2013). As Friederici cally into a domain-general and a syntax-specific component, with the latter being
and Singer (2015) pinpoint, the basic neural mechanisms behind located in Broca’s area (Beese, Meyer, Vassileiou & Friederici, in preparation). This
once again speaks rather for an anatomical distinction.
T. Goucha et al. / Neuroscience and Biobehavioral Reviews 81 (2017) 213–224 221

Fig. 3. Phylogeny and ontogeny of the dorsal stream. The arcuate fascicle has a particular profile in both comparative and developmental research. The upper panel shows
how the dorsal connections from the IFG to the posterior temporal cortex are highly strengthened in humans in comparison with chimpanzees. The lower panel displays the
fronto-temporal structural connectivity of newborns and adults, where we can observe that the arcuate fascicle is not yet fully myelinated at birth, in contrast to the other
fiber tracts involved in language processing, which are trackable from birth.

Fig. 4. Recent developments in the neural implementation of syntactic processing. A. In the absence of all semantic cues, including derivational morphology, only the
posterior part of the IFG, BA 44, is activated. These data indicate that BA 44 is the core area for abstract syntax processing. B. Even in the processing of local structures (3-word
phrases or sentences) involving lexical items, there is a co-activation of the IFG and the posterior temporal cortex. C. The current framework focuses on the arcuate fascicle
as the structure that enables the cross-talk between abstract structural building in the IFG and the lexically-based argument structure in the posterior temporal cortex.

One important question that remains to be addressed empiri- studies that isolated the abstract language structure regardless of
cally is the precise function of the arcuate fascicle in hierarchical meaning (Goucha and Friederici, 2015; Fig. 4a) or in the local syn-
structure building. The literature points to its crucial role in linguis- tactic combination of a determiner and a pseudoword resulting
tic syntax because its degree of myelination determines syntactic from a single Merge (Zaccarella and Friederici, 2015c), discussed in
performance in processing syntactically complex sentences in another paper in this issue (Zaccarella and Friederici, 2016). Note
patients (Wilson et al., 2011) and in children (Skeide et al., 2015). that the combination of two words in language can be seen as a
The arcuate fascicle connects the IFG and the posterior temporal hierarchical structure since this paradigm is productive, as dis-
cortex, which in adults are often simultaneously involved in the cussed by Yang (2013), that is, a determiner phrase (DP) allows
processing of hierarchically complex sentences (Friederici et al., for all combinations of determiners and nouns. Do to the genera-
2009; Meyer et al., 2012; Pattamadilok et al., 2015; Skeide et al., tive properties of this syntactic element, we can already talk about
2015), and also in structure building of phrases (Zaccarella et al., hierarchy—a structure that is impossible in adjacent relationships
2015; Fig. 4b). However, the posterior temporal cortex is not in an artificial grammar. Therefore, according to our interpreta-
involved in most artificial grammar studies and also in language tion, this single combination of two words already encompasses
222 T. Goucha et al. / Neuroscience and Biobehavioral Reviews 81 (2017) 213–224

the labeling of two linguistic elements that have previously been References
concatenated. It appears that in order to recruit the temporal cor-
tex the second element needs to be a real noun—a lexical element Amunts, K., Schleicher, A., Bürgel, U., Mohlberg, H., Uylings, H., Zilles, K., 1999.
Broca’s region revisited: cytoarchitecture and intersubject variability. J. Comp.
with a clear label triggering lexical access. The word categories that Neurol. 412 (2), 319–341.
constitute the labels are delivered by the lexicon, which is located Amunts, K., Lenzen, M., Friederici, A.D., Schleicher, A., Morosan, P.,
in the middle temporal cortex (Lau et al., 2008; Price, 2012) in the Palomero-Gallagher, N., Zilles, K., 2010. Broca’s region: novel organizational
principles and multiple receptor mapping. Plos Biol. 8 (9), e1000489.
ventral stream. This raises the question at which step the arcuate Badre, D., D’Esposito, M., 2009. Is the rostro-caudal axis of the frontal lobe
fascicle comes into play. The framework put forward in this paper hierarchical? Nat. Rev. Neurosci. 10 (9), 659–669.
highlights the role of white matter pathways in the establishment Badre, D., 2008. Cognitive control, hierarchy, and the rostro-caudal organization of
the frontal lobes. Trends Cogn. Sci. 12 (5), 193–200.
of brain computations across distant brain regions. One hypothe- Bahlmann, J., Schubotz, R.I., Friederici, A.D., 2008. Hierarchical artificial grammar
sis for the function of the arcuate fascicle would be to allow that processing engages Broca’s area. Neuroimage 42 (2), 525–534.
the IFG operates hierarchically on the labels and matches them to Bahlmann, J., Schubotz, R.I., Mueller, J.L., Koester, D., Friederici, A.D., 2009. Neural
circuits of hierarchical visuo-spatial sequence processing. Brain Res. 1298,
lexically-based argument structure frames stored in the posterior
161–170.
temporal cortex. The integration of these information types is made Beese, C., Meyer, L., Vassileiou, B., Friederici, A.D., (in preparation). Resting state
possible through the cross-talk established via the arcuate fascicle theta power as lifespan marker of sentence comprehension.
(as portrayed in Fig. 4c). Bemis, D.K., Pylkkänen, L., 2013a. Basic linguistic composition recruits the left
anterior temporal lobe and left angular gyrus during both listening and
reading. Cereb. Cortex, bhs170.
Bemis, D.K., Pylkkänen, L., 2013b. Combination across domains: an MEG
investigation into the relationship between mathematical, pictorial, and
linguistic processing. Front. Psychol. 3, 583.
4. Conclusions: Is there really something special about
Berwick, R., Chomsky, N., 2011. The biolinguistic program: the current stage of its
language? development. In: di Sciullo, Anna Maria, Boeckx, Cedric (Eds.), The Biolinguistic
Enterprise. Oxford University Press, Oxford, pp. 19–41.
It is undeniable that language is a uniquely human property, Berwick, R.C., Chomsky, N., 2016. Why Only Us: Language and Evolution. MIT
Press, Cambridge, MA.
but it is less clear whether there is a language-specific faculty Berwick, R.C., Friederici, A.D., Chomsky, N., Bolhuis, J.J., 2013. Evolution, brain, and
behind or if language is simply enabled by the general cognitive the nature of language. Trends Cognit. Sci. 17 (2), 89–98.
capacity of humans, including enhanced memory and more sophis- Bickerton, D., 2014a. More than nature needs. In: Language, Mind, and Evolution.
Harvard University Press, Cambridge, MA.
ticated sociality. Despite claims about the linearity of language Bickerton, D., 2014b. Some problems for biolinguistics. Biolinguistics 8, 73–96.
and the possibility of processing linguistic strings in a parallel Boeckx, C., Benítez-Burraco, A., 2014. The shape of the human language-ready
non-hierarchical fashion (Frank et al., 2012), there are a num- brain. Front. Psychol. 5.
Boeckx, C., 2013. Biolinguistics: forays into human cognitive biology. J. Anthropol.
ber of linguistic phenomena that can only fully be accounted for Sci. 91, 1–28.
by representing utterances as structures with relations of domi- Boeckx, C., 2014. Elementary Syntactic Structures: Prospects of a Feature-Free
nance between their elements (Everaert et al., 2015). Following the Syntax. Cambridge University Press, Cambridge, http://dx.doi.org/10.1017/
CBO9781139524391.
rationale of the Minimalist Program, we aimed to identify the fun-
Bolhuis, J.J., Tattersall, I., Chomsky, N., Berwick, R.C., 2014. How could language
damental computational basis of hierarchical structure building in have evolved? Plos Biol. 12 (8), e1001934.
language. By integrating linguistic theory in an evolutionary per- Bornkessel, I., Zysset, S., Friederici, A.D., von Cramon, D.Y., Schlesewsky, M., 2005.
Who did what to whom? The neural basis of argument hierarchies during
spective, we isolated the need for a labeling algorithm operating on
language comprehension. Neuroimage 26 (1), 221–233.
the output of the elementary combinatorial computation. While Bornkessel-Schlesewsky, I., Schlesewsky, M., Small, S.L., Rauschecker, J.P., 2015.
grounding this labeling algorithm and linguistic structure build- Neurobiological roots of language in primate audition: common computational
ing in neural circuits, we reinstated a functional divide between properties. Trends Cogn. Sci. 19 (3), 142–150.
Brauer, J., Anwander, A., Friederici, A.D., 2011. Neuroanatomical prerequisites for
dorsal and ventral white matter pathways in language processing. language functions in the maturing brain. Cereb. Cortex 21 (2), 459–466.
The ventral pathway, already developed in NHPs, allows for serial Brodmann, K., 1909. Vergleichende Lokalisationslehre der Grosshirnrinde in ihren
concatenation and is able to combine meanings based on their asso- Prinzipien dargestellt auf Grund des Zellenbaues. Barth.
Buckner, R.L., Krienen, F.M., 2013. The evolution of distributed association
ciative strength, accounting for a non-hierarchical dimension of networks in the human brain. Trends Cogn. Sci. 17 (12), 648–665.
language (Jackendoff and Wittenberg, 2014) and non-human pri- Buzsáki, G., Logothetis, N., Singer, W., 2013. Scaling brain size: keeping timing:
mate communication. In turn, the dorsal pathway is responsible evolutionary preservation of brain rhythms. Neuron 80 (3), 751–764.
Catani, M., Jones, D.K., 2005. Perisylvian language networks of the human brain.
for the full-fledged hierarchical structures that characterize lan- Ann. Neurol. 57 (1), 8–16.
guage, allowing for local combinations of words with a labeling Charrier, C., Joshi, K., Coutinho-Budd, J., Kim, J.E., Lambert, N., De Marchena, J.,
algorithm that can be recursively applied and generate complex et al., 2012. Inhibition of SRGAP2 function by its human-specific paralogs
induces neoteny during spine maturation. Cell 149 (4), 923–935.
hierarchical structures. Anatomically, we highlight the crucial role
Chesi, C., Moro, A., 2015. The subtle dependency between competence and
of the arcuate fascicle for this rule-based processing mechanism, performance. In: 50 Years Later: Reflections on Chomsky’s Aspects. MIT
with its important evolutionary path (Rilling, 2014), integrated in Working Papers in Linguistics, Cambridge, M.
Chomsky, N., 1965. Aspects of the Theory of Syntax. MIT Press, Cambridge, MA.
general principles of brain evolution (Geschwind and Rakic, 2013).
Chomsky, N., 1995. The Minimalist Program, Vol. 1765. MIT Press, Cambridge, MA.
This evolution from transitional probabilities to the rule-based con- Chomsky, N., 2007. Approaching UG from Below. Interfaces+
struction of linguistic trees (Fitch, 2014; Dehaene et al., 2015) Recursion = Language?, Vol. 89. Mouton de Gruyter, Berlin, pp. 1–30.
should constitute the focus of further research. At the interface Chomsky, N., 2013. Problems of projection. Lingua 130, 33–49.
Christiansen, M.H., Chater, N., 2015. The language faculty that wasn’t: a
between linguistic theory and neurolinguistics, we finally believe usage-based account of natural language recursion. Front. Psychol. 6.
that future neurobiological investigations aiming at describing the Clark, A., 2013. Whatever next? Predictive brains, situated agents, and the future of
specific nature of linguistic computations in the human brain will cognitive science. Behav. Brain Sci. 36 (03), 181–204.
Clarke, A., Tyler, L.K., 2015. Understanding what we see: how we derive meaning
have to consider the fundamental dichotomy between competence from vision. Trends Cogn. Sci. 19 (11), 677–687.
and performance (Chomsky, 1965; Chesi and Moro 2015). Any true Clerget, E., Winderickx, A., Fadiga, L., Olivier, E., 2009. Role of Broca’s area in
endeavor in biolinguistics must be able to reconcile empirical neu- encoding sequential human actions: a virtual lesion study. Neuroreport 20
(16), 1496–1499.
robiological research on the basis of performative evidence with the Collins, A.M., Loftus, E.F., 1975. A spreading-activation theory of semantic
formal constraints imposed by the internal linguistic competence. processing. Psychol. Rev. 82 (6), 407.
We therefore hope to have provided a constructive hypothesis Comrie, B., 1989. Language Universals and Linguistic Typology: Syntax and
Morphology. University of Chicago Press, Chicago, IL.
concerning a human specific language faculty that is both com-
Dapretto, M., Bookheimer, S.Y., 1999. Form and content: dissociating syntax and
putationally sound and anatomically grounded. semantics in sentence comprehension. Neuron 24 (2), 427–432.
T. Goucha et al. / Neuroscience and Biobehavioral Reviews 81 (2017) 213–224 223

den Ouden, D.B., Saur, D., Mader, W., Schelter, B., Lukic, S., Wali, E., Thompson, C.K., familiarity and length in overt picture naming. J. Cogn. Neurosci. 19 (4),
2012. Network modulation during complex syntactic processing. Neuroimage 617–631.
59 (1), 815–823. Griffiths, J.D., Marslen-Wilson, W.D., Stamatakis, E.A., Tyler, L.K., 2013. Functional
De Vries, M.H., Barth, A.C., Maiworm, S., Knecht, S., Zwitserlood, P., Flöel, A., 2010. organization of the neural language system: dorsal and ventral pathways are
Electrical stimulation of Broca’s area enhances implicit learning of an artificial critical for syntax. Cereb. Cortex, bhr386.
grammar. J. Cogn. Neurosci. 22 (11), 2427–2436. Groszer, M., Keays, D.A., Deacon, R.M., De Bono, J.P., Prasad-Mulcare, S., Gaub, S.,
DeWitt, I., Rauschecker, J.P., 2012. Phoneme and word recognition in the auditory et al., 2008. Impaired synaptic plasticity and motor learning in mice with a
ventral stream. Proc. Natl. Acad. Sci. U. S. A. 109 (8), E505–E514. point mutation implicated in human speech deficits. Curr. Biol. 18 (5),
Dehaene, S., Meyniel, F., Wacongne, C., Wang, L., Pallier, C., 2015. The neural 354–362.
representation of sequences: from transition probabilities to algebraic patterns Hagoort, P., Indefrey, P., 2014. The neurobiology of language beyond single words.
and linguistic trees. Neuron 88 (1), 2–19. Ann. Rev. Neurosci. 37 (1), 347–362.
Ding, N., Melloni, L., Zhang, H., Tian, X., Poeppel, D., 2016. Cortical tracking of Hagoort, P., 2005. On Broca, brain, and binding: a new framework. Trends Cogn.
hierarchical linguistic structures in connected speech. Nat. Neurosci. 19 (1), Sci. 9 (9), 416–423.
158–164. Hanna, J., Mejias, S., Schelstraete, M.A., Pulvermüller, F., Shtyrov, Y., Van der Lely,
Dubois, J., Poupon, C., Thirion, B., Simonnet, H., Kulikova, S., Leroy, F., et al., 2015. H.K., 2014. Early activation of Broca’s area in grammar processing as revealed
Exploring the early organization and maturation of linguistic pathways in the by the syntactic mismatch negativity and distributed source analysis. J. Cogn.
human infant brain. Cereb. Cortex, bhv082. Neurosci. 5 (2), 66–76.
Enard, W., 2011. FOXP2 and the role of cortico-basal ganglia circuits in speech and Hauser, M.D., Chomsky, N., Fitch, W.T., 2002. The faculty of language: what is it,
language evolution. Curr. Opin. Neurobiol. 21 (3), 415–424. who has it, and how did it evolve? Science 298 (5598), 1569–1579.
Everaert, M.B., Huybregts, M.A., Chomsky, N., Berwick, R.C., Bolhuis, J.J., 2015. Herrmann, B., Maess, B., Hasting, A.S., Friederici, A.D., 2009. Localization of the
Structures, not strings: linguistics as part of the cognitive sciences. Trends syntactic mismatch negativity in the temporal cortex: an MEG study.
Cogn. Sci. 19 (12), 729–743. Neuroimage 48 (3), 590–600.
Ferreira, F., Patson, N.D., 2007. The ‘good enough’approach to language Herrmann, B., Obleser, J., Kalberlah, C., Haynes, J.D., Friederici, A.D., 2012.
comprehension. Lang. Linguist. Compass. 1 (1–2), 71–83. Dissociable neural imprints of perception and grammar in auditory functional
Fitch, W.T., Hauser, M.D., 2004. Computational constraints on syntactic processing imaging. Hum. Brain Mapp. 33 (3), 584–595.
in a nonhuman primate. Science 303 (5656), 377–380. Heyes, C., 2012. Simple minds: a qualified defence of associative learning. Philos.
Fitch, W., Martins, M.D., 2014. Hierarchical processing in music, language, and Trans. R. Soc. B 367 (1603), 2695–2703.
action: lashley revisited. Ann. N.Y. Acad. Sci. 1316 (1), 87–104. Hornstein, N., 2009. A Theory of Syntax: Minimal Operations and Universal
Fitch, W.T., 2014. Toward a computational framework for cognitive biology: Grammar. University Press Cambridge, Cambridge.
unifying approaches from cognitive neuroscience and comparative cognition. Howard, M.W., MacDonald, C.J., Tiganj, Z., Shankar, K.H., Du, Q., Hasselmo, M.E.,
Phys. Life Rev. 11 (3), 329–364. Eichenbaum, H., 2014. A unified mathematical framework for coding time,
Flöel, A., de Vries, M.H., Scholz, J., Breitenstein, C., Johansen-Berg, H., 2009. White space, and sequences in the hippocampal region. J. Neurosci. 34 (13),
matter integrity in the vicinity of Broca’s area predicts grammar learning 4692–4707.
success. Neuroimage 47 (4), 1974–1981. Jackendoff, R., Wittenberg, E., 2014. In: Newmeyer, Fritz, Preston, Lauren (Eds.),
Forkstam, C., Hagoort, P., Fernández, G., Ingvar, M., Petersson, K.M., 2006. Neural What You can Say Without Syntax: a Hierarchy of Grammatical Complexity.
correlates of artificial syntactic structure classification. Neuroimage 32 (2), Measuring Grammatical Complexity. Oxford University Press, Oxford.
956–967. Jeon, H.-A., Friederici, A.D., 2013. Two principles of organization in the prefrontal
Frank, S.L., Bod, R., Christiansen, M.H., 2012. How hierarchical is language use? cortex are cognitive hierarchy and degree of automaticity. Nat. Commun. 4,
Proc. R. Soc. Lond. B Biol., rspb20121741. 2041.
Freedman, D.J., Riesenhuber, M., Poggio, T., Miller, E.K., 2003. A comparison of Jeon, H.A., Friederici, A.D., 2015. Degree of automaticity and the prefrontal cortex.
primate prefrontal and inferior temporal cortices during visual categorization. Trends Cogn. Sci. 19 (5), 244–250.
J. Neurosci. 23 (12), 5235–5246. Kayne, Richard S., 1994. The Antisymmetry of Syntax. Linguistic Inquiry
French, C.A., Jin, X., Campbell, T.G., Gerfen, E., Groszer, M., Fisher, S.E., Costa, R.M., Monograph Twenty-Five. MIT Press, Cambridge, MA.
2012. An aetiological Foxp2 mutation causes aberrant striatal activity and Koechlin, E., Basso, G., Pietrini, P., Panzer, S., Grafman, J., 1999. The role of the
alters plasticity during skill learning. Mol. Psychiatry 17 (11), 1077–1085. anterior prefrontal cortex in human cognition. Nature 399 (6732), 148–151.
Frey, S., Campbell, J.S., Pike, G.B., Petrides, M., 2008. Dissociating the human Langston, R.F., Wood, E.R., 2010. Associative recognition and the hippocampus:
language pathways with high angular resolution diffusion fiber tractography. J. differential effects of hippocampal lesions on object-place, object-context and
Neurosci. 28 (45), 11435–11444. object-place-context memory. Hippocampus 20 (10), 1139–1153.
Friederici, A.D., Gierhan, S.M.E., 2013. The language network. Curr. Opin. Neurobiol. Lau, E.F., Phillips, C., Poeppel, D., 2008. A cortical network for semantics:
23 (2), 250–254. (de)constructing the N400. Nat. Rev. Neurosci. 9, 920–933.
Friederici, A.D., Singer, W., 2015. Grounding language processing on basic Makris, N., Kennedy, D.N., McInerney, S., Sorensen, A.G., Wang, R., Caviness, V.S.,
neurophysiological principles. Trends Cogn. Sci. 19 (6), 329–338. Pandya, D.N., 2005. Segmentation of subcomponents within the superior
Friederici, A.D., Opitz, B., von Cramon, D.Y., 2000a. Segregating semantic and longitudinal fascicle in humans: a quantitative, in vivo, DT-MRI study. Cereb.
syntactic aspects of processing in the human brain: an fMRI investigation of Cortex 15 (6), 854–869.
different word types. Cereb. Cortex 10 (7), 698–705. Makuuchi, M., Bahlmann, J., Anwander, A., Friederici, A.D., 2009. Segregating the
Friederici, A.D., Wang, Y., Herrmann, C.S., Maess, B., Oertel, U., 2000b. Localization core computational faculty of human language from working memory. Proc.
of early syntactic processes in frontal and temporal cortical areas: a Natl. Acad. Sci. U. S. A. 106 (20), 8362–8367.
magnetoencephalographic study. Hum. Brain Mapp. 11 (1), 1–11. Martins, M.J., Fischmeister, F.P., Puig-Waldmueller, E., Oh, J., Geißler, A., Robinson,
Friederici, A.D., Rüschemeyer, S.A., Hahne, A., Fiebach, C.J., 2003. The role of left S., et al., 2014. Fractal image perception provides novel insights into
inferior frontal and superior temporal cortex in sentence comprehension: hierarchical cognition. Neuroimage 96, 300–308.
localizing syntactic and semantic processes. Cereb. Cortex 13 (2), 170–177. Martins, M.D.J.D., Muršič, Z., Oh, J., Fitch, W.T., 2015. Representing visual recursion
Friederici, A.D., Bahlmann, J., Heim, S., Schubotz, R.I., Anwander, A., 2006. The brain does not require verbal or motor resources. Cogn. Psychol. 77, 20–41.
differentiates human and non-human grammars: functional localization and McKenzie, S., Frank, A.J., Kinsky, N.R., Porter, B., Rivière, P.D., Eichenbaum, H., 2014.
structural connectivity. Proc. Natl. Acad. Sci. U. S. A. 103 (7), 2458–2463. Hippocampal representation of related and opposing memories develop within
Friederici, A.D., Makuuchi, M., Bahlmann, J., 2009. The role of the posterior superior distinct, hierarchically organized neural schemas. Neuron 83 (1), 202–215.
temporal cortex in sentence comprehension. Neuroreport 20 (6), 563–568. Mesulam, M.M., Thompson, C.K., Weintraub, S., Rogalski, E.J., 2015. The Wernicke
Friederici, A.D., 2009. Allocating functions to fiber tracts: facing its indirectness. conundrum and the anatomy of language comprehension in primary
Trends Cogn. Sci. 13 (9), 370–371. progressive aphasia. Brain 138 (8), 2423–2437.
Friederici, A.D., 2011. The brain basis of language processing: from structure to Meyer, L., Obleser, J., Anwander, A., Friederici, A.D., 2012. Linking ordering in
function. Physiol. Rev. 91 (4), 1357–1392. Broca’s area to storage in left temporo-parietal regions: the case of sentence
Friederici, A.D., 2012. The cortical language circuit: from auditory perception to processing. NeuroImage 62 (3), 1987–1998.
sentence comprehension. Trends Cogn. Sci. 16 (5), 262–268. Moro, A., 2000. Dynamic antisymmetry (No. 38). MIT press.
Fujita, K., 2014. Recursive merge and human language evolution. In: Roeper, T., Moro, A., 2014. On the similarity between syntax and actions. Trends Cogn. Sci. 18,
Speas, M. (Eds.), Recursion: Complexity in Cognition. Springer International 109–110.
Publishing, Switzerland, pp. 243–264. Murphy, E., 2015a. Labels, cognomes and cyclic computation: an ethological
Geschwind, D.H., Rakic, P., 2013. Cortical evolution: judge the brain by its cover. perspective. Front. Psychol. 6, 715.
Neuron 80 (3), 633–647. Murphy, E., 2015b. The brain dynamics of linguistic computation. Front. Psychol. 6.
Gigerenzer, G., Goldstein, D.G., 1996. Reasoning the fast and frugal way: models of Narita, H., 2014. Endocentric Structuring of Projection-free Syntax, Vol. 218. John
bounded rationality. Psychol. Rev. 103 (4), 650. Benjamins Publishing Company.
Giraud, A.L., Poeppel, D., 2012. Cortical oscillations and speech processing: Neubert, F.X., Mars, R.B., Thomas, A.G., Sallet, J., Rushworth, M.F., 2014. Comparison
emerging computational principles and operations. Nat. Neurosci. 15 (4), of human ventral frontal cortex areas for cognitive control and language with
511–517. areas in monkey frontal cortex. Neuron 81 (3), 700–713.
Goucha, T.B., Friederici, A.D., 2015. The language skeleton after dissecting meaning: Nucifora, P.G., Verma, R., Melhem, E.R., Gur, R.E., Gur, R.C., 2005. Leftward
a functional segregation within Broca’s area. Neuroimage 114 (6), 294–302. asymmetry in relative fiber density of the arcuate fasciculus. Neuroreport 16
Graves, W.W., Grabowski, T.J., Mehta, S., Gordon, J.K., 2007. A neural signature of (8), 791–794.
phonological access: distinguishing the effects of word frequency from
224 T. Goucha et al. / Neuroscience and Biobehavioral Reviews 81 (2017) 213–224

Opitz, B., Friederici, A.D., 2003. Interactions of the hippocampal system and the Saur, D., Kreher, B.W., Schnell, S., Kümmerer, D., Kellmeyer, P., Vry, M.S., et al.,
prefrontal cortex in learning language-like rules? Neuroimage 19 (4), 2008. Ventral and dorsal pathways for language. Proc. Natl. Acad. Sci. U. S. A.
1730–1737. 105 (46), 18035–18040.
Opitz, B., Friederici, A.D., 2004. Brain correlates of language learning: the neuronal Scharinger, M., Bendixen, A., Herrmann, B., Henry, M.J., Mildner, T., Obleser, J.,
dissociation of rule-based versus similarity-based learning. J. Neurosci. 24 (39), 2015. Predictions interact with missing sensory evidence in semantic
8436–8440. processing areas. Hum. Brain Mapp. 37 (2), 704–716.
Opitz, B., Friederici, A.D., 2007. Neural basis of processing sequential and Schenker, N.M., Hopkins, W.D., Spocter, M.A., Garrison, A.R., Stimpson, C.D., Erwin,
hierarchical syntactic structures. Hum. Brain Mapp. 28 (7), 585–592. J.M., et al., 2010. Broca’s area homologue in chimpanzees (Pan troglodytes):
Overath, T., McDermott, J.H., Zarate, J.M., Poeppel, D., 2015. The cortical analysis of probabilistic mapping, asymmetry, and comparison to humans. Cereb. Cortex
speech-specific temporal structure revealed by responses to sound quilts. Nat. 20 (3), 730–742.
Neurosci. 18 (6), 903–911. Schmahmann, J.D., Pandya, D.N., Wang, R., Dai, G., D’Arceuil, H.E., de Crespigny, A.J.,
Pattamadilok, C., Dehaene, S., Pallier, C., 2015. A role for left inferior frontal and Wedeen, V.J., 2007. Association fibre pathways of the brain: parallel
posterior superior temporal cortex in extracting a syntactic tree from a observations from diffusion spectrum imaging and autoradiography. Brain 130
sentence. Cortex. (3), 630–653.
Perani, D., Saccuman, M.C., Scifo, P., Anwander, A., Spada, D., Baldoli, C., Friederici, Schreiweis, C., Bornschein, U., Burguière, E., Kerimoglu, C., Schreiter, S.,
A.D., 2011. Neural language networks at birth. Proc. Natl. Acad. Sci. U. S. A. 108 Dannemann, M., et al., 2014. Humanized Foxp2 accelerates learning by
(38), 16056–16061. enhancing transitions from declarative to procedural performance. Proc. Natl.
Petersson, K.M., Hagoort, P., 2012. The neurobiology of syntax: beyond string sets. Acad. Sci. U. S. A. 111 (39), 14253–14258.
Philos. Trans. R. Soc B 367 (1598), 1971–1983. Sciullo, A.M.D., Jenkins, L., 2016. Biolinguistics and the human language faculty.
Petersson, K.M., Folia, V., Hagoort, P., 2012. What artificial grammar learning Language 92 (3), e205–e236.
reveals about the neurobiology of syntax. Brain Lang. 120 (2), 83–95. Skeide, M.A., Brauer, J., Friederici, A.D., 2015. Brain functional and structural
Petkov, C.I., Kikuchi, Y., Milne, A.E., Mishkin, M., Rauschecker, J.P., Logothetis, N.K., predictors of language performance. Cereb. Cortex, bhv042.
2015. Different forms of effective connectivity in primate frontotemporal Spocter, M.A., Hopkins, W.D., Garrison, A.R., Bauernfeind, A.L., Stimpson, C.D., Hof,
pathways. Nat. Commun. 6. P.R., Sherwood, C.C., 2010. Wernicke’s area homologue in chimpanzees (Pan
Piai, V., Anderson, K.L., Lin, J.J., Dewar, C., Parvizi, J., Dronkers, N.F., Knight, R.T., troglodytes) and its relation to the appearance of modern human language.
2016. Direct brain recordings reveal hippocampal rhythm underpinnings of Proc. R. Soc. Lond. B Biol. Sci. 277.
language processing. Proc. Natl. Acad. Sci. 201603312. Suzuki, W.A., 2007. Making new memories. Ann. N.Y. Acad. Sci. 1097 (1), 1–11.
Price, C.J., 2012. A review and synthesis of the first 20 years of PET and fMRI studies Szalontai, A., Csiszar, K., 2013. Genetic insights into the functional elements of
of heard speech, spoken language and reading. Neuroimage 62 (2), 816–847. language. Hum. Genet. 132 (9), 959–986.
Pulvermüller, F., 2013. How neurons make meaning: brain mechanisms for Szczepanski, S.M., Knight, R.T., 2014. Insights into human behavior from lesions to
embodied and abstract-symbolic semantics. Trends Cogn. Sci. 17 (9), 458–470. the prefrontal cortex. Neuron 83 (5), 1002–1018.
Röder, B., Stock, O., Neville, H., Bien, S., Rösler, F., 2002. Brain activation modulated Takaya, S., Kuperberg, G.R., Liu, H., Greve, D.N., Makris, N., Stufflebeam, S.M., 2015.
by the comprehension of normal and pseudo-word sentences of different Asymmetric projections of the arcuate fasciculus to the temporal cortex
processing demands: a functional magnetic resonance imaging study. underlie lateralized language function in the human brain. Front. Neuroanat. 9.
Neuroimage 15 (4), 1003–1014. Tuennerhoff, J., Noppeney, U., 2016. When sentences live up to your expectations.
Reber, A.S., 1967. Implicit learning of artificial grammars. J. Verb. Learn. Verb. Neuroimage 124, 641–653.
Behav. 6, 855–863. Uddén, J., Folia, V., Forkstam, C., Ingvar, M., Fernández, G., Overeem, S., Petersson,
Reinhart, T.M., 1976. The Syntactic Domain of Anaphora (Doctoral Dissertation). K.M., 2008. The inferior frontal cortex in artificial syntax processing: an rTMS
Massachusetts Institute of Technology. study. Brain Res. 1224, 69–78.
Rilling, J.K., Glasser, M.F., Preuss, T.M., Ma, X., Zhao, T., Hu, X., Behrens, T.E., 2008. Uddén, J., Bahlmann, J., 2012. A rostro-caudal gradient of structured sequence
The evolution of the arcuate fasciculus revealed with comparative DTI. Nat. processing in the left inferior frontal gyrus. Philos. Trans. R. Soc. Lond. B Biol.
Neurosci. 11 (4), 426–428. Sci. 367 (1598), 2023–2032.
Rilling, J.K., Glasser, M.F., Jbabdi, S., Andersson, J., Preuss, T.M., 2011. Continuity, Ullman, M.T., 2001. A neurocognitive perspective on language: the
divergence, and the evolution of brain language pathways. Front. Evol. declarative/procedural model. Nat. Rev. Neurosci. 2 (10), 717–726.
Neurosci. 3. Vernooij, M.W., Smits, M., Wielopolski, P.A., Houston, G.C., Krestin, G.P., van der
Rilling, J.K., 2014. Comparative primate neuroimaging: insights into human brain Lugt, A., 2007. Fiber density asymmetry of the arcuate fasciculus in relation to
evolution. Trends Cogn. Sci. 18 (1), 46–55. functional hemispheric language lateralization in both right-and left-handed
Rodd, J.M., Vitello, S., Woollams, A.M., Adank, P., 2015. Localising semantic and healthy subjects: a combined fMRI and DTI study. Neuroimage 35 (3),
syntactic processing in spoken and written language comprehension: an 1064–1076.
activation likelihood estimation meta-analysis. Brain Lang. 141, 89–102. Wang, L., Uhrig, L., Jarraya, B., Dehaene, S., 2015. Representation of numerical and
Rohrmeier, M., Fu, Q., Dienes, Z., 2012. Implicit learning of recursive context-free sequential patterns in macaque and human brains. Curr. Biol. 25 (15),
grammars. PloS One 7 (10), e45885. 1966–1974.
Rolheiser, T., Stamatakis, E.A., Tyler, L.K.1, 2011. Dynamic processing in the human Wilson, S.M., Dronkers, N.F., Ogar, J.M., Jang, J., Growdon, M.E., Agosta, F.,
language system: synergy between the arcuate fascicle and extreme capsule. J. Gorno-Tempini, M.L., 2010. Neural correlates of syntactic processing in the
Neurosci. 31 (47), 16949–16957. nonfluent variant of primary progressive aphasia. J. Neurosci. 30 (50),
Saffran, J., Hauser, M., Seibel, R., Kapfhamer, J., Tsao, F., Cushman, F., 2008. 16845–16854.
Grammatical pattern learning by human infants and cotton-top tamarin Wilson, S.M., Galantucci, S., Tartaglia, M.C., Rising, K., Patterson, D.K., Henry, M.L.,
monkeys. Cognition 107 (2), 479–500. et al., 2011. Syntactic processing depends on dorsal language tracts. Neuron 72
Sakai, T., Mikami, A., Tomonaga, M., Matsui, M., Suzuki, J., Hamada, Y., et al., 2011. (2), 397–403.
Differential prefrontal white matter development in chimpanzees and Yang, C., 2013. Ontogeny and phylogeny of language. Proc. Natl. Acad. Sci. 110 (16),
humans. Curr. Biol. 21 (16), 1397–1402. 6324–6327.
Sakai, T., Hirata, S., Fuwa, K., Sugama, K., Kusunoki, K., Makishima, H., et al., 2012. Zaccarella, E., Friederici, A.D., 2015a. Syntax in the brain. In: Toga, A. (Ed.), Brain
Fetal brain development in chimpanzees versus humans. Curr. Biol. 22 (18), Mapping: An Encyclopedic Reference. Elsevier, pp. 461–468.
R791–R792. Zaccarella, E., Friederici, A.D., 2015b. Reflections of word processing in the insular
Sakai, T., Matsui, M., Mikami, A., Malkova, L., Hamada, Y., Tomonaga, M., et al., cortex: a sub-regional parcellation based functional assessment. Brain Lang.
2013. Developmental patterns of chimpanzee cerebral tissues provide 142, 1–7.
important clues for understanding the remarkable enlargement of the human Zaccarella, E., Friederici, A.D., 2015c. Merge processing in the human brain: a
brain. Proc. R. Soc. Lond. B Biol. 280 (1753), 20122398. sub-region based functional investigation in the left pars opercularis. Front.
Sanides, F., 1962. Entwicklungsprinzipien des menschlichen Stirnhirns. Psychol. 6, 1818.
Naturwissenschaften 49 (7), 160–161. Zaccarella, E., Meyer, L., Makuuchi, M., Friederici, A.D., 2015. Building by syntax:
Santi, A., Grodzinsky, Y., 2010. fMRI adaptation dissociates syntactic complexity the neural basis of minimal linguistic structures. Cereb. Cortex, Bhv234.
dimensions. Neuroimage 51 (4), 1285–1293. Zaccarella, E., Friederici, A.D., 2016. 2016. The neurobiological nature of syntactic
hierarchies. Neurosci. Biobehav. Rev.

You might also like