1 Introduction

As explained in a recent critical review [1], generally accepted fundamentals and methodologies of traffic and transportation theory have failed by their applications for traffic network optimization and control in the real world. In comparison with the above-mentioned review with about 540 references, this brief critical review is mostly devoted to formulation and critical discussion of empirical and theoretical fundamentals, which can be used for the development of reliable automatic driving, control and optimization of vehicular traffic and transportation networks.

Therefore, objectives of this critical review are as follow:

  1. (i)

    We show why the fundamental empirical basis for automatic driving, reliable control and optimization of traffic and transportation networks is the set of empirical features of traffic breakdown at a road bottleneck.

  2. (ii)

    We explain why the classical generally accepted traffic flow models and associated well-known traffic simulation tools cannot be used for analyses of automatic driving, reliable control and optimization of vehicular traffic and transportation networks.

  3. (iii)

    We present results of three-phase traffic theory that can show these empirical features of traffic breakdown. Therefore, three-phase traffic flow models can be used as reliable tools for the development of automatic driving vehicles that improve traffic safety and decrease the probability of traffic congestion in traffic and transportation networks.

2 Generally accepted fundamentals and methodologies of traffic and transportation theory

Traffic researchers have developed a huge number of traffic theories for optimization and control of traffic and transportation networks. In particular, generally accepted fundamentals and methodologies of traffic and transportation theory are as follow:

  1. i.

    The Lighthill-Whitham-Richards (LWR) model introduced in 1955–1956 [2, 3]. Daganzo introduced a cell-transmission model (CTM) that is consistent with the LWR model [4, 5]. Currently, Daganzo’s CTM is widely used for simulations of traffic and transportation networks (see references in [1]).

  2. ii.

    A traffic flow instability that causes a growing wave of a local reduction of the vehicle speed. This classical traffic flow instability was introduced in 1959–1961 in the General Motors (GM) car-following model by Herman, Gazis, Montroll, Potts, Rothery, and Chandler [68] (see also other references in the book [9]).

    With the use of very different mathematical approaches, this classical traffic flow instability [68] has been incorporated in a huge number of traffic flow models that can be considered belonging to the GM model class. This is because (as found firstly by Kerner and Konhäuser in [10, 11]) in all these very different traffic flow models the traffic flow instability that causes a growing wave of a local reduction of the vehicle speed leads to a moving jam (J) formation in free flow (F) (called \(\mathrm{F} \to \mathrm{J}\) transition; see Fig. 1).

    Fig. 1.
    figure 1

    Simulations of traffic breakdown at on-ramp bottleneck in the GM model class: Speed in space and time during the spontaneous emergence of wide moving jam (J) in initial free flow (F) at on-ramp bottleneck (\(\mathrm{F} \to \mathrm{J}\) transition). The flow rate on the main road upstream of the bottleneck \(q_{\mathrm{in}}\) is larger in (a) than that in (b), whereas on-ramp inflow \(q_{\mathrm{on}}\) is smaller in (a) than that in (b) at the same other model parameters (see Sect. 10.3 of the book [32] for more detail)

    Examples of the well-known models that belong to the GM model class are as follows: Gipps’s model, Payne’s model, Newell’s optimal velocity (OV) model, Wiedemann’s model (used in VISSIM traffic simulation tool), Whitham’s model, the Nagel-Schreckenberg (NaSch) cellular automaton (CA) model, Bando et al. OV model, Treiber’s IDM, Krauß model (used in SUMO tool), the Aw-Rascle model and many other well-known microscopic and macroscopic traffic-flow models. These models are the basis of a number of traffic simulation tools widely used by traffic engineers and researchers (see e.g., references in reviews [9, 1220]).

  3. iii.

    The understanding of highway capacity as a particular value. This understanding of road capacity was probably introduced in 1920–1935 (see the classical paper by Greenshields [21] and references in [12, 15, 16, 20]). Recently, due to empirical results of [2229] it has been assumed this the particular highway capacity is a stochastic value (see Appendix A).

  4. iv.

    Wardrop’s user equilibrium (UE) and system optimum (SO) principles for traffic and transportation network optimization and control introduced in 1952 [30]. The Wardrop’s UE and SO principles are the basis for a huge number of models for dynamic traffic assignment, control and optimization of traffic and transportation networks (see references in [1, 31]).

3 The fundamental empirical basis for automatic driving as well as for reliable control and optimization of vehicular traffic and transportation networks

Vehicular traffic is a spatiotemporal process because it occurs in space and time. Traffic and transportation networks are usually very complex. Therefore, it is not surprising that in empirical studies of traffic data measured in vehicular traffic a diverse variety of empirical spatiotemporal traffic phenomena have been discovered. Obviously, each of the traffic and transportation theories and models can explain some real traffic phenomena and each of the models exhibits a limited region of the applicability for the explanation of real traffic and transportation phenomena. Therefore, the following question arises:

  • Whether is there an empirical traffic phenomenon that can be considered the fundamental empirical basis of transportation science?

Users of traffic and transportation networks would expect that through the use of traffic control, dynamic traffic assignment and other methods of dynamic optimization traffic breakdown can be prevented, i.e., free flow can be maintained in the network. This is because due to traffic breakdown congested traffic occurs in which travel time, fuel consumption as well as other travel costs increased considerably in comparison with travel costs in free flow.

Therefore, any traffic and transportation theory, which is claimed to be a basis for the development of reliable methods and strategies for dynamic traffic assignment as well as network optimization and control should be consistent with the fundamental empirical features of traffic breakdown at a road bottleneck.

The fundamental empirical basis for automatic driving as well as for reliable control and optimization of vehicular traffic and transportation networks is the set of empirical features of traffic breakdown at a road bottleneck. Consequently, we can also make the following conclusion:

  • Traffic and transportation theories, which are not consistent the set of the fundamental empirical features of traffic breakdown at a bottleneck, cannot be applied for the development of reliable management, control, and organization of traffic and transportation networks.

4 The set of fundamental empirical features of traffic breakdown at highway bottlenecks

The set of fundamental empirical features of traffic breakdown at a highway bottleneck, which is found from a study of traffic breakdown at road bottlenecks during many different days (and years) of traffic breakdown observations, is as follows [32, 33]:

  1. 1.

    Traffic breakdown at a highway bottleneck is a local phase transition from free flow (F) to congested traffic whose downstream front is usually fixed at the bottleneck location (see, e.g., [12, 15, 20, 2229] and references there). In three-phase traffic theory, such congested traffic is called synchronized flow (S) [32, 33] (Fig. 2).

    Fig. 2.
    figure 2

    Examples for empirical spontaneous (a) and empirical induced traffic breakdown (b) at highway bottlenecks (1 min averaged field data measured by road detectors on freeway A5-South in Germany). Taken from [33]. This empirical data explains the qualitative difference between two phases in congested traffic—wide moving jam and synchronized flow as follows. The downstream front of the wide moving jam (J) in (b) propagates upstream of the bottleneck with a constant mean velocity. In contrast, the downstream front of synchronized flow is fixed at the bottleneck. This different behavior of the downstream front of congested traffic is the basis for the definition of the wide moving jam and synchronized flow phases in congested traffic (more detailed explanations as well as the motivation of these definitions see in Sects. 2.3–2.7 of the book [32]; the concept of synchronized flow has also been explained in Sect. 4.2 of the book [33])

  2. 2.

    At the same bottleneck, traffic breakdown can be either spontaneous (Fig. 2(a)) or induced (Fig. 2(b)) [32, 33] (see Appendix B).

  3. 3.

    As found firstly by Elefteriadou et al. in 1995 [22], traffic breakdown exhibits a probabilistic nature: At the same bottleneck, traffic breakdown is observed on different days at very different flow rates. In 1998, Persaud et al. [25] found that probability of traffic breakdown is an increasing flow rate function. This result has been confirmed in empirical studies of data measured in different countries [2529, 34].

  4. 4.

    There is a well-known hysteresis phenomenon associated with traffic breakdown and a return transition to free flow (e.g., [12, 20, 2329]).

5 Explanation of failure of classical traffic and transportation theories

As emphasized in [1], there are many achievements of the generally accepted fundamentals and methodologies of traffic and transportation theory (Sect. 2), which have made a great impact on the understanding of many traffic phenomena. Because of these achievements of generally accepted classical traffic and transportation theories, a question arises:

  • Why does the author state in [1] that the generally accepted classical traffic and transportation theories are not consistent with the set of empirical features of traffic breakdown and, therefore, they are not applicable for a reliable description of traffic breakdown, capacity, effect of automatic driving on real traffic flow, traffic control, and optimization of real traffic and transportation networks?

The failure of the generally accepted classical traffic flow theories is explained as follows [1]:

  1. 1.

    The LWR-theory [25] failed because this theory cannot show induced traffic breakdown observed in real traffic (Fig. 2(b)) (see the proof of this statement in Sect. 4 of [1]).

  2. 2.

    Traffic flow models of the GM model class (see references in [9, 1214, 1720]) failed because traffic breakdown in the models of the GM class is an \(\mathrm{F} \to \mathrm{J}\) transition (Fig. 1): Due to traffic breakdown, a wide moving jam(s) appears spontaneously in an initially free flow at a bottleneck (Fig. 1). In contrast with this model result, real traffic breakdown is a phase transition from free flow (F) to synchronized flow (S) (\(\mathrm{F} \to \mathrm{S}\) transition): Rather than a wide moving jam(s), due to traffic breakdown in real traffic, synchronized flow occurs whose downstream front is fixed at the bottleneck (Fig. 2) (a more detailed explanation of the definitions and features of the \(\mathrm{F} \to \mathrm{S}\) and \(\mathrm{F} \to \mathrm{J}\) transitions can be found, respectively, in Chap. 3 and Sect. 10.3 of the book [32]).

  3. 3.

    The understanding of highway capacity as a particular value [12, 15, 16, 2129] failed because this assumption about the nature of highway capacity contradicts the empirical evidence that traffic breakdown can be induced at a highway bottleneck (Fig. 2(b)) (see Appendixes AD).

  4. 4.

    Dynamic traffic assignment or/and any kind of traffic optimization and control based on Wardrop’s SO or UE principles (see references in [1, 31]) failed because of possible random transitions between the free flow and synchronized flow at highway bottlenecks. Due to such random transitions, the minimization of travel cost in a traffic network is not possible.

This can explain why network optimization and control approaches based on these fundamentals and methodologies failed by their applications in the real world. Even several decades of a very intensive effort to improve and validate network optimization models have no success. Indeed, there can be found no examples where on-line implementations of the network optimization models based on these fundamentals and methodologies could reduce congestion in real traffic and transportation networks.

This is due to the fact that the fundamental empirical features of traffic breakdown at highway bottlenecks have been understood only during last 20 years. In contrast, the generally accepted fundamentals and methodologies of traffic and transportation theory have been introduced in the 50–60 s.

Thus the scientists whose ideas led to these classical fundamentals and methodologies of traffic and transportation theory (see Sect. 2 above) could not know the set of empirical features of real traffic breakdown. It should be noted that many of the diverse driver behavioral characteristics related to real traffic as well as some of the mathematical approaches to traffic flow modeling, which have been discovered in classical approaches to traffic flow theory, are also used in three-phase traffic theory and associated microscopic traffic flow models (for more details, see Sect. 11 of [1]).

6 Basic theoretical fundament for the development of reliable control and optimization of traffic and transportation networks

To explain the set of the fundamental empirical features of traffic breakdown at network bottlenecks, the author has introduced three-phase traffic flow theory [32, 33, 3537].

  • The main reason for the three-phase traffic theory is the explanation of the set of the fundamental empirical features of traffic breakdown at highway bottlenecks.

In three-phase traffic theory, an \(\mathrm{F} \to \mathrm{S}\) transition explains traffic breakdown at a highway bottleneck: The terms “\(\mathrm{F} \to \mathrm{S}\) transition” and “traffic breakdown” are synonyms. The \(\mathrm{F} \to \mathrm{S}\) transition (traffic breakdown) occurs in metastable free flow (Figs. 3 and 4) [33] (see explanation of the term “metastable free flow with respect to an \(\mathrm{F} \to \mathrm{S}\) transition” in Appendixes BD). The metastability of free flow explains both spontaneous (Fig. 3(a)) and induced (Fig. 3(b)) traffic breakdowns leading to the emergence of synchronized flow at the bottleneck (empirical features 1 and 2 of traffic breakdown of Sect. 4).

Fig. 3.
figure 3

Simulations of spontaneous (a) and induced (b) breakdown at on-ramp bottleneck with a microscopic stochastic traffic flow model in the framework of three-phase traffic theory [38]

Fig. 4.
figure 4

Explanations of the fundamental empirical features of traffic breakdown at on-ramp bottleneck based on the assumption of three-phase traffic theory about the metastability of free flow at the bottleneck with respect to an \(\mathrm{F} \to \mathrm{S}\) transition (taken from [1]): (a) Simulations of the probability of spontaneous traffic breakdown at the bottleneck on a single-lane road taken from [39]. (b) Qualitative Z-speed-flow-rate characteristic for traffic breakdown; F—free flow, S—synchronized flow. Figure (b) has been qualitatively drawn in accordance with simulations of the Z-speed-flow-rate characteristic for traffic breakdown at on-ramp bottleneck shown in Fig. 3.17(b) of the book [32] in which the flow rate in free flow downstream of the bottleneck \(q_{\mathrm{sum}} = q_{\mathrm{in}} + q_{\mathrm{on}}\) changes through a change in the on-ramp inflow \(q_{\mathrm{on}}\) at a given flow rate in free flow upstream of the bottleneck \(q_{\mathrm{in}}\); in these simulations, due to the increase in \(q_{\mathrm{on}}\) at constant \(q_{\mathrm{in}}\), average synchronized flow speed decreases slightly (in Fig. 3.17(b) of the book [32], the synchronized flow speed is related to a virtual road detector located 200 m upstream of the on-ramp merging region)

The theoretical probability of spontaneous traffic breakdown at the bottleneck found firstly from simulations of a microscopic stochastic three-phase traffic flow model [39] (Fig. 4(a)) is a growing flow-rate function as discovered by Persaud et al. [25] in field data measured by road detectors (empirical feature 3 of traffic breakdown of Sect. 4). The theoretical dependence of the probability \(P^{(B)}(q_{\mathrm{sum}})\) of spontaneous traffic breakdown at the on-ramp bottleneck on the flow rate \(q_{\mathrm{sum}}\) in free flow at the bottleneck is well fitted by a function [39]

$$ P^{(B)}(q_{\mathrm{sum}}) = \frac{1}{1 + \exp [\alpha \left( q_{P} - q_{\mathrm{sum}} \right)]}, $$
(1)

where \(q_{\mathrm{sum}} = q_{\mathrm{in}} + q_{\mathrm{on}}\) is the flow rate downstream of the bottleneck, parameters \(\alpha\) and \(q_{P}\) depend on the on-ramp inflow rate \(q_{\mathrm{on}}\) and a time interval within which traffic breakdown is studied, \(q_{\mathrm{in}}\) is the flow rate in free flow on the main road upstream of the bottleneck. In other words, in accordance with empirical data, in simulations has been found that the on-ramp inflow rate \(q_{\mathrm{on}}\) and the flow rate in free flow on the main road upstream of the bottleneck \(q_{\mathrm{in}}\) have different contributions to traffic breakdown (see Fig. 18(b, c) of [39]). Qualitatively the same growing flow-rate function for the breakdown probability (1) has later been found in field data measured by road detectors in [2729].

The possibility of empirical induced traffic breakdown at a bottleneck (Fig. 2(b)) leads to the following conclusion of three-phase traffic theory: At the same flow rate on a network link, traffic flow at the bottleneck can be either in the free flow phase (F) or in the synchronized flow phase (S) (Fig. 4(b)). In three-phase theory, this empirical fact is responsible for the existence at any time instant of the range of the infinite number of highway capacities, which are within the flow-rate range between a minimum highway capacity \(C_{\min}\) and a maximum highway capacity \(C_{\max}\) (Fig. 4); within this flow-rate range, traffic breakdown can be induced at the bottleneck. Thus, the theoretical fundament resulting from three-phase traffic theory for the development of reliable control and optimization of traffic and transportation networks is as follows.

  • At any time instant, there are the infinite number of the flow rates in free flow at a bottleneck at which traffic breakdown can be induced at the bottleneck. These flow rates are the infinite number of the capacities of free flow at the bottleneck. The range of these capacities of free flow at the bottleneck is limited by the minimum highway capacity \(C_{\min}\) and the maximum highway capacity \(C_{\max}\) (Fig. 4).

The sense of the infinite number of highway capacities of free flow at the bottleneck is as follows (see also Appendix C). When the flow rate in free flow at the bottleneck is within the flow rate range

$$ C_{\min} \le q_{\mathrm{sum}} < C_{\max}, $$
(2)

free flow is in a metastable state with respect to traffic breakdown (\(\mathrm{F}\to\mathrm{S}\) transition) at the bottleneck. This means that traffic breakdown can occur. Therefore, all flow rates satisfying conditions (2) are highway capacities.

In general, there can be four ranges I, II, III, and IV of the flow rate (Fig. 4) within which free flow at a road bottleneck exhibits qualitatively different features with respect to traffic breakdown [32, 33].

In the range I of the flow rate related to condition

$$ q_{\mathrm{sum}} < C_{\min}, $$
(3)

free flow is stable with respect to traffic breakdown (Fig. 4). This means that highway capacity does not reach.

In the range II of the flow rate related to condition

$$ C_{\min} \le q_{\mathrm{sum}} < q_{\mathrm{th}}^{(B)}, $$
(4)

free flow is metastable with respect to traffic breakdown; however, the breakdown can be induced only. This is because under condition (4) the probability of spontaneous breakdown during a given time interval is equal to zero: \(P^{(B)} = 0\). In conditions (4), \(q_{\mathrm{sum}} = q_{\mathrm{th}}^{(B)}\) (Fig. 4(a)) is a threshold flow rate for spontaneous traffic breakdown at which the breakdown probability \(P^{(B)}(q_{\mathrm{th}}^{(B)})\) is very small but it is still larger than zero.

In the range III of the flow rate related to condition

$$ q_{\mathrm{th}}^{(B)} \le q_{\mathrm{sum}} < C_{\max}, $$
(5)

free flow is also metastable with respect to traffic breakdown. However, in contrast with range II (conditions (4)) traffic breakdown can occur spontaneously at the bottleneck during a given time interval with the probability \(0 < P^{(B)}(q_{\mathrm{sum}}) < 1\). Naturally, under conditions (5) the breakdown can also be induced.

In the range IV of the flow rate related to condition

$$ q_{\mathrm{sum}} \ge C_{\max}, $$
(6)

free flow can be considered unstable with respect to traffic breakdown. This is because the breakdown does occur with probability \(P^{(B)} = 1\) during a time interval that is the shorter, the more the flow rate exceeds the maximum capacity.

Recently, the theoretical conclusion about the existence of the infinite number of road capacities as well as conditions (2)–(6) have been generalized for a city bottleneck due to traffic signal [4042]. In particular, it has been found that the flow rate function of the probability of traffic breakdown (transition from under-saturated to oversaturated traffic) at the signal is also related to that shown in Fig. 4(a) [4042]:

$$ P^{(B)}(\bar{q}_{\mathrm{in}}) = \frac{1}{1 + \exp [\beta \left( q_{s} - \bar{q}_{\mathrm{in}} \right)]}, $$
(7)

where \(\bar{q}_{\mathrm{in}}\) is the average arrival flow rate at the approach, parameters \(\beta\) and \(q_{s}\) depend on signal characteristics.

However, in accordance with three-phase traffic theory of city traffic introduced by the author [4042], synchronized flow patterns occurring in undersaturated traffic flow at traffic signal are quite different from those at highway bottlenecks. As a result, rather than Fig. 4(b) that is valid for highway bottlenecks only, qualitatively different characteristics for a study of the metastability of under-saturated city traffic at the signal should be used.

It must be emphasized that for an empirical study of city traffic phenomena in undersaturated city traffic (metastability, synchronized flow, infinity number of signal capacities) theoretically predicted in [4042], spatiotemporal distributions of the flow rate and speed should be measured both upstream and downstream of the signal over many days in which traffic breakdown (transition from under- to oversaturated traffic) are observed.

Unfortunately, to the knowledge of the author, such field data measured in city traffic is not available. In other words, the set of empirical features of the breakdown at the signal could not still be studied. Therefore, in contrast with highway traffic (Sect. 2 and Appendix B), there is \(\mathit{no}\) empirical proof of the theoretical result about the infinite number of capacities of traffic signal [41, 42]. For this reason, a more detailed consideration of probabilistic theory of city traffic [42] as well as of a comparison of this stochastic theory with classical two-regime models for under- and oversaturated traffic made in [43] are out of scope of this brief review.

It should be noted that a consideration of features of synchronized flow resulting from the breakdown both in highway traffic [32, 33] and in oversaturated city traffic [44, 45] is also out of scope of this review. This is because this review is solely devoted to a brief consideration of the set of fundamental empirical features of traffic breakdown occurring in free flow at highway bottlenecks as well as to the impact of these empirical features of the breakdown on theoretical fundamentals for reliable control and optimization of vehicular traffic and transportation networks.

7 Incommensurability of three-phase traffic theory and classical traffic-flow theories

Due to the criticism of classical traffic-flow theories made in Sect. 4, a question arises:

  • May some of the classical traffic-flow theories be relatively easily adjusted to take into account the empirical evidence of the induced transition from free flow to synchronized flow and the flow-rate dependence of the breakdown probability?

The explanation of traffic breakdown at a highway bottleneck by an \(\mathrm{F} \to \mathrm{S}\) transition in a metastable free flow at the bottleneck is the basic assumption of three-phase traffic theory (Fig. 4) [32, 33, 3537]. None of classical traffic-flow theories (see for review, e.g., [9, 1220]) incorporates an \(\mathrm{F}\rightarrow\mathrm{S}\) transition in a metastable free flow at the bottleneck.

For this reason, the classical traffic-flow models cannot describe the \(\mathrm{F}\rightarrow\mathrm{S}\) phase transition in metastable free flow at highway bottleneck. However, the transition does explain the empirical evidence of the induced transition from free flow to synchronized flow and the flow-rate dependence of the breakdown probability.

In accordance with the classical book by Kuhn [46], this shows the incommensurability of three-phase traffic theory and the classical traffic-flow theories (for more detail, see [47]):

  • The existence in three-phase traffic theory of the minimum highway capacity \(C_{\min}\) at which traffic breakdown (\(\mathrm{F}\rightarrow\mathrm{S}\) phase transition) can still be induced at a highway bottleneck has \(\mathit{no}\) sense for classical traffic flow theories.

The term incommensurability has been introduced by Kuhn in his classical book [46] to explain a paradigm shift in a scientific field.

It must also be noted that the existence of these two phases F and S (Fig. 4) does not result from the stochastic nature of traffic: Even if there were no stochastic processes in vehicular traffic, the states F and S do exist at the same flow rate. For this reason, stochastic approaches to traffic control (see, e.g., [48, 49]), which do not assume a possibility of an \(\mathrm{F}\rightarrow\mathrm{S}\) phase transition in metastable free flow and, respectively, the existence of the flow range between the minimum and maximum capacities of three-phase traffic theory, cannot resolve the above-discussed problem of the inconsistence of classical traffic theories with the set of empirical features of real traffic breakdown.

However, it should be noted that the stochastic nature of traffic influences crucially on the probability of random transitions between the phases F and S. At a given flow rate, this probability can change in several orders of magnitude when stochastic characteristics of traffic change.

8 Future reliable control and optimization of vehicular traffic with the use of three-phase traffic theory

Thus within the flow rate range (5), traffic breakdown at a highway bottleneck can occur with some probability regardless of traffic control. This explains the criticism of generally accepted methods of traffic control of Sect. 5. Therefore, a question arises:

  • How will the three-phase traffic theory assist in providing reliable control and optimization of vehicular traffic?

The three-phase traffic theory provides the following future directions for traffic control and optimization theory:

  1. a)

    the minimization of breakdown probability in free flow at network bottlenecks based on the breakdown minimization principle (BM) principle for the control and optimization of transportation networks [50]. The BM principle should be applied for those parts of a traffic network that are not influenced by congestion together with

  2. b)

    a spatial limitation or/and dissolution of congestion in congested parts of the network.

However, a consideration of the BM principle [50] and methods for the spatial limitation or/and dissolution of congestion at bottlenecks based on three-phase traffic theory [32, 33, 5154] is out of scope of this brief review.

9 Control of congested traffic

Up to now we have discussed empirical traffic breakdown as the empirical fundament for automatic driving as well as for reliable control and optimization of transportation networks. However, as mentioned in Sect. 8, for reliable network optimization and control, one can apply the minimization of the probability of traffic breakdown in the network with the BM principle that should be combined with a spatial limitation of congestion growth and/or congestion dissolution in congested network links.

Moreover, in many real traffic networks there are not enough alternative routes to avoid traffic congestion at large enough traffic demand. In these cases, it can be expected that even if traffic congestion cannot be avoided in some parts of traffic networks, nevertheless, the application of ITS can change characteristics of traffic congestion with the objective to increase traffic safety and comfort while moving in congested traffic. Therefore, the following questions arise:

  1. 1.

    What are empirical features of congested patterns that can be influenced for reliable spatial limitation of congestion growth and/or for congestion dissolution?

  2. 2.

    How can driver behavior change features of congested patterns with the aim to increase safety and comfortable driving?

  3. 3.

    What are vehicle systems for automatic driving as well as other ITS-applications that can help to increase safety and comfort while driving in congested traffic?

The three-phase traffic theory has answered these questions [32, 33]. In particular, in 1998 based on an analysis of real field traffic data measured on German highways the author found out that one of the most important features of spatiotemporal complexity of traffic congestion is a sequence of \(\mathrm{F}\rightarrow\mathrm{S}\rightarrow\mathrm{J}\) phase transitions (Fig. 5) [55] that are as follows. Firstly, an \(\mathrm{F}\rightarrow\mathrm{S}\) transition (traffic breakdown) occurs at a highway bottleneck as discussed in Sect. 6 above. Synchronized flow propagates upstream of the bottleneck. Within the emergent synchronized flow, at some distance upstream of the bottleneck moving jams emerge spontaneously (called as an \(\mathrm{S}\rightarrow\mathrm{J}\) transition) (Fig. 5).

Fig. 5.
figure 5

Empirical sequence of \(\mathrm{F}\rightarrow\mathrm{S}\rightarrow\mathrm{J}\) phase transitions in real traffic [55, 56]: (a) Speed in space and time within a congested pattern occurring at on-ramp bottleneck (1-min averaged data measured on March 23, 1998 on freeway A5-South in Germany with road detectors installed along a three-lane road); data is presented in space and time with averaging method described in Sect. C.2 of [57] (arrows \(\mathrm{F}\rightarrow\mathrm{S}\) and \(\mathrm{S}\rightarrow\mathrm{F}\) mark symbolically, respectively, the \(\mathrm{F}\rightarrow\mathrm{S}\) and \(\mathrm{S}\rightarrow\mathrm{F}\) transitions at the location of on-ramp bottleneck, arrows \(\mathrm{S}\rightarrow\mathrm{J}\) mark symbolically \(\mathrm{S}\rightarrow\mathrm{J}\) transitions related to the emergence of two first wide moving jams within synchronized flow). (b–d) Empirical double Z-characteristics, which are the presentation of phase transitions shown in (a) in the speed–density (b), speed–flow-rate (c) and flow–density planes (d) (points J in (b–d) are related to the second of the wide moving jams in (a))

Both the \(\mathrm{F}\rightarrow\mathrm{S}\) transition and the \(\mathrm{S}\rightarrow\mathrm{J}\) transition, i.e., the sequence of \(\mathrm{F}\rightarrow\mathrm{S}\rightarrow\mathrm{J}\) transitions (Fig. 5(a)), which occur in the reality in space and time, can alternatively be presented either in the speed–density plane (Fig. 5(b)), or in the speed–flow-rate plane (Fig. 5(c)), or else in the flow–density plane (Fig. 5(d)) by double Z-characteristic for phase transitions [33]. In other words, arrows \(\mathrm{F}\rightarrow\mathrm{S}\) and \(\mathrm{S}\rightarrow\mathrm{J}\) shown in these planes (Fig. 5(b–d)) mark, respectively, the presentation of the \(\mathrm{F}\rightarrow\mathrm{S}\) transition and the \(\mathrm{S}\rightarrow\mathrm{J}\) transition associated with real \(\mathrm{F}\rightarrow\mathrm{S}\) transition and the \(\mathrm{S}\rightarrow\mathrm{J}\) transition shown in Fig. 5(a). Obviously, there are also return \(\mathrm{S}\rightarrow\mathrm{F}\) and \(\mathrm{J}\rightarrow\mathrm{S}\) transitions between the three traffic phases (Fig. 5).

10 Effect of automatic driving on traffic flow

Traffic flow simulations should be able to answer the following questions, which arise due to the development of automatic driving vehicles:

  1. 1.

    What is the effect of automatic driving vehicles on traffic flow consisting of usual (non-automatic) vehicles?

  2. 2.

    What are features of vehicle systems for automatic driving that can improve traffic safety and decrease the probability of traffic congestion?

As explained in this review, to answer these questions, traffic flow models used for simulations should be able to explain the set of empirical features of traffic breakdown at a road bottleneck. Because the classical generally accepted traffic flow models cannot show these empirical features, the application of these models and associated simulation tools leads to incorrect conclusions that cannot be used for the development of systems for the future automatic driving vehicles.

For this reason, studies the effect of adaptive vehicle cruise control (ACC) and other vehicle systems for automatic driving on traffic flow with well-known traffic flow models and simulation tools like VISSIM (Wiedemann model), SUMO (Krauß model) as well as all other traffic simulation tools based on classical traffic flow models (see Sect. 2), which have been made and/or reviewed, for example, in [19, 5863], are invalid for the real word. Therefore, simulation approaches of [19, 5863] lead to incorrect conclusions about the impact of ACC vehicles as well as of automatic driving on real traffic flow.

As explained in Sect. 23.6 of the book [33] and Sect. 9.6 of the book [32], to perform a reliable simulations of the effect of ACC and other vehicle systems for automatic driving on traffic flow on traffic flow, a three-phase traffic flow model, which can explain the set of the empirical features of traffic breakdown at highway bottlenecks (Sect. 4), is needed.

One of such three-phase traffic flow models is Kerner-Klenov stochastic microscopic three-phase traffic flow model [6466]. One of the results of simulations of the effect of ACC on traffic flow made with this model is shown in Figs. 6 and 7.

Fig. 6.
figure 6

Simulations of the effect of adaptive vehicle cruise control (ACC) on traffic flow on single-lane road with an on-ramp bottleneck made with Kerner-Klenov microscopic stochastic traffic flow model [6466] in the framework of three-phase traffic theory. Vehicle speed in space and time (a, c, e) and the same speed data presented by regions with variable shades of gray (b, d, f) (in white regions the speed is equal to 105 km/h, in black regions the speed is equal to zero). (a, b) No ACC vehicles. (c, d) 10 % ACC vehicles. (e, f) 20 % ACC vehicles. Explanations of ACC used for simulations is given in Sect. 23.6 of the book [33]; desired time gap \(\tau_{d}^{(\mathrm{ACC})}\) that is a given parameter of ACC is equal to 1.1 s; other parameters of ACC are the same as those in caption of Fig. 23.18 of the book [33]. Arrows \(\mathrm{F}\rightarrow\mathrm{S}\) in (a, b) mark the \(\mathrm{F}\rightarrow\mathrm{S}\) transition (traffic breakdown) at the location of on-ramp bottleneck. Flow rate in free flow upstream of the bottleneck is equal to \(q_{\mathrm{in}} = 2000~\mbox{vehicles/h}\), the flow rate to the on-ramp is equal to \(q_{\mathrm{on}} =320~\mbox{vehicles/h}\). F—free flow, S—synchronized flow, J—wide moving jam

Fig. 7.
figure 7

Simulations of the effect of ACC on traffic flow on single-lane road with an on-ramp bottleneck in the framework with three-phase traffic theory (continue of Fig. 6). Vehicle speed in space and time (a, c) and the same speed data presented by regions with variable shades of gray (b, d) (in white regions the speed is equal to 105 km/h, in black regions the speed is equal to zero). (a, b) 20 % ACC vehicles with a longer time of observation of traffic flow than that in Fig. 6(e, f). (c, d) 30 % of ACC vehicles. Other parameters and explanations are the same as those in Fig. 6

At chosen flow rates \(q_{\mathrm{in}}\) and \(q_{\mathrm{on}}\) (Fig. 6), in traffic flow without ACC vehicles, traffic breakdown (\(\mathrm{F}\rightarrow\mathrm{S}\) transition) occurs at an on-ramp bottleneck after a random time delay for traffic breakdown \(T^{(B)}\) (Fig. 6(a, b)) as this is well-known in three-phase traffic theory and empirical observations [3234].

If there are 10 % ACC vehicles with quick speed adaptation (see explanations of the terms “quick speed adaptation of ACC” and “slow speed adaptation of ACC” in Sect. 23.6 of the book [33]), which are randomly distributed in traffic flow of usual vehicles without ACC, no considerable change occurs in traffic flow: Traffic breakdown occurs after a time delay at the on-ramp bottleneck (Fig. 6(c, d)).

The situation changes qualitatively, if we further increase the percentage of the ACC vehicles: At 20 % of ACC vehicles there is no traffic breakdown during the observation time (Fig. 6(e, f)).

However, if we increase the time of observation of traffic flow, we find that at 20 % of ACC vehicles traffic breakdown can nevertheless occur with a considerably longer mean time delay (Fig. 7(a, b)). If we increase the percentage of the ACC vehicles up to 30 %, no traffic breakdown occurs at the bottleneck any more (Fig. 7(c, d)).

Thus, we can conclude that when the percentage of ACC vehicles with quick speed adaptation increases in traffic flow, firstly the mean time delay in traffic breakdown increases (Fig. 7(a, b)). A further increase in the percentage of the ACC vehicle prevents traffic breakdown at a highway bottleneck (Fig. 7(c, d)).

11 Conclusions

  1. 1.

    The fundamental empirical basis for automatic driving as well as for reliable control and optimization of vehicular traffic and transportation networks is the set of empirical features of traffic breakdown at a road bottleneck.

  2. 2.

    The theoretical fundament resulting from three-phase traffic theory for the development of reliable control and optimization of traffic and transportation networks is the existence of the range of the infinite number of highway capacities: At any time instant, there are the infinite number of highway capacities within a range of the flow rate between the minimum capacity and the maximum capacity (Fig. 4); within this flow range, traffic breakdown can be induced at the bottleneck.

  3. 3.

    The explanation of traffic breakdown at a highway bottleneck by an \(\mathrm{F} \to \mathrm{S}\) transition in a metastable free flow introduced in three-phase traffic theory is responsible for the incommensurability of three-phase traffic theory with all other traffic flow theories.

  4. 4.

    Classical traffic theories failed to explain the set of empirical features of traffic breakdown at a highway bottleneck. For this reasons, traffic flow models, which are based on these classical traffic theories, cannot be used for a reliable analysis of the impact of automatic driving and/or other ITS-applications on traffic flow. Examples of these traffic flow models are the classical LWR-model, Daganzo’s Cell Transmission Model, GM car-following model of Herman, Gazis et al., Newell optimal velocity (OV) model, Gipps model, Bando et al. OV model, Payne’s macroscopic model, Aw-Rasche macroscopic model, Treiber’s IDM, Nagel-Schreckenberg cellular automaton (CA) model as well as well-known and generally used traffic simulation tools (like VISSIM (Wiedemann-model) and SUMO (Krauß-model)): Simulations of the effect of automatic driving and/or other ITS-applications on traffic flow with the use of such traffic simulation models and tools lead to incorrect results and invalid conclusions.

  5. 5.

    To perform reliable analysis of the impact of automatic driving and/or other ITS-applications on traffic flow, traffic flow models in the framework of three-phase traffic theory should be used. This is because these models can explain the set of empirical features of traffic breakdown at a road bottleneck.